Está en la página 1de 426

  

ANUNCIO

Soporte de acceso remoto COVID-19: Obtenga más información sobre el acceso ampliado a
la investigación de Publicaciones ACS.

VOLVER A EDICIÓN

 ANTERIOR REVISIÓN SIGUIENTE 

Máquinas moleculares arti ciales


Sundus Erbas-Cakmak, David A. Leigh*, Charlie T. McTernany Alina L. Nussbaumer

Ver información del autor 

 Cite esto: Chem. Rev. 2015 , 115 , 18 , 10081–10206


Fecha de publicación : 8 de septiembre de 2015 
https://doi.org/10.1021/acs.chemrev.5b00146
Copyright © 2015 Sociedad Química Estadounidense
DERECHOS Y PERMISOS ACS AuthorChoice
con CC-BY licencia

Vistas de artículos Altmétrico Citas

48238 9 916
MÁS INFORMACIÓN SOBRE ESTAS MÉTRICAS

Compartirañadir Exportar

  RIS

 PDF (62 MB)

 Get e-Alerts

ASIGNATURAS: Rotaxanos , Estructura molecular , 


BIOGRAFÍA
  

Sündüs Erbaş Çakmak obtuvo su B.Sc. Licenciado en Biología Molecular y Genética en 2007 de la
Universidad Boğaziçi en Estambul, Turquía. Realizó su Ph.D. Licenciada (2013) en Ciencia de
Materiales y Nanotecnología en el Instituto Nacional de Nanotecnología de la Universidad de Bilkent,
bajo la supervisión del Prof. Engin U. Akkaya, investigando la concatenación y el uso de puertas
lógicas moleculares para la modulación de la actividad de agentes de terapia fotodinámica. Trabajó
bajo la supervisión de Amar H. Flood en la Universidad de Indiana durante 6 meses durante su Ph.D.
trabajando en el desarrollo de sensores de aniones uorescentes. En 2013 se incorporó al grupo del
Prof. David A. Leigh como becaria posdoctoral intraeuropea Marie-Curie. Sus intereses de
investigación incluyen máquinas moleculares arti ciales, catálisis supramolecular y computación
molecular.

BIOGRAFÍA
  

David Leigh nació en Birmingham (Reino Unido) y obtuvo su B.Sc. y Ph.D. grados de la Universidad de
She eld. Después de una investigación postdoctoral en Ottawa (1987-1989), fue designado para una
cátedra en el Instituto de Ciencia y Tecnología de la Universidad de Manchester (Reino Unido).
Después de una temporada en las universidades de Warwick y Edimburgo, regresó a Manchester en
2012, donde actualmente ocupa la cátedra Sir Samuel Hall de Química. Los premios y galardones por
las contribuciones de investigación de su grupo incluyen el Premio Internacional Izatt-Christensen
2007 de Química Macrocíclica, el Premio Feynman 2007 de Nanotecnología, el Premio Descartes de
la UE 2007 para la Investigación, el Premio RSC Merck 2009, el Premio RSC Tilden 2010, la
Conferencia y Premio Panadero de la Royal Society 2013 y el premio RSC Pedler Prize 2014. En 2009
fue elegido miembro de la Beca de la Royal Society (Londres).

BIOGRAFÍA
  

Charlie Thomas McTernan nació en Londres y obtuvo su M.Chem. de la Universidad de Oxford. Su


proyecto de la Parte II se llevó a cabo bajo la supervisión del Prof. Tim Donohoe, investigando la
síntesis de motivos de isoquinolina usando α-arilación catalizada por paladio. Se unió al grupo de
David Leigh en 2013, nanciado por un premio Dean's Faculty Award de la Universidad de Manchester.
Sus intereses de investigación incluyen la síntesis de máquinas moleculares arti ciales, catalizadores
intercambiables y rotaxanos.

BIOGRAFÍA
  

Alina Laura Nussbaumer nació en Berna (Suiza) y obtuvo su B.Sc. y M.Sc. grados de la Universidad de
Berna. Completó su Ph.D. bajo la supervisión del Prof. Robert Häner en la misma Universidad,
investigando el ADN como andamio para el ensamblaje de cromóforos hacia el desarrollo de nuevos
materiales basados en ADN y la producción de polímeros supramoleculares basados en conjugados
de ADN que muestran una fuerte ampli cación de quiralidad. En 2012, se unió al grupo de David Leigh
como becaria postdoctoral de la Swiss National Science Foundation. Sus intereses de investigación
incluyen el movimiento molecular controlado y el desarrollo de nuevas tácticas sintéticas hacia
moléculas topológicamente complejas.

1. Introducción Salta a

El uso generalizado de máquinas moleculares en biología ha sugerido durante mucho


tiempo que se podrían obtener grandes recompensas al cerrar la brecha entre los sistemas
moleculares sintéticos y las máquinas del mundo macroscópico. En las últimas dos
décadas, se ha demostrado que es posible diseñar sistemas moleculares sintéticos con
arquitecturas en las que se producen cambios posicionales de gran amplitud de
componentes submoleculares. Quizás la mejor manera de apreciar el potencial tecnológico
del movimiento controlado a nivel molecular es reconocer que los nanomotores y las
máquinas a nivel molecular se encuentran en el corazón de todo proceso biológico
signi cativo. Durante miles de millones de años de evolución, la naturaleza no ha elegido
repetidamente esta solución para realizar tareas complejas sin una buena razón. Cuando la
humanidad aprenda a construir estructuras arti ciales que puedan controlar y explotar el
movimiento a nivel molecular e interconectar sus efectos directamente con otras  
subestructuras de nivel molecular y el mundo exterior, tendrá un impacto potencial en todos
los aspectos del diseño funcional de moléculas y materiales. Seguramente seguirá una
mejor comprensión de la física y la biología.

Los primeros pasos en el largo camino hacia la invención de las máquinas moleculares
arti ciales se dieron posiblemente en 1827 cuando el botánico escocés Robert Brown
observó el movimiento fortuito de pequeñas partículas bajo su microscopio.(1, 2) La
explicación del movimiento browniano, que es causado por el bombardeo de las partículas
por moléculas como consecuencia de la teoría cinética de la materia, fue proporcionada
más tarde por Einstein, seguida de la veri cación experimental por Perrin.(3, 4)El movimiento
térmico aleatorio de las moléculas y sus implicaciones para las leyes de la termodinámica
inspiraron a su vez Gedankenexperiments (“experimentos mentales”) que exploraron la
interacción (y aparentes paradojas) del movimiento browniano y la Segunda Ley de la
Termodinámica. La famosa conferencia de Richard Feynman de 1959 “ Hay mucho espacio
en la parte inferior ” describió algunas de las promesas que podrían cumplir las máquinas
moleculares arti ciales.(5, 6)Sin embargo, la charla de Feynman se produjo en un momento
antes de que los químicos tuvieran las herramientas sintéticas y analíticas necesarias para
fabricar máquinas moleculares. Si bien el interés entre los químicos sintéticos comenzó a
crecer en las décadas de 1970 y 1980, el progreso se aceleró en la década de 1990,
particularmente con la invención de métodos para hacer sistemas moleculares entrelazados
mecánicamente (catenanos y rotaxanos) y controlar y cambiar las posiciones relativas de
sus componentes.(7-24)

Aquí, revisamos los movimientos desencadenados de gran amplitud en las estructuras


moleculares y los cambios en las propiedades que pueden producir. Nos concentramos en
cambios conformacionales y con guracionales en moléculas totalmente unidas
covalentemente y en catenanos y rotaxanos en los que el cambio es provocado por varios
estímulos (luz, electroquímica, pH, calor, polaridad del solvente, unión catiónica o aniónica,
efectos alostéricos, temperatura, covalente reversible formación de enlaces, etc.).
Finalmente, discutimos las últimas generaciones de so sticados sistemas de máquinas
moleculares sintéticas en las que el movimiento controlado de subcomponentes se utiliza
para realizar tareas complejas, allanando el camino a las aplicaciones y la realización de una
nueva era de “nanotecnología molecular”.

1.1 El lenguaje utilizado para describir las máquinas


moleculares
La terminología debe de nirse de manera adecuada y apropiada y estos signi cados deben
usarse de manera consistente para transmitir conceptos cientí cos de manera efectiva. En
ninguna parte es más evidente la necesidad de un lenguaje cientí co preciso que en el
campo de las máquinas moleculares. Gran parte de la terminología utilizada para describir
las máquinas de nivel molecular tiene su origen en observaciones realizadas por biólogos y
físicos, y sus descubrimientos y descripciones a menudo han sido mal interpretados y mal 
comprendidos por los químicos. En 2007 formalizamos de niciones de algunos términos
comunes utilizados en el campo (por ejemplo, "máquina", "interruptor", "motor", "trinquete",
etc.) para que los químicos pudieran usarlos de una manera consistente con los signi cados
entendidos por biólogos y físicos que estudian máquinas de nivel molecular.(14)

La palabra "máquina" implica un movimiento mecánico que realiza una tarea útil. Esta
revisión se concentra en sistemas en los que un estímulo desencadena el movimiento
controlado, de amplitud relativamente grande (o direccional) de un componente molecular o
submolecular en relación con otro que potencialmente puede resultar en la realización de
una tarea neta. Las máquinas moleculares se pueden clasi car en varias clases, como
"motores" e "interruptores", cuyo comportamiento di ere signi cativamente.(14)Por ejemplo,
en un “conmutador” basado en rotaxano, el cambio de posición de un macrociclo en el hilo
del rotaxano in uye en el sistema solo en función del estado. Devolver los componentes de
un interruptor molecular a su posición original deshace cualquier trabajo realizado y, por lo
tanto, un interruptor no se puede usar de manera repetitiva y progresiva para realizar el
trabajo. Un "motor", en cambio, in uye en un sistema en función de la trayectoria, es decir,
cuando los componentes de un motor molecular vuelven a sus posiciones originales, por
ejemplo, después de una rotación direccional de 360 °, cualquier trabajo que se haya
realizado no se deshace a menos que el motor se gire posteriormente 360 ° en la dirección
inversa. Esta diferencia de comportamiento es signi cativa; No se puede utilizar ninguna
máquina molecular "basada en interruptores" para realizar el trabajo progresivamente de la
forma en que lo hacen los motores biológicos, como los de la kinesina,(14)

1.2 Los efectos de la escala


Las máquinas deben diseñarse de acuerdo con el entorno en el que deben funcionar. El
signi cado y las consecuencias del movimiento térmico aleatorio, la disipación de calor, la
solvatación, el momento, la inercia, la gravedad, etc., di eren signi cativamente a nivel
molecular y macroscópico, lo que signi ca que las máquinas a nanoescala no pueden
simplemente imitar los mecanismos de sus hermanos macroscópicos.

Las fuerzas con mayor in uencia en la dinámica del nano-mundo no son aquellas en las que
comúnmente con amos en el mundo macroscópico. Para los objetos grandes, los términos
inerciales, que dependen de la masa de la partícula, dominan el movimiento. A medida que el
tamaño de las partículas disminuye hasta o por debajo de la escala micrométrica, las
fuerzas viscosas y el movimiento browniano se vuelven dominantes, mientras que el
impulso y la gravedad se vuelven cada vez más irrelevantes. Este efecto se puede cuanti car
mediante el número de Reynolds ( R ) para una partícula de radio a , velocidad v en un medio
de densidad ρ y viscosidad μ:(25, 26)

(1)
Debido a que el número de Reynolds disminuye con el radio, las máquinas de tamaño
molecular normalmente operan en condiciones de bajo número de Reynolds. Además,
  la 
mayor relación super cie: volumen de las partículas pequeñas las hace cada vez más
adhesivas, disminuyendo la movilidad. Estos efectos deben considerarse cuidadosamente
en el diseño de máquinas moleculares.(14)

1.3 Máquinas de pensamiento que exploran el movimiento


browniano
Los motores moleculares se basan en la explotación de uctuaciones térmicas aleatorias
para el movimiento direccional mediante el empleo de mecanismos de trinquete.
Experimentos mentales como el demonio de Maxwell,(27, 28) La trampilla de Smoluchowski,
(29) y el trinquete y el trinquete de Feynman(30) han investigado posibles formas de
provocar el movimiento direccional de las partículas brownianas.(31)

La Segunda Ley de la Termodinámica establece que la entropía de un sistema aislado tiende


a aumentar, lo que lleva a una distribución de equilibrio con máxima entropía. Para lograr
cualquier distribución que no sea el equilibrio termodinámico, se debe trabajar en el sistema.
Se han propuesto varios experimentos mentales que intentan violar esta premisa y alejar a
un sistema del equilibrio sin gastar trabajo. En el experimento mental del demonio de
Maxwell, las partículas de un recipiente son clasi cadas por un "guardián inteligente", el
"demonio" ( Figura 1 ).(32)Las partículas de gas se distribuyen en un recipiente dividido en
dos secciones en un sistema aislado. La formación espontánea de un gradiente de calor o
presión conduciría a una disminución de la entropía y, por tanto, violaría la Segunda Ley. El
guardián puede detectar la velocidad de cada partícula y puede controlar la puerta en
consecuencia. El demonio permite que partículas con velocidades superiores a la media
(mostradas en rojo en la Figura 1 a) pasen de la sección derecha a la izquierda, pero no de la
izquierda a la derecha. Lo contrario es válido para partículas más lentas (que se muestran en
azul en la Figura 1un). El portero provoca así una distribución no uniforme de partículas
entre las dos secciones, creando un gradiente de temperatura. De manera similar, un
demonio capaz de detectar la dirección de la partícula y abrir la puerta en consecuencia,
puede concentrar partículas en una sección y generar así un gradiente de presión ( Figura 1
b). Si la puerta no tiene fricción, entonces en ambos casos se ha logrado una distribución de
desequilibrio aparentemente sin hacer trabajo, lo que violaría la Segunda Ley. La solución a
la aparente paradoja radica en considerar la información que requiere el demonio para saber
cuándo abrir la puerta. Como el demonio debe medir la velocidad o dirección de cada
partícula que se acerca a la puerta(33) y no puede tener memoria in nita,(34)en algún
momento el demonio debe "olvidar" esta información. La destrucción de información tiene
un costo entálpico mínimo asociado(35) como se ha veri cado experimentalmente,(36)que
siempre supera la disminución de entropía en el contenedor. Por lo tanto, una disminución
local en la entropía se paga mediante un aumento generalizado al eliminar la información.

Figura 1
  

Figura 1. Demonios de Maxwell. (a) Un "demonio de la temperatura" clasi ca las partículas


según la velocidad, generando un gradiente de temperatura. (b) Un “demonio de la presión”
clasi ca las partículas de acuerdo con su dirección de movimiento, generando un gradiente
de presión.(37)

Smoluchowski propuso un sistema que no dependía de un observador inteligente basado en


una puerta unidireccional cargada por resorte, abierta por uctuaciones térmicas ( Figura 2
a).(29, 38, 39)Sin embargo, se demostró que el transporte direccional es imposible en este
sistema, ya que la puerta no puede disipar la energía de las colisiones con las partículas de
gas. Esto conduce a períodos abiertos cada vez más largos para la compuerta y, por lo
tanto, al equilibrio de las partículas a través del contenedor.

Figura 2

Figura 2. (a) En la Trampilla de Smoluchowski, una puerta con resorte separa las dos
secciones de un contenedor. La falta de disipación de energía conduce a una mayor
duración de la apertura y, por lo tanto, a una distribución uniforme de partículas entre las
secciones.(29)(b) Dispositivo de trinquete y trinquete de Feynman. La asimetría de los
dientes del engranaje de trinquete estaba destinada a impulsar el movimiento direccional.
Sin embargo, cuando T 1 = T 2 , la rotación no es direccional.(30)
En el trinquete y el trinquete de Feynman, los dientes dentada asimétricos podrían haber
permitido la rotación direccional impulsada por el movimiento browniano ( Figura 2 b). 
(30, 40-44)Aunque de acuerdo con la segunda ley, no se pudo extraer trabajo cuando T 1 = T
2 (el trinquete está a la misma temperatura que el resto del dispositivo y, por lo tanto, uctúa
entre los estados abierto y cerrado y, por lo tanto, no actúa como un trinquete) , pero se
puede trabajar cuando T 1 > T 2 . Este es un punto importante para el diseño de máquinas a
nanoescala. A medida que el sistema pasa de un estado de no equilibrio al equilibrio ( T 1 = T
2), se puede realizar trabajo direccional. Por lo tanto, si se puede crear una situación de
desequilibrio, su relajación puede impulsar el movimiento direccional. En otras palabras, se
requiere una entrada de energía para realizar el trabajo y romper el equilibrio detallado (ver
sección 4.4 ).

1.4 Inspiración de la naturaleza


Quizás la lección más importante para aprender de los sistemas biológicos es que las
máquinas moleculares son viables y pueden realizar tareas extremadamente complejas. En
la naturaleza, las máquinas moleculares desempeñan un papel vital, ya que participan en
casi todos los procesos biológicos importantes y permiten realizar una amplia gama de
tareas químicas y mecánicas.(45-47) Las formas en que la naturaleza ha superado los
problemas de escala, movimiento browniano, viscosidad, distribución de desequilibrio y
medio ambiente proporcionan algunas direcciones generales para el diseño de máquinas
moleculares.

Los procesos celulares están compartimentados por barreras lipídicas, que permiten que se
acumulen y mantengan distribuciones en desequilibrio. Una gama de máquinas moleculares
controlan el transporte de diversas especies cargadas y polares a través de estas
membranas mediante canales de retransmisión o entidades portadoras móviles.(48-50)Esta
transferencia puede ser impulsada por difusión pasiva hacia abajo de un gradiente de
concentración o por la generación de un gradiente electroquímico (donde una especie se
mueve en contra de su gradiente de concentración pero hacia abajo por otro gradiente,
como el potencial eléctrico). Si el movimiento es impulsado por el transporte acoplado de
otro sistema, las dos especies pueden migrar en la misma (simportación) o en direcciones
opuestas (antipuerto) a través de la membrana. El transporte altamente selectivo a través de
las bicapas lipídicas es posible gracias al gran número de canales y portadores especí cos
en los sistemas biológicos impulsados por la relajación de los gradientes electroquímicos
transmembrana hacia el equilibrio termodinámico.

Estos gradientes electroquímicos son mantenidos por solo unas pocas bombas de iones,
más comúnmente ATPasas que hidrolizan ATP y generan transporte direccional utilizando la
liberación de energía resultante. Aunque el mecanismo del movimiento direccional en estas
proteínas motoras aún no se comprende completamente, los cambios conformacionales
que conducen a a nidades de unión alteradas parecen ser particularmente importantes.
(51-60) El acceso a los sitios de unión de estas bombas a menudo se "cierra" por cambios
conformacionales después de la unión de iones para aumentar la direccionalidad del
proceso.(58, 61)   
El bombeo de protones a través de las membranas es un proceso particularmente
importante, ya que crea la fuerza motriz del protón que impulsa la producción de ATP de las
ATP sintasas, que a su vez impulsa una gran variedad de procesos de transporte activo. Este
potencial transmembrana generado por protones es comúnmente creado por cadenas de
transporte de electrones donde se utilizan reacciones redox para separar cargas a través de
una membrana.(60)El bombeo de protones impulsado directamente por la luz, como en la
bacteriorrodopisina (donde la isomerización cis / trans conduce a la direccionalidad), es
menos común.(62-67) Las estructuras entrelazadas a menudo se explotan en biología para
ayudar a garantizar una alta secuencialidad, por ejemplo, en el ribosoma,(68-70) y diversas
ADN polimerasas.(71)

Las máquinas moleculares biológicas se utilizan para transportar carga alrededor de una
célula (p. Ej., Kinesina), para impulsar el movimiento de organismos (p. Ej., Motores
agelares bacterianos), para sintetizar proteínas (p. Ej., El ribosoma) y para separar hebras
de ADN (p. Ej., helicasas). Hay muchas diferencias cruciales entre estas máquinas y las que
conocemos en el mundo macroscópico, tanto en términos de las tareas que realizan como
de la forma en que lo hacen. A partir de estas máquinas naturales se pueden extrapolar
varios conceptos pertinentes a la síntesis de máquinas moleculares arti ciales.

(1) Las máquinas biológicas no pueden utilizar gradientes térmicos debido a la


rápida disipación de calor en la escala en la que operan.

(2) Los motores biológicos a menudo usan energía química de eventos favorables
de formación / ruptura de enlaces (por ejemplo, hidrólisis de ATP) o gradientes
de concentración para impulsar su funcionamiento.

(3) Las biomáquinas operan en solución y sobre super cies bajo condiciones de
alta viscosidad.

(4) Las máquinas biológicas explotan el movimiento browniano en lugar de luchar


contra él. Ellos "trincan" el movimiento térmico aleatorio de sus componentes y
sustratos. El movimiento browniano también asegura la mezcla rápida de
máquinas, combustible y sustratos.

(5) La fricción es irrelevante en un medio tan viscoso, donde el movimiento


browniano constante asegura la movilidad de las partes individuales.

(6) Las biomáquinas móviles, como la kinesina, a menudo operan a lo largo de las
vías para reducir los grados de libertad disponibles de la máquina. La
asociación cinética de la máquina en la vía durante la duración de la operación
  
es vital para lograr un transporte dirigido.

(7) Las interacciones no covalentes son a menudo características importantes de


la estructura y el funcionamiento de los sistemas biológicos.

(8) Las biomáquinas se fabrican a menudo mediante procesos de autoensamblaje


a partir de una gama limitada de motivos básicos como aminoácidos, lípidos,
ácidos nucleicos y sacáridos.

(9) A menudo se requiere la compartimentación para permitir que los sistemas se


mantengan y operen fuera del equilibrio.

(10) Los pequeños eventos de unión a menudo pueden causar grandes cambios
conformacionales.

Aunque las máquinas moleculares biológicas tienen un nivel de complejidad inalcanzable


por la generación actual de máquinas moleculares, están constreñidas por los procesos
evolutivos que las dieron origen y el entorno en el que operan. La selección natural asegura
que la primera solución exitosa a un problema tiende a retenerse y mejorarse gradualmente
a lo largo de muchas iteraciones. Estas restricciones, en términos de química y soluciones,
no se aplican a los sistemas arti ciales. Las biomáquinas también operan en un entorno
muy desordenado y, por lo tanto, deben exhibir una tolerancia extremadamente alta a las
colisiones con especies no relacionadas o no reactivas. Además, el estudio de sistemas
arti ciales menos elaborados puede dar lugar a una mayor comprensión de los so sticados
sistemas biológicos que los inspiraron.

1.5 Transporte direccional y trabajo bajo movimiento


browniano
El principio del equilibrio detallado proporciona importantes restricciones sobre la relación
entre el equilibrio y las constantes de velocidad. La comprensión de este concepto es
importante en el campo de las máquinas moleculares sintéticas y para el diseño de posibles
mecanismos de movimiento direccional.(72-74)El Principio del Balance Detallado establece
que, en el equilibrio, cada transición tiene la misma probabilidad (y por lo tanto la tasa) de
ocurrir en una dirección hacia adelante o hacia atrás (por ejemplo, k 1 = k –1 ). Esto signi ca
que en el equilibrio no se genera ningún ujo neto a través de ninguna barrera y no se puede
realizar ninguna tarea. Esto descarta el mantenimiento de un equilibrio mediante un proceso
cíclico como A → B → C → A; más bien, cada sustrato debe estar en equilibrio con todos los
demás con los que intercambia, es decir, A⇆ B + B ⇆ C + C ⇆ A en su lugar. De ello se
deduce que la trampilla de Smoluchowski no podría funcionar, ya que rompería el principio
de reversibilidad microscópica. Romper el equilibrio detallado es un requisito paratrabajar
 
en las escalas de longitud en las que se produce el movimiento browniano.(75) El bombeo
estocástico rompe el equilibrio detallado utilizando un parámetro oscilante para generar un
ujo neto y mantener un estado estable de no equilibrio, y se ve en la naturaleza, por
ejemplo, en las ATPasas.(76, 77) Se han desarrollado varios marcos teóricos para explicar
varios modos de transporte impulsados por uctuaciones, como el bombeo estocástico, que
gobierna el transporte en motores y trinquetes brownianos.(78-126) A continuación se
comentan algunos ejemplos particularmente relevantes.

1.6 Mecanismos de trinquete


Los mecanismos de trinquete browniano se dividen en dos clases generales: trinquetes de
energía (los trinquetes pulsantes e inclinados son subcategorías de los trinquetes de
energía) y trinquetes de información.(127-129)

1.6.1 Trinquetes pulsantes

En trinquetes pulsantes, los mínimos y máximos de energía potencial se varían de manera


periódica o estocástica sin referencia a la posición de la partícula en la super cie de energía
potencial.(128)La forma más simple es un trinquete de encendido y apagado en el que el
potencial se enciende y apaga repetidamente más rápidamente que la difusión de las
partículas brownianas sobre su super cie de energía potencial. Esto da como resultado el
transporte direccional neto de partículas a través de la super cie ( Figura 3 ).

gura 3

Figura 3. Un trinquete de encendido-apagado: (a) las partículas están ubicadas en un mínimo


de energía, (b) el potencial se apaga para que la difusión pueda ocurrir por un corto tiempo, y
(c) el potencial se enciende nuevamente. Como el potencial es asimétrico, las partículas
tienen una mayor probabilidad de quedar atrapadas en un pozo adyacente a la derecha del
original que a la izquierda. (d) La relajación en los mínimos de energía local conduce a la
posición promedio de las partículas que se mueven hacia la derecha.(128)

Un trinquete intermitente es un subtipo del trinquete pulsante, que consta de una serie
repetida de máximos y mínimos. El descenso y aumento secuencial de partes de esta
super cie de energía potencial conduce al transporte direccional ( Figura 4 ).
Figura 4
  

Figura 4. Un trinquete intermitente. En (a) y (c), la partícula comienza en un pocillo verde o


naranja, respectivamente. El aumento de estos mínimos de energía mientras se reducen los
máximos y mínimos adyacentes desencadena el movimiento mediante el movimiento
browniano (b) a (c) o (d) a (e). Variando continuamente las alturas relativas de las barreras
de energía y los mínimos de los pozos de energía, la partícula se puede transportar
direccionalmente.(128)

Una super cie de energía potencial asimétrica no es estrictamente necesaria para generar
un trinquete pulsante.(128)Se puede utilizar una matriz periódica de mínimos locales que
viajan a una velocidad constante (es decir, una onda viajera) para formar un trinquete de
potencial viaje. En términos de sistemas realistas, el mecanismo de trinquete de potencial de
viaje tiene mayor relevancia para los procesos impulsados por el campo discutidos en la
sección 5 y los mecanismos de autopropulsión de la sección 6 .

1.6.2 Trinquetes basculantes

En un trinquete de inclinación, la super cie de energía potencial subyacente permanece sin


cambios y el equilibrio detallado se rompe mediante la aplicación de una fuerza impulsora
insesgada a las partículas, como el calor.(128)Cuando la fuerza aplicada es calor, los
trinquetes también se denominan trinquetes de temperatura o difusión ( Figura 5 ). En su
forma más simple, el mecanismo es muy similar al del trinquete de encendido y apagado.
Aquí, las partículas no pueden cruzar los máximos de energía a baja temperatura, pero un
breve aumento de temperatura permite la difusión de partículas a través de la super cie.
Una disminución de la temperatura atrapa este movimiento de una manera similar al
trinquete pulsante de encendido y apagado que conduce al transporte direccional. El período
de temperatura elevada debe ser breve para evitar la difusión general, que destruiría
cualquier direccionalidad ganada inicialmente.

Figura 5
  

Figure 5. A temperature or diffusion ratchet. (a) The particles are located in an energy
minima on the potential-energy surface, with energy barriers ≫kBT1. (b) The temperature is
increased so that the height of the barriers is ≪kBT2, and free diffusion is allowed to occur
for a short time. (c) As the temperature is lowered again, the asymmetric potential energy
surface means that the particles have a greater probability of being trapped to the right of
their initial position. (d) Relaxation to the local energy minima.(128)

El movimiento direccional también se puede lograr mediante la aplicación de una fuerza


direccional periódica a una super cie de energía potencial asimétrica en un trinquete
oscilante. Aunque la fuerza impulsora neta promedia cero, el movimiento direccional se
impulsa aprovechando la asimetría de la super cie de energía potencial ( Figura 6 ).

Figura 6

Figura 6. Un trinquete oscilante. (a) Las partículas están ubicadas en un mínimo de energía
en la super cie de energía potencial. (b) Se aplica una fuerza direccional hacia la izquierda.
(c) Se aplica una fuerza direccional igual y opuesta a la derecha. (d) La eliminación de la
fuerza y la relajación a un mínimo de energía conduce a la posición promedio de las
partículas que se mueven hacia la derecha.(128)

De manera similar, en un trinquete de fuerza uctuante, la fuerza impulsora aplicada se


genera aleatoriamente y varía estocásticamente. Finalmente, en un trinquete de inclinación
asimétrico, las partículas se encuentran en una super cie de energía potencial simétrica,
pero el potencial aplicado proporciona la asimetría requerida. Por ejemplo, en un trinquete
oscilante, la inclinación en una dirección sería más larga que en la otra.

1.6.3 Trinquetes de información

Los trinquetes de energía operan mediante la aplicación de una fuerza externa o la


modulación de una super cie de energía potencial independientemente de la posición de la
partícula.(128)Los trinquetes de información, por otro lado, se basan en subir o bajar una
barrera de energía según la posición de la partícula en la super cie de energía potencial (
Figura 7 ). Esto da como resultado que la distribución de partículas se aleje del equilibrio.
  
Este proceso requiere la transferencia de información de la partícula a la super cie de
energía potencial, de una manera que recuerda al demonio de Maxwell.

Figura 7

Figura 7. Un trinquete de información. En (a) y (d) las líneas punteadas representan la


transferencia de información sobre la posición de la partícula. (b) La posición de la partícula
baja la barrera de energía al movimiento hacia el pozo de la derecha pero no hacia la
izquierda. (c) La partícula se mueve por movimiento browniano.(128)

1.6.4 Lenguaje necesario para describir el funcionamiento de las máquinas


moleculares

Four terms are useful for describing the relationship of the components/substrates of
molecular machines in terms of dynamics: balance, linkage, ratcheting, and escapement.(75)
Having the components/substrates in a state of “balanced/unbalanced” or “linked/unlinked”
plays a crucial role in determining whether or not a machine can perform a task. “Balance” is
the thermodynamically preferred distribution of a particle (or submolecular component) in a
molecular system. The trigger for net transportation comes from the balance being broken
(i.e., the system being in a nonequilibrium state) (section 4.4). “Linkage” is the
communication necessary for the transport of a particle between parts of the machine.
However, the ability to exchange a particle between two parts of a machine does not
necessarily allow a task to be performed, as a driving force is also required. Linking and
unlinking operations are not limited to the addition or removal of steric bulk. They refer to any
action that has an in uence on the rate of a reaction. The movement of a particle up an
energy gradient is described as ratcheting. It involves capturing the transitory positional
displacement of a particle through the imposition of a kinetic energy barrier. Escapement is
the counterpart to ratcheting in which the lowering of a kinetic energy barrier allows a
ratcheted substrate to relax from a statistically unbalanced system toward a thermodynamic
sink.
2 Controlling Motion in Covalently Bonded Jump To

  
Molecular Systems

2.1 Controlling Conformational Changes


2.1.1 Correlated Motion through Nonbonded Interactions

At room temperature, organic compounds typically uctuate between rapidly equilibrating


stereoisomeric structures through the perturbation (rotation, lengthening, and shortening) of
covalent bonds. The relative stability and energy barrier to interconversion of con gurational
and conformational isomers depend on intramolecular and intermolecular interactions. The
rotational and vibrational freedom of a single substituent of a molecule is not only
determined by its connectivity, shape, and size, but also its interactions with neighboring
substituents.(130-133) For example, Mislow et al. demonstrated the steric in uence of
neighboring groups on rotation as part of their work on molecular propellers such as
compound 1 where two or more aryl rings are attached to a single atomic center (Z) (Figure
8). The rotation of one aryl ring about a C–Z axis is sensed by the other rings. Mislow
suggested that the motion of the rings is coupled in the sense that none of the rings move
independently of the others and described such correlated motion as “sympathic”.
(130, 134-138)

Figure 8
  

Figure 8. Chemical structure of a triaryl molecule as an example of a molecular propeller.


(130, 139)

Öki’s group conducted pioneering investigations into restricted rotation about sterically
hindered single bonds in molecules such as 9-aryl uorenes and triptycenes.(140-149) These
molecules showed very high rotational barriers, and, in some cases, it was possible to isolate
the rotational isomers (rotamers). Inspired by this work, Mislow and Iwamura simultaneously
introduced the concept of dynamic gearing of proximate substituents in crowded molecules
in molecular analogues of propellers and bevel gears.(150-167) The rst molecular bevel
gear consisted of two 9-triptycyl groups joined through their bridge head carbons to a central
atom (Figure 9). Experimental and theoretical studies con rmed that the triptycyl groups are
tightly intermeshed and undergo correlated disrotatory motion.(150, 156)

Figure 9
  

Figure 9. (a) Molecular bevel gear 2, consisting of two 9-triptycyl groups joined through a
bridge head carbon to a central atom.(150) (b) X-ray structure of the molecular bevel gear
(side view). (c) X-ray structure of the molecular bevel gear (top view). Adapted with
permission from ref 156. Copyright 1984 American Chemical Society.

Many other examples of dynamic gearing have since been published including a range of
triptycene derivatives,(168-174) metallocene-based gears,(175, 176) conformationally
restricted amides,(177-181) and a series of propeller molecules and other molecules with
selective rotation,(132, 133, 139, 182-194) such as 1,2-di-o-tolylnaphthalene moieties.
(195, 196)

The sandwich-shaped trinuclear silver(I) complex [Ag3(3)2] (Figure 10) developed by


Shionoya and co-workers(197-199) is an example of a different type of a rotational device.
The complex consists of two disk-shaped ligands, 3, consisting of three thiazolyl or 2-pyridyl
ligands and three p-tolyl groups attached to a central benzene ring. The three p-tolyl groups
force the neighboring ligands to adopt a nonplanar arrangement with respect to the central
aromatic ring, and make the coordination sites more accessible for silver(I) cations. The
exterior rings tilt at 30° in the same sense and thus form helical structures with P and M
geometries. A change from P to M helicity, and vice versa, was induced by 120° relative
rotation of the ligand in these complexes (Figure 10). These observations were further
studied using a heterotopic system [Ag3(3)(4)] consisting of a hexa-monodentate ligand 4
and a tris-monodentate thiazolyl ligand 3, which complex three silver(I) cations.(200) In the
case of the hexa-monodentate thiazolyl ligand 4, only every second ring is coordinated to a
silver cation. Correlated ipping motion of the coordinating rings (and ligand exchange)
results in the conversion from M to P and a 60° rotation of the two disks (Figure 10d).

Figure 10
  

Figure 10. (a) Chemical structure of tris-monodentate disk shaped ligand 3 and hexakis-
monodentate ligand 4. (b) Schematic representation of complex [Ag3(3)2] shown as its M-
helical enantiomer. (c) X-ray structure of complex [Ag3(3)2].(197) (d) Schematic
representation of the ipping motion of the rings attached to the disks and subsequent
ligand exchange from 1-A, 3-E, and 5-C in M to 1-B, 3-F, and 5-D in P. The direction of rotation
in the M to P transition is opposite to that of a subsequent P to M′ transition.(200) Reprinted
with permission from refs 197 and 200. Copyright 2003 and 2004 Wiley-VCH Verlag GmbH &
Co. KGaA, Weinheim.

Related heterotopic complexes were later used to transform the rotational motion of the
sandwich-shaped trinuclear silver(I) complex into translational motion using a molecular
crank mechanism. In molecular crank 5, a [2]rotaxane was attached as a translational
segment (Figure 11).(201)

Figure 11
  

Figure 11. (a) Schematic illustration of a crank mechanism that translates linear motion into
rotary motion in cylinder engines. (b) Schematic representation of a molecular crank
mechanism. (c) Chemical structure of a synthetic molecular crank 5.(201) Reprinted with
permission from ref 201. Copyright 2010 Royal Society of Chemistry.

Recently, Anderson et al. reported a zinc-porphyrin macrocycle coordinated by two templates


containing multiple pyridines (wheels) forming a caterpillar track complex.(202) NMR
exchange spectroscopy (EXSY) experiments showed that the ring underwent correlated
motion with correlated motion of the templating “wheels”.

The examples shown in this section clearly show the in uence that steric interactions can
have on submolecular motion. However, at equilibrium, these motions are nondirectional,
even during a partial rotation.

2.1.2 Stimuli-Induced Control of Conformation about a Single Bond

An early example of controlled random rotary motion about a single bond was provided by
Kelly et al.,(203-205) who utilized a cation binding event to halt the free rotation of a
triptycene moiety in “molecular brake” 6 (Scheme 1). The two pyridine groups must be
coplanar to maximize binding with Hg2+, thus raising the energy barrier to rotation by “putting
a stick in the spokes”. Sulfur oxidation has similarly been used to inhibit free rotation in 7
with mono- or dioxidation signi cantly slowing the rotation.(206) Rotation about a single
bond has also been promoted by substrate protonation by Shimizu and co-workers in 8.(207)
Protonation of the quinoline nitrogen led to an increase in the rate of rotation by 7 orders of
magnitude, through a proposed hydrogen-bonding stabilization of the rotation transition
  
state (Scheme 1). The same group observed that utilizing an acetate guest’s hydrogen-
bonding interactions to “turn on” rotation provided a more modest accelerating effect.(208)
Again, this was suggested to be via a lowering of the energy of the rotational transition state.
Other examples have also been reported.(209, 210)

Scheme 1

Scheme 1. Molecular Brakes Operated by (a) Hg2+ Binding and (b) Sulfur Oxidation; and (c)
Proposed Transition State Stabilization in Shimizu’s System(203, 206, 207)

Singlet oxygen has been used to promote rotation in combination with thermal relaxation in
a system based on rotation between the cis and trans isomers of 9 (Scheme 2). Reaction
with singlet oxygen forms the cis-substituted 9,10-endoperoxide selectively. Thermal
reversion generates the cis-anthracene, which upon extensive heating furnishes the more
thermodynamically stable trans-anthracene 9, completing the cycle.(211)

Scheme 2

Scheme 2. A Molecular Switch Flipped by Singlet Oxygen(211)

Kelly and co-workers reported an attempt to create a Feynman adiabatic ratchet-and-pawl.


(212-214) The design utilized a helicene pawl whose inherent helical chirality was intended to
bias the rotation of the triptycene cog. Although this design realizes an asymmetric potential
energy surface, 1H NMR showed no directional bias with equal rates of rotation in both
directions. This is in line with Feynman’s thought experiment(215) and illustrates that the rate
of rotation depends solely on transition state energy and temperature, and not on the shape
of the barrier to rotation. State functions, for example, free energy, do not depend on the
system’s history (Figure 12).
Figure 12
  

Figure 12. (a) Kelly’s ratchet-and-pawl 10. (b) Enthalpy change on rotation of helicene.
Reprinted with permission from ref 212. Copyright 1997 Wiley-VCH Verlag GmbH & Co.
KGaA, Weinheim.

The key factor required to break detailed balance and allow directional motion of the
triptycene cog is an energy input to drive the system away from equilibrium. Although a rapid
periodic variation in temperature could in theory cause the system to act as a temperature
ratchet, it would be very di cult to verify this experimentally. As such, Kelly et al. instead
proposed the modi ed structure 11,(216, 217) where a chemical reaction drives the system
from equilibrium (Scheme 3). The helicene pawl oscillates in the energy minima between
blades, and, ignoring the amine group, all three minima are identical. Formation of the
isocyanate from reaction with phosgene activates the system. In the course of random
uctuations, sometimes the helicene is close enough to the isocyanate to react. After
reaction, the helicene is trapped part way up the energy barrier; that is, the helicene is
“ratcheted” part way to rotation. Random thermal uctuations may then overcome the
smaller barrier to rotation, and the urethane can be cleaved to give overall 120° rotation. This
machine is a landmark achievement in the realization of chemically fueled directed rotation
in molecular machines and demonstrates most of the basic tenants of a directional motor.
However, attempts to modify the rotor to allow 360° rotation have thus far failed.(218)

Scheme 3
  

Scheme 3. Chemically Driven Directional 120° Rotation of Kelly’s Triptycene Rotor 11(216)

Mock and Ochwat have proposed a minimal molecular motor in which epimerization of a
stereogenic center, that is, a formal 180° rotation, is driven by the hydration of a ketenimine
fuel.(219) A biaryl lactone motif reported by Branchaud has been used to control motion
about a single bond in a nondirectional rotor based on the facially selective ring opening of a
chiral lactone.(220) They later improved on this system to report 180° directional rotation.
(221) Feringa et al. achieved full 360° directional motion about a single bond in a related
biaryl system (Scheme 4).(222, 223) Rotation was driven by four sequential chemical “power
strokes” with felicity of rotation for each step being between 90% and 100%. The cycle
involved four conformationally restricted intermediates, 12–15, with 12 and 14 locked by
covalent bonds, and 13 and 15 locked by their steric bulk. Although 13 and 15 are free to
dynamically equilibrate between helical forms, the chiral information in the reducing agent is
used to discriminate between them and thus drive directional motion. The rst step involved
asymmetric ring opening with the (S)-CBS reagent, and regioselective protection of the
resulting phenol with an allyl group. After oxidation and PMB deprotection with spontaneous
lactonization, 180° rotation was achieved. Another stereoselective ring opening and
regioselective phenol protection as the PMB ether followed, before oxidation and ring closure
furnished the product of 360° rotation in high delity. Additionally, rotation in the opposite
sense can be achieved by utilizing the (R)-CBS reagent and reversing the order of PMB/allyl
protection.

Scheme 4
  

Scheme 4. Directional 360° Rotation about a Single Bond via Four States A–Da

Scheme a Rotation is restricted in A and C by covalent bonds, although helical inversion is


allowed, and in B and D by nonbonded interactions, with directional control of rotation being
provided by stereospeci c covalent bond cleavage.(222)

2.1.3 Stimuli-Induced Conformational Control in Organometallic Systems

Control of the rotary motion of ligands in metal sandwich or double-decker complexes is


somewhat conceptually similar to controlling rotation around covalent single bonds. Aida
and co-workers demonstrated that porphyrin ligands in double decker complexes such as 16
can rotate with respect to each other depending on the central metal atom and their steric
bulk (Figure 13a).(224, 225) Studies on a variety of metal complexes found that the central
metal had an in uence on the speed of rotation.(226) Complexes synthesized from two D2h
symmetry porphyrin free-bases are chiral, and therefore their rotary motion corresponds to
enantiomerization and can thus be studied by CD spectroscopy. Racemization can be
induced by protonation or reduction of the metal center. The second reduces π–π
interactions between the two ligands as a consequence of the larger ionic radius of the metal
center.(227)

Figure 13
  

Figure 13. (a) X-ray crystal structure of porphyrin double decker complex 16 with cerium(IV).
Reprinted with permission from ref 228. Copyright 1989 Wiley-VCH Verlag GmbH & Co.
KGaA, Weinheim. (b) Representation of the cooperative binding of guest molecules to
porphyrin double decker complex 17.(229)

It was shown that the rotation of the cerium(IV) bis[tetrakis(4-pyridyl)porphyrinate] 17 could


be suppressed by successive cooperative guest binding of dicarboxylic acids (Figure 13b).
(229, 230) The double decker complex shows a large allosteric effect in this molecular
recognition event (once rotation has been suppressed by the rst guest molecule, guest
binding becomes increasingly favorable). Further guest molecules have been tested such as
β-dicarboxylate anions,(231) potassium ions,(232) mono- and oligosaccharides,(233-235)
and silver cations.(236) In the case of silver ions, it was shown that their cooperative binding
leads to a progressive and nonlinear increase in the rate of rotation.(237-244) The rotation of
a similar cerium double decker complex has been monitored on a single molecule level, by
the attachment of a bead visible under optical microscopy.(245)

Ever more complicated systems such as multidecker pophyrin complexes(246) and


mechanically interlocked porphyrin systems like 18 have been synthesized (Figure 14).
(247-249) The latter consist of cerium(IV) or lanthanum(III) bis(porphyrinate) double decker
complexes and one or two orthogonal porphyrin molecules. In these “molecular gears”, the
rotation of the top and side units could be partially coupled through mechanical interactions
between the different units. A folding ruler based on similar principles has been reported.
(250) A dodecanuclear four level complex that mimics a double ball bearing has also been
synthesized.(251)   

Figure 14

Figure 14. Cerium(IV) bis(porphyrinate) double decker (top unit) and a rhodium(III) porphyrin-
based side cog. The two units are connected through a coordination bond between
rhodium(III) and a pyridyl group.(249)

Ferrocene is a prototypical sandwich complex with an iron atom symmetrically aligned


between two C5H5 (cyclopentadienyl) rings.(252, 253) The barrier to rotation of the two rings
about the C5 axis is very small in the gas phase. Using electrospray ionization and theoretical
studies, it was shown that ferrocene derivative 192– can serve as a two-state rotary switch,
which works through proton transfer (Scheme 5). Differing electrostatic interactions lock
192– and [19·H]− in two different conformations. The trans conformation of 192– is more
stable because of Coulombic repulsion between the negative charges, while the cis
conformation of [19·H]− is stabilized through hydrogen bonding (Scheme 5).

Scheme 5
  

Scheme 5. Control of Rotation in Ferrocene Complex 19 through Protonation(253)

This nding was followed by studies of many more ferrocene derivatives, and the control of
their rotary motion through different stimuli such as electron transfer and photochemistry.
(254, 255) Ferrocene-based rotors have recently been used to form self-assembled
nanostructures.(256)

Metallocarboranes tend to have rather high barriers to rotation around the metal–ligand axis,
but some complexes, for example, nickel complex [Ni(20)2], can be used as
electrochemically controlled rotary switches (Figure 15).(257-261)

Figure 15

Figure 15. (a) Dicarbollide ligand 202–. (b) Metallocarborane [Ni(20)2], an electrochemically
controlled rotary switch.(261)

2.1.4 Stimuli-Induced Conformational Control over Several Covalent Bonds

In biological systems, host–guest binding is often used to induce conformational and


functional changes. These changes can vary from small, localized, bond deformations to
long-range rearrangements across multiple covalent bonds. These deformations are
frequently used to alter binding at distal sites, that is, allosteric regulation. There are several
synthetic examples where host–guest binding causes a su ciently large amplitude change
to merit discussion in this Review. Some of the earliest examples of allosteric receptors
came from Rebek and co-workers.(262) For example, the binding of tungsten by negative
  
allosteric receptor 21 reduces the receptor’s a nity for potassium ions, as in this more rigid
conformation binding to potassium would induce a degree of strain (Scheme 6).(263, 264)
The positive allosteric receptor 22 developed by the same group depends on the initial
mercury binding event preorganizing the linked receptor for a second, more favorable binding
event.(265) These early examples have sired an extraordinary and diverse array of synthetic
allosteric receptors.(262, 264, 266-279)

Scheme 6

Scheme 6. (a) Negative Allosteric Receptor 21, Where Binding of Tungsten Forces Ring
Contraction and Steric Clash of the 3- and 3′-Substituents of the Bipyridyl Unit; and (b)
Positive Allosteric Receptor 22(263, 264)

Organic guests can also cause large-scale changes to host molecules, as they maximize
favorable interactions such as in molecular tweezer 23 (Figure 16) where the distance
between sidewalls halves on binding certain guests.(280, 281) A large amplitude change was
observed in the exible system 24 with an extended conformation being exchanged for a
more rigid “sandwich” form to maximize guest binding.(282) Guest-induced change is also
observed in clip 25 where the sa form predominates in solution in the absence of a guest,
but the aa form dominates in the presence of a suitable guest (Figure 16).(283-285) This
system has been further modi ed in 26 where the binding of a potassium ion to each crown
ether preorganizes the system for organic guest binding with increased a nity.(286)

Figure 16
  

Figure 16. (a) Molecular tweezer 23, arrows indicate direction of contraction. (b) Clip 25,
where the aa form is favored in the presence of a guest. (c) Clip 26, where ion binding
enhances guest a nity. (d) Crystal structure of uncomplexed and complexed 24.
(282, 283, 286, 287) X-ray crystal structure reprinted with permission from ref 282. Copyright
2007 American Chemical Society.

Metal coordination has been used by Lehn et al. to switch from an arrangement where
parallel anthracene units can bind electron-poor 2,4,7-trinitro-9- uorenone (TNF) in a
tweezer-like manner, to an open form 27 (Scheme 7). Here, coordination of copper in a
bidentate manner by the aromatic nitrogens causes a rotation about the heterocyclic axis
and opens the guest binding cavity, preventing TNF coordination. In related 28, zinc
coordination is a prerequisite of TNF binding.(288)

Scheme 7
  

Scheme 7. (a) Prevention of Organic Guest Binding on Complexation of Copper to 27; and (b)
Organic Guest Is Only Bound in the Zinc-Complexed Form of 28(288)

A notable example of chelation control of conformation was provided by trisaccharide 29


(Scheme 8). Although the tetra-equatorial isomer is the dominant form in solution,(289) the
addition of Pt(II) led to coordination, and a preference for the tetra-axial form, the better to
bind the metal ion. This process could be reversed by removal of the Pt(II) with excess
NaCN.(290, 291) This process has been replicated with a variety of cyclic substrates using
both metal ion binding(292-294) and pH variation.(295-297) Sauvage reported the collapse
of a porphyrin containing [4] rotaxane on copper removal.(298) Haberhauer’s molecular
hinge, 30, based on a 2,2′-bipyridine motif, is another example of large amplitude motion
controlled by coordination of a metal ion. In this case, the open system was free to rotate
over 180° with its rotation being constrained by an in exible backbone. Upon coordination of
Cu(II), the hinge shut. Upon copper removal, the hinge was again able to open, but only in one
direction due to the constraints of the backbone.(299) A light-driven switch and a chirality
pendulum based on a similar system have been reported by the same group.(300, 301)

Scheme 8

Scheme 8. Chelation Control of Equatorial/Axial Conformers and Haberhauer’s Molecular


Hinge(290, 299),a

Scheme a Reprinted with permission from ref 299. Copyright 2008 Wiley-VCH Verlag GmbH
& Co. KGaA, Weinheim.
Cavitands derived from resorcin[4]arenes have been shown to undergo large conformational
changes between open “kite” and closed “vase” forms of 31 (Scheme 9).(302) Because   of 
solvation effects, the kite form is normally favored at lower temperatures and the vase at
higher.(302-304) However, this change has also been achieved using protonation of or metal
coordination to the quinoxaline nitrogens,(305-308) or by altering the design to incorporate
redox-active centers.(309, 310) Substitution of the cavitand with amide groups can lead to
stabilization of the vase form by dimerization or guest complexation.(311-320) As an
alternative to coordination of a guest causing a change in conformation, complexation can
be intramolecular as in macrocycle 32 reported by Stoddart et al. The system rests in the
self-complexed form until reduction of the cyclophane decreases subcomponent
interactions, followed by release of the system to an unrestrained, unencapsulated form.
(321-326) A similar system, 33, was realized by Feringa and Qu(327) with directional rotation
inhibited in the complexed state, but allowed when uncomplexed (Scheme 9).

Scheme 9

Scheme 9. Resorcin[4]arene 31, Vase and Kite Conformers, Stoddart’s Self-Complexing Lock
32, and Feringa’s Self-Complexing Rotor 33(98, 321, 322, 327),a

Scheme a Reprinted with permission from refs 321 and 327. Copyright 1997 and 2010 Wiley-
VCH Verlag GmbH & Co. KGaA, Weinheim.

Hamilton and co-workers introduced conformational switch 34 (Scheme 10) where in the
initial state the equilibrium is mildly biased toward the benzamide station (1.3:1). Upon
protonation of the dimethylamine with TFA, a dramatic reversal of this bias is observed
(1:99).(328) The oxidation of a copper center has been used to trigger coupled rotation and
copper exchange between two independent switches (based on the structure shown in
Scheme 45).(329)

Scheme 10
  

Scheme 10. Hamilton’s Acid Sensitive Switch(328)

Scheme 11

Scheme 11. pH-Driven Conformational Change in Oligomeric System 35(330)

Oligomeric systems have been used to great effect in generating large-scale conformational
changes, often exploiting the helical secondary structures that can be formed by many
species.(330-367) Polyamide 35 resides in a helical arrangement, but protonation of four
diaminopyridine units causes a dramatic extension to the linear form of 35 (Scheme 11).
Further protonation leads to another helical form, and the system can be returned to its initial
form by deprotonation.(330) An extension of this system was shown to expand from 12.5 to
57 Å upon protonation.(368) An alternating pyridine and pyrimidine oligomer formed helices
(due to the favored disposition of neighboring rings being transoid), but upon addition of
Pb(II) to 36, coordination led to a linear form with a concurrent increase in length from 7.5 to
38 Å (Scheme 12). This process could be made switchable by sequestering the Pb(II) until
needed in pH responsive cryptand 37.(369) A similar process has been achieved using Ag(I)
to modulate expansion.(370)   

Scheme 12

Scheme 12. Large-Scale Extension of a Ligand Strand upon Pb(II) Complexation(369, 370)

An interesting recent example was provided by Stadler and Lehn(371) where a linear and a
helical domain were synthesized in the same molecule. In this two-domain system,
coordination of Pb(II) led to the unfurling of the helical domain with concurrent curling of the
linear portion of the molecule. Upon removal of Pb(II), the system reverted to its original
state, thus representing a reversible device where an extension process is coupled to a
furling one, on both complexation and decomplexation of a metal ion.

2.2 Controlling Con gurational Changes


Changes in con guration, especially cis/trans isomerization, have been used in many
molecular systems to control motion.(372) Although the sometimes small amplitude motion
is not always useful for the design of molecular machines, it represents an interesting tool
for inducing directional motion in synthetic and supramolecular systems. The most
prominent examples of this type of system are the photoisomerization of stilbenes and
azobenzenes,(373-377) the reversible electrocyclization of diarylethenes,(378-382) the
photochromic reactions of fulgides,(383, 384) the interconversion of spiropyrans with
merocyanin,(385) as well as chiroptical switching of overcrowded alkenes.(386-388) In
recent years, this topic has become an extensive area of research, and a wide range of
applications exploiting switchable systems have been proposed and realized.(389-395) Most
of these applications rely on intrinsic electronic and spectroscopic changes on
interconversion between the two species, but some use small con gurational changes in a
more mechanical fashion, and the resulting devices can be seen as molecular machines.

Con gurational changes, especially the photoisomerization of azobenzenes, have been used
to induce changes in the structures of biologically relevant molecules such as antibiotics,
peptides, and DNA,(374, 396-406) but also as part of small-molecule systems such as
molecular scissors, tweezers, and other molecular machines.(407-412) Some of them will be
introduced in the following section.   

Scheme 13

Scheme 13. (a) Molecular Scissors 38;a and (b) Structure of Molecular Hinge 39(414)

Scheme a Irradiation at 350 nm initiates photoisomerization from the trans to cis isomer of
the azobenzene moiety (closed to open molecular scissor); irradiation at >400 nm induces
the reverse.(413)

Molecular scissors 38, described by Aida and co-workers, consist of an azobenzene unit and
a ferrocene unit (Scheme 13a).(413, 415) The open and closed forms were interconverted by
photoisomerization of the double bond of the azobenzene unit, which led to an angular
change of 49° around the cyclopentadienyl rings of the ferrocene. Another example was the
synthesis of the molecular hinge 39 where two xanthene rings were connected by −N═N–
linkers (Scheme 13b).(414) Photoisomerization at 366 nm from the trans/trans, which is
almost planar, to the cis/cis state resulted in a change of approximately 90° in the angle
between the two aromatic rings. Reisomerization could be achieved by irradiation at 436 nm.
Azobenzenes have been used for the construction of photoswitchable azo-macrocycles.
(389, 416) Azo-macrocycles have found applications in host–guest chemistry as they can
selectively and reversibly bind ions. Cyclic azobenzenes have been used to switch on or off
the rotation of subunits.(417, 418) A recent example showed that the rotation of a 2,5-
dimethylbenzene rotor in the cyclic azobenzene 40 could be switched off in the E-isomer of
the azobenzene, while rotation was allowed in the Z-isomer (Figure 17).

Figure 17
  

Figure 17. (a) Chemical structure of a molecular brake 40 with a 2,5-dimethylbenzene unit as
the rotor. (b) X-ray crystal structure. (c) Schematic representation of this photoinduced
molecular brake. X-ray crystal structure reprinted with permission from ref 418. Copyright
2011 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

Azobenzenes have also been used to control motion in molecular shuttles (see also section
4.3) and to control the threading and dethreading rate of a macrocycle when attached to the
ends of a thread. In a recently published example, E → Z photoisomerization of azo-end
groups slowed the threading–dethreading of a ring and turned the pseudorotaxane into a
kinetically inert rotaxane.(419)

The substitution of an alkene can lead to steric overcrowding around a double bond. When
the substituents are large enough, the planarity of the double bond can be disrupted and a
variety of twisted and folded conformations can be adopted, including helices. The area
where the substituents come into close proximity is often referred to as the fjord region.
These types of compounds still undergo photoisomerization and can be used as chiroptical
molecular switches.(386, 420, 421) Feringa et al. have developed several molecular motors
exploiting overcrowded alkenes as chiroptical molecular switches.(422-428) The rst
generation of this molecular motor (Scheme 14, compound 41) featured two identical halves
connected by a central double bond.(429-436) Each of the halves had a stereocenter, which
was crucial to controlling the rotational process. Rotation of 360° around the double bond
was repetitive and directional. Two light induced cis–trans isomerizations led to a 180°
rotation and were each followed by thermally controlled helicity inversion, which effectively
blocked reverse rotation. The cis–trans photoisomerization was evident from 1H NMR, and
the simultaneous exchange of P to M was detected by circular dichroism (CD). A number of
different structures have been synthesized to optimize the rotation process and to
explore
 
the limits of the system.(437-444)

Scheme 14

Scheme 14. Minimization of Steric Interactions between Aromatic Groups and Substituent
Results in (P,P)-(trans)-41-Stable Ground-State Conformation, with Axial Substituentsa

Scheme a (i) Photochemical isomerization leaves substituents in the unstable equatorial


conformation (M,M)-(cis)-41-unstable. (ii) Steric strain is released by a thermally activated
helicity inversion leading to (P,P)-(cis)-41-stable. (iii) Photoisomerization from cis to trans.
(iv) Helical inversion completes a full rotation.(429)

In the second generation of Feringa’s molecular motor 42, two halves that are not identical
were connected by a central double bond.(445-447) One half was replaced by a tricyclic
aromatic group and there was only one stereogenic center present, but the rotation operated
according to the same principles as the rst generation motor 41 (Figure 18). Some dramatic
enhancements of rotary speed were achieved in the second generation of motors.
(428, 448-453)

Figure 18
  

Figure 18. Comparison of the general structures for the rst (41) and second (42)
generations of Feringa’s molecular motors.(425)

Feringa and co-workers have shown that transmission of molecular rotation from one part of
a molecule to another is possible. In motor 43, a xylyl unit was attached to the lower half of
an overcrowded alkene-based switch. By switching between the cis- and trans-forms, the
rotation of the xylyl group could be controlled (Figure 19).(454, 455) Imines have recently
been used as directional light-driven rotors by Lehn and Greb.(456)

Figure 19

Figure 19. Structure of molecular brake 43.(454)

It has long been known that irradiation of hydrazones can induce E → Z isomerization
around the imine bond (C═N). However, the Z-isomer is usually thermodynamically unstable
and thus short-lived.(457, 458) Intramolecular hydrogen bonding in various hydrazone
derivatives can kinetically stabilize the Z-isomer, which is formed by irradiation.(459-461) The
Aprahamian group has studied the E/Z isomerization of hydrazones using chemical
switching inputs such as protons and zinc ions.(462-467)

Scheme 15
  

Scheme 15. Hydrazone-Based Switchesa

Scheme a (a) Protonation of the pyridyl nitrogen of compound (E)-44 induces E → Z


isomerization.(468) (b) The quinoline unit in (E)-45 provides a binding site for a Zn2+ ion,
which is now the trigger for E → Z isomerization.(469)

In their original system, E→ Z isomerization was induced by protonation of a pyridyl ring that
initially acts as an intramolecular hydrogen-bond acceptor in the E-isomer (Scheme 15,
compound (E)-44).(468, 470) In the Z-isomer, the carbonyl oxygen stabilizes the Z-H+ form of
the molecule because the pyridyl unit is no longer available as a hydrogen-bond acceptor.
Upon deprotonation, the process is reversed. However, the Z-isomer is metastable, and as a
consequence the removal of the proton does not lead to an immediate change. The
thermodynamically more stable E-isomer reappears with time. The Aprahamian group has
also replaced the naphthalene unit with a quinoline unit (see (E)-45), thus introducing an
additional hydrogen-bond acceptor and an additional metal ion binding pocket next to the
hydrazone.(469) In this system, a Zn2+ ion now triggers E → Z isomerization. The most
recent system showed that on addition of Zn2+ to (E)-46 and (E)-47, a switching cascade
could be initiated.(471) Coordination of Zn2+ to (E)-46 lowers the pKa of the imidazole N–H
by 3 units and allows the protonation of (E)-47, which in turn allows switching of this moiety.
Without metal coordination, the imidazole is not acidic enough to catalyze switching in
(E)-47, making this sequence an elegant example of one switching event acting as the input
for another, in one pot. Because the stator and rotor are held in xed relative positions, the
switches developed by Aprahamian and co-workers represent very promising tools for the
control of directional motions in molecular machines (Scheme 16). A hydrazone switch has
been used to dope a liquid crystal with the color of the liquid changing upon switching.(472)
  

Scheme 16

Scheme 16. Switching Cascade of Hydrazone Rotary Switches(471)

3 Controlling Motion in Supramolecular Systems Jump To

The use of supramolecular systems in molecular machines is challenging as competing


processes such as the association and dissociation of subcomponents from and to the bulk
must be controlled.(473, 474) These processes can interfere with the integrity and/or
processivity of the machine unless the binding is part of the motion generation process itself
or the effect of unbound elements is restricted. If host–guest interactions are used to
assemble a machine, then exchange with free hosts or guests in solution should be avoided
to maintain the integrity and kinetic stability of the machine. Disassembly or exchange can,
in certain cases, be intentionally exploited for functional purposes and to provide
communication with external species. Controlled motion in supramolecular systems must
be built on cooperative interactions of individually weak but collectively strong noncovalent
interactions.

In biology, large-scale molecular motion generated upon guest binding is a common


process. Acetylcholine neurotransmitters are among such guests and bind to a
transmembrane channel receptor, inducing a large conformational change, resulting in the
controlled ux of ions through the membrane.(475) In this case, any guests in the bulk can
act as an inducer of molecular motion and at the same time as a messenger of the
microenvironment enabling dynamic communication with the surroundings. The stabilities
of most biological functional assemblies are well maintained by restriction of bulk exchange.
If this was not so, rapid exchange of the ribosome complex or individual subunits of the
ribosome during protein synthesis would detach the enzyme from the substrate mRNA and
release a partially synthesized peptide, unlikely to be of any use. However, mRNA and the
ribosome form a kinetically stable complex, subunit exchange with the bulk is restricted, and
the enzyme processively produces large proteins.(476)   

3.1 Guest Binding Generating, or Assisted by, Large-Scale


Intramolecular Motion
Considering the above discussion, molecular machines and devices based on noncovalent
assemblies require careful design. The relatively large molecular level motions generated
upon guest binding are of great importance in the design of practical molecular machines.
Guest-induced molecular motion has been widely studied in the literature with organic
compound,(477) anion,(478) cation,(479-481) metal,(464, 471, 482, 483) or proton guests.
(484, 485) Switches based on pseudorotaxanes have also been reported.(486-518) These
supramolecular complexes display interesting operational properties that are useful for
device development.(495, 519-564) These switches can be used to build relatively
sophisticated devices such as the molecular tweezers 23–26.(280, 281, 287, 288, 565-575)
This concept conventionally refers to receptors having a cavity with two binding motifs. The
binding pocket can either readily encapsulate the guest (lock and key type) or evolve to a
more suitable conformation upon guest binding (induced t type). Molecular machines can
be regulated by responsive tweezers, with motion induced and regulated through an external
stimulus such as the binding of guest molecules.

Scheme 17

Scheme 17. Light-Operated Molecular Tweezers(575)

One of the earliest examples of a molecular machine was a “molecular tweezer” based on
the photoisomerization-dependent binding of cations in a crown ether-bearing azobenzene,
developed by Shinkai et al. (Scheme 17).(575, 576) The open trans form can bind smaller
cations such as Na+ selectively, whereas the cis form prefers to bind larger Rb+ ions by
forming a sandwich complex with the cation. Moreover, the cis isomer can be stabilized by
the sandwich complex in the presence of Rb+, and the photostationary state biased toward
the cis isomer, causing the molecule to relax to the trans isomer much more slowly. Hence,
in addition to photoisomerization-dependent changes in the a nity of the receptor, guest
binding can alter the kinetics of the photoisomerization process.

Scheme 18
  

Scheme 18. Decreased Drug Binding A nity by a pH-Induced Conformational Change(573)

Intercalation of electron-poor aromatic guests by molecular tweezers having two π-systems


has also been explored by several groups.(577-581) The binding and release of the guest is
modulated by the motion of the receptor components. Intramolecular motion has been
induced allosterically by means of various stimuli including metals,(582-584) anions,
(585, 586) electrochemical reduction,(587) and pH change.(573) 2,6-Substituted pyridine 49
has been shown to change conformation upon protonation of the pyridine, such that the
methoxy groups form hydrogen bonds with the pyridinium proton (Scheme 18).(573) This
hydrogen bonding results in a rotation around the single bond moving the naphthyl moieties
away from each other. Intramolecular motion results in the release of the hydrophobic drug
mitoxanthron, which had been bound by π–π interactions between two naphthyl receptors.

Figure 20

Figure 20. Orthogonal exclusion of an azobenzene and a viologen encapsulated by CB[8].


(588)

Extracting work from machines depending solely on supramolecular interactions would be


di cult due to the rapid exchange of bound and unbound guests. However, some
supramolecular host–guest systems are moderately stable and might be suitable for use in
the assembly of molecular devices. A methylviologen (compound 50, MV2+) and a trans-
azobenzene derivative (compound 51, trans-AB) were encapsulated by a cucurbit[8]uril
(CB[8]) host (Figure 20). Each guest has distinct properties: MV2+ is redox active, and the
azobenzene can be isomerized to its cis-form (cis-AB) on irradiation.(588) One electron
reduction of methylviologen (MV+•) results in a binary encapsulation of the reduced species
in the host and trans-AB is kicked out. Photoisomerization of azobenzene to its cis -isomer
 
results in its ejection due to steric repulsion. Therefore, the initial heteroternary complex of
MV2+, trans-AB, and CB[8] can be orthogonally switched to give two distinct complexes. In
one complex, one-electron reduction of MV2+ forces out the azobenzene to form a
homoternary complex (MV+•)2CB[8], while in the other photoisomerization of the azobenzene
ejects cis-AB, while MV2+ stays encapsulated to form (MV2+)CB[8].

These three distinct, stimuli-dependent, modes of complexation allow greater control over
molecular behavior. To transfer this to the macroscopic world, thiol-functionalized trans-
azobenzenes were attached to a gold surface. Upon complexation by CB[8], together with the
uorescently labeled MV2+, the surface became uorescent, and due to the charged nature
of viologen, a substantial increase in wettability was observed. Upon irradiation, the
azobenzene was kicked out as its cis isomer, returning the surface to a non uorescent,
hydrophobic state, after the unbound uorescent viologen–CB[8] complex was washed away
from the surface. The orthogonal control of switching processes is a vital tool for the
creation of even more complex molecular devices, especially in memory device construction.
(588)

One of the interesting emerging properties of supramolecular host–guest systems is the


stimuli-dependent expansion and contraction of molecular constructs. This behavior will be
revisited in rotaxane systems in section 8. In double-stranded helicate 52 (bridged by
spiroborates at each end), sodium ion-induced reversible extension and contraction was
observed with an accompanying directional twisting (Figure 21).(589) In the presence of Na+,
the molecule was forced to twist its central tetraphenol moiety to better coordinate the
cation. The handedness of the helicate was preserved during this process. This process
could be reversed by the sequestration of Na+ by [2.2.1]cryptand.

Figure 21

Figure 21. Na+-induced twisting in helicate 52. Reprinted with permission from ref 589.
Copyright 2010 Nature Publishing Group.
The tetraphenolic central compartment could be replaced by two porphyrins to create a
hydrophobic cavity for aromatic guests to stack inside.(590) Upon intercalation ofelectron-
 
de cient aromatic guests between the porphyrin rings, the distance between the two
porphyrin units increased from 4.1 to 6.8 Å. This expansion in the central region induced a
rotation, which reduced the torsion angle, and, due to steric constraints, the capping
spiroborates unwound in a directional manner, demonstrating the rst guest-induced rotary
motion with accompanying corkscrewing motion. Guest control of screw motion, involving
simultaneous linear translation and rotation, has been reported.(591) In this case, the guest
acted as a template for the formation of helicates based on hydrogen bonding and biased
the formation of one of the two equilibrium structures.

Stimulus-driven molecular level conformational change and subsequent control over the
binding a nity of guest molecules have been used to regulate anion concentration in
solution.(592-594) Flood et al. developed a triazole-based foldamer bearing an azobenzene
moiety in the structure (Figure 22).(595) Hydrogen bonding between the triazole units and a
chloride anion together with π-stacking within the foldamer backbone kept the anion buried
in the interior.(592, 596-602) Photoisomerization of the azobenzene moieties disrupted the
stacking, inducing unfolding and releasing the anions. This photoinduced release and
capture of anions is completely reversible. Calculated K values in acetonitrile decrease
signi cantly from 3000 to 380 M–1 upon irradiation with UV-light. In addition, the conductivity
of the irradiated sample increased due to release of anions into solution. Helical inversion
caused by ion binding has also been observed.(603) Stimuli responsive switches have great
potential in drug release and ion-cargo transport applications. Nitschke et al. have reported a
complex system, where the addition of various signal molecules to a helical complex led to a
cascade of changes in the self-assembled system and various product distributions could be
formed.(604)

Figure 22
  

Figure 22. Control of anion concentration in solution by a photoresponsive foldamer.


Reprinted with permission from ref 595. Copyright 2010 American Chemical Society.

In the above examples, binding properties are regulated by an external stimulus or by


allosteric control. Large-scale molecular motions regulated by guest binding are of particular
interest in molecular machine design, because the presence of the regulator acts as a
chemical communication between the environment and the machine. This dynamic
interaction would allow the development of smart molecular machinery that performs
advanced operations dependent on the chemical environment, or perhaps allows
communication between machines.

3.2 Intramolecular Ion Translocation


The molecular basis for the conversion (in archea) of photon energy into more useful
potential energy relies on proton translocation between a photoisomerized retinal molecule
and the amino acid residues on the protein scaffold in which it is buried. This process occurs
with a well-de ned directionality, leading to light-mediated directional proton pumping across
a membrane.(384, 605) In supramolecular systems, for accurate translocation to take place,
the ion to be translocated should be more kinetically available for intramolecular motion than
the bulk ion. An electrochemical gradient cannot be achieved if each ion acceptor takes ions
from the bulk. Ion translocation can be controlled by changing the a nity of organic
receptors or metal centers for the ion of interest, through external stimuli such as light or
reduction/oxidation.(606) In the case of proton transfer, some molecules have distinct
photophysical properties such as a change in the pKa of the molecule in the excited state,
which enables rapid dissociation of the proton from one part of the molecule and
reassociation with another, intramolecular acceptor, in an “excited-state intramolecular
proton transfer” (ESIPT).(607) This process has been extensively studied using 2-(2-
hydroxyphenyl)-benzothiazole (HBT, Scheme 19).(608) Fast proton transfer from the hydroxyl
group to the nearby thiazole nitrogen takes place upon excitation of the molecule to generate
 
a cis-keto form with a distinct, solvent-dependent, spectroscopic character. In this molecule,
intramolecular hydrogen bonding mediates this fast proton transfer between different
regions of the molecule.

Scheme 19

Scheme 19. Excited-State Intramolecular Proton Transfer in HBT(607, 608)

Metals having more than one stable oxidation state can be translocated by oxidation or
reduction or incorporated into a scaffold along with other ions that can be translocated. For
either process, the molecule should bear more than one binding site, with different a nities,
to drive the motion of the ion. Flexible receptors are most often used for ion translocation,
but rigid aromatic systems can be suitable as well. These processes are usually
accompanied by a visible color change due to changes in absorption after a change in the
oxidation state of the metal, or due to ligand exchange. The reversible translocation of a
chloride anion from a Cu(II) center to a Ni(III) center was an early example of a exible
system (Scheme 20a).(609) The concentration-independent nature of the process indicated
that anions translocate from the Cu(II) center in an intramolecular fashion rather than via
exchange with the bulk.

Scheme 20

Scheme 20. Intramolecular Anion Translocation Mediated by (a) Redox(609) and (b) Ligand
Exchange(610)
Work by Mirkin et al. showed reversible intramolecular shuttling of a chloride anion from a
bisurea binding site to Rh(I) in the presence of carbon monoxide (Scheme 20b).(610) The 
hydrogen-bonding ability of the urea was further enhanced by attaching electron-
withdrawing 3,5-bis(tri uoromethyl)benzyl groups. Rh(I) adopts a distorted square-planar
geometry with these phosphine and thioether ligands, as shown by the crystal structure.
Initially, chloride was encapsulated between the urea tweezers. With the aid of polar solvents
(dimethyl sulfoxide or dimethylformamide) and in the presence of CO, chloride was
translocated to Rh(I) with accompanying CO coordination. The Rh–S bonds were cleaved
concurrently. Translocation was driven by both the weakening of hydrogen-bonding
interactions in the urea and the more favored and less distorted geometry adopted at the Rh
center with the new ligands. This is an important example of ion translocation in view of the
fact that it involves two different types of interaction (hydrogen bonding and metal–ligand
coordination) to enable shuttling and may open a path to the development of further
molecular devices by the orthogonal control of these two interactions.

The rst intramolecular cation translocation driven by auxiliary redox reactions was reported
by Shanzer et al.(611) A triple-stranded helical scaffold was used with ditopic hydroxamate
and bpy ligands, which bind preferentially to Fe(III) and Fe(II), respectively (Scheme 21).
Chemical reduction of Fe(III) with ascorbic acid relocated the cation to the bpy ligand with a
visually observable color change from pale brown to violet-red. Reversibility was obtained by
oxidation with (NH4)2S2O8. Even though the process takes place slowly (minutes to hours),
the overall movement of the cation is intramolecular.

Scheme 21

Scheme 21. Redox-Mediated Intramolecular Ion Translocation(611)


Several other examples of anion and cation translocation between ligands separated by
exible linkers have been investigated.(480, 612-614) Metals are also known to translocate
  
2
over rigid sp hybridized carbon surfaces.(615) Sandwich complexes of two dipalladium
units (four palladium nucleii arranged in two separate regions) clustered between π-
conjugated ligands showed redox-driven reversible translocation, resulting in a
tetrapalladium complex (Scheme 22).(616) In a complex with further extended conjugation,
redox reaction conditions resulted in reversible carbon–carbon formation between the
dipalladium clusters. This barrier (the new C–C bond) built by the mobility driving force used
in the rst complex (redox reaction) prevented translocation in the second complex and
hence con ned the metals to their original positions.

Scheme 22

Scheme 22. Redox-Mediated Pd Translocation in an Aromatic Sandwich Complex(616)

Supramolecular host–guest systems with distinct stimuli responsive switching behavior,


guest sorting character, and emergent photophysical properties have been developed with
implications for the design of molecular machines. To enable controlled mechanical motion,
these systems should be kinetically stable enough to avoid exchange of components with
the bulk on the time scale of the operation. Mechanically bonded molecular assemblies are
viable candidates for machine architectures, considering their stable interlocked structure,
and the relative ease of controlling their mobility.

4 Shuttling in Rotaxanes: Inherent Dynamics Jump To

Motion in rotaxanes is constrained by the nature of the mechanical bond. Very limited
motion is permitted orthogonal to the axle, while a “shuttling” motion “powered” by random
Brownian motion can be observed along the axle (as bounded by large stoppers). As the
templating methods typically used to form interlocked structures normally leave recognition
motifs,(23, 617-676) it is rare, although not unknown, to form a rotaxane without residual
interactions between thread and macrocycle.(504, 672, 677-689) These residual interactions
are typically viewed as “stations” on the rotaxane, and the shuttling of the macrocycle
between these stations grants a useful handle in these molecules. The rates of shuttling
between stations and their occupancy can be controlled by the strength of station–
macrocycle interactions.
4.1 Shuttling in Degenerate, Two-Station, Molecular Shuttles
  
The rst two-station degenerate [2]rotaxane 61 was reported by Stoddart et al. and exhibited
temperature-dependent shuttling between the two equivalent hydroquinol groups, as shown
by 1H NMR.(690) Many rotaxanes based on similar macrocycles and stations have since
been reported.(691, 692) Similar processes are observed in the amide rotaxanes 62–64
where two diglycine units are separated by various linkers. In 62–64, rapid shuttling of the
macrocycle between the degenerate stations was observed at 298 K. Only when the linker
was replaced with a bulky N-tosyl group was shuttling inhibited (removal of the tosyl group
restored rapid shuttling) (Scheme 23).(693)

Scheme 23

Scheme 23. (a) The First “Molecular Shuttle”, 61;(690) and (b) Degenerate Peptide-Based
Molecular Shuttles 62–64 of Varying Linker Length(693)

As movement between the binding sites must involve at least partial rupture of the
hydrogen-bonding network, the migration can be represented by the simpli ed energy
diagram in Figure 23. Hydrogen-bonding solvents have been shown to disrupt macrocycle–
thread interactions in single station rotaxanes,(694) and here addition of 5% [D4]methanol
increased the rate of shuttling 100-fold, consistent with lowering the energy barrier to
migration by disrupting station–macrocycle interactions and thus raising ground-state
energies.(693, 695-697) The effect of water on the rate of shuttling has been investigated
and was found to be greatly superior to that of other protic solvents.(698) This effect was
attributed to the ability of water to form three-dimensional hydrogen-bonding networks.
  

Figure 23

Figure 23. Idealized model of binding in degenerate molecular shuttles. Barrier height is
dependent on the energy required to break interactions between macrocycle and station and
a distance-dependent diffusional component.(14)

The use of a phenol/phenolate central group where the formation of a phenolate/cation pair
signi cantly slowed shuttling(699) has also been explored. Cleavage of a nitrogen protecting
group has been used to induce degenerate shuttling.(700) A large barrier to shuttling could
be introduced in 65 where the dimerization of the rotaxane, as templated by Cu(I) addition,
prevented shuttling (Scheme 24). This could be reversed by copper removal with an ion-
exchange resin.(701, 702) Berna and co-workers have recently reported a system in which
the binding of a guest hydrogen-bond acceptor–donor–acceptor molecule to the thread led
to the trapping of the macrocycle to the linker region between the two stations.(703) Upon
removal, by external complexation, of the guest, interstation shuttling was restored. Control
of shuttling rate has also been achieved by the oxidation/reduction of hydrogen-bonding
stations(704) and by photochemical ring contraction of a macrocycle.(705)

Scheme 24
  

Scheme 24. Complexation of Copper Leads to the Introduction of a Large Kinetic Barrier to
Shuttling(701)

4.2 Physical Models of Degenerate, Two-Station Rotaxanes


A decrease in the rate of shuttling was observed upon lengthening the alkyl chain of 63 as
compared to 62, the magnitude of which corresponds to an increase in activation barrier of
ca. 5 kJ mol–1. This effect was attributed to the increased distance the macrocycle must
travel between binding stations.(693) This process can be modeled by considering the
macrocycle as a particle con ned to a one-dimensional potential energy surface (the
mechanically interlocked nature of the rotaxane preventing signi cant motion of the
macrocycle orthogonal to the thread). Assuming that thread–macrocycle interactions
between stations are negligible leads to the energy pro le shown in Figure 24. The rate of
escape from the station energy well can then be modeled by an Arrhenius equation with a
contribution from a distance-dependent diffusion factor to the overall rate of shuttling. A
quantum mechanical treatment of this system has found that, as the lengthening of the
spacer has no effect on the activation for breaking the hydrogen bonds, the effect on the rate
of shuttling is due to the widening of the overall potential energy well. This leads to a greater
density of states per unit energy, thermal population of a greater number of energy levels,
and thus a larger partition function and activation energy.(706) An alternate explanation of
the dependence of shuttling rate on linker length was proposed by Brouwer and Günbaş who
proposed that shuttling in a similar system was facilitated by hydrogen bonding between the
macrocycle and the station to which it is not currently bound.(707) As the linker length
increases, this bridging conformation becomes increasingly entropically unfavorable.
 Some
 
support for this proposal comes from recent examples from Hirose et al. and Sissel and co-
workers,(708, 709) and Stoddart and co-workers,(710) who in different systems with rigid
spacing units found no correlation between spacer length and shuttling rate.

Figure 24

Figure 24. Idealized potential energy surface for macrocycle shuttling in a degenerate, two-
station molecular shuttle.

4.3 Stimuli Responsive Molecular Shuttles


Molecular motion in mechanically interlocked and thus kinetically stable rotaxanes can be
controlled using multiple binding sites with a nities for the macrocycle that vary under
different conditions. The conditions can be modi ed by electrochemical redox processes,
light, pH, and environmental changes. Once detailed balance is broken by the applied
stimulus, as with temperature difference in Feynmann’s ratchet-and-pawl, useful mechanical
work can be done, provided there is an additional operation to prevent it being undone. By
weakening the existing interaction or increasing the a nity of a competing binding site, the
macrocycle can be driven to shuttle to a new equilibrium position. For a reverse or further
translational motion to take place, a new stimulus is required to perturb the new equilibrium.

4.3.1 Stimuli Responsive Molecular Shuttles with Single Binding Site

Shuttling in a single binding site system can be obtained by reducing the a nity of either the
macrocycle for the station or the station to the macrocycle, by the application of an external
stimuli. Light controlled switching was observed in a rotaxane made from a “bluebox”
macrocycle, an electron-rich dioxyerene station, and redox-active ferrocenyl stoppers.(711)
Light-induced electron transfer from the station to the macrocycle was accompanied by
electron hole transfer to one of the ferrocenyl stoppers, thereby inducing unfolding of the
 
rotaxane and shuttling of the macrocycle away from the station. Photodriven shuttling in
single station rotaxanes was also observed with the increase of steric bulk generated upon
photoisomerization of an azobenzene or stilbene station.(712, 713) Among these examples,
directional shuttling of cyclodextrin (CD) on a symmetrical thread is of particular importance
due to the fact that it serves as a ratchet for a directional bias maintained purely by the
asymmetric nature of the CD rims.(713) Shuttling was biased to the direction that locates the
6-rim of the CD in a position facing the central stilbene unit (Scheme 25).

Scheme 25

Scheme 25. Light-Driven Directional CD Shuttling in a [2]Rotaxane(712)

Solvent-dependent shuttling is generally achieved by switching between hydrogen-bonding


and non-hydrogen-bonding solvents.(714) Hydrogen bonding between macrocycle and
thread is weakened in the presence of a competing solvent. In a peptide-based rotaxane with
a benzyl amide macrocycle, hydrogen-bonding interactions were weakened by addition of
MeOH (Scheme 26).(715) Because of a solvent-induced relocation of the macrocycle, the
chiral center located on the thread now had an in uence on the macrocycle. The original
system could be restored in CHCl3. More recently, a rotaxane based on a pillar[5]arene
macrocycle was shown to shuttle upon either solvent exchange or heating.(716) This system
was found to form supramolecular gels in pure DMSO, and in a similar system this shuttling
was exploited to form a solvent-driven molecular spring.(717)

Scheme 26
  

Scheme 26. Solvent-Dependent Shuttling and Induced CD Response in a [2]Rotaxane(715)

4.3.2 Stimuli Responsive Molecular Shuttles with Two or More Binding Sites

At a given temperature, the macrocycle of a rotaxane distributes itself among the existing
stations according to the binding a nity of each. Perturbation of the binding energies of any
of the stations results in the redistribution of the macrocycle toward a new equilibrium state,
as driven by thermal motion. Changing the position of the macrocycle in a well-determined
way is possible, by making the binding a nity of the less occupied station more favorable
(Figure 25, transformation of station B to B′),(14) or by destabilizing the interaction between
the most occupied station and the macrocycle (Figure 25, transformation of station A to A′).
Stimuli responsive modi cation of the macrocycle can also drive such shuttling processes.
Redox, pH, light, and microenvironment (temperature, solvent, etc.) switches are commonly
used to control translational movement in rotaxane architectures, and will be discussed in
the following sections.

Figure 25
  

Figure 25. Potential energy diagram of a rotaxane-based bistable molecular shuttle.(14)

4.3.2.1 pH-Driven Molecular Shuttles


Variation in pH is one of the most useful stimuli used to drive translational motion in
rotaxanes because hydrogen bonding, electrostatic, and ion–dipole interactions can be
effectively controlled by protonation or deprotonation. Indeed, the rst molecular shuttle
developed by Stoddart et al. depended on acid-driven relocation of the electron poor,
positively charged, cyclophane macrocycle from a protonated benzidine station to a biphenol
station (Scheme 27).(718) Initially, under neutral pH and at 229 K, the macrocycle
preferentially rested on the benzidine station (with 84:16 distribution), stabilized by donor–
acceptor interactions. Upon protonation of this station, a decrease in the strength of the
interaction and an increased electrostatic repulsion caused the macrocycle to relocate to the
other station with a new equilibrium distribution of more than 98:2 in favor of the biphenol
station. Complexation of the cyclophane macrocycle in similar systems has inspired a
variety of computational research.(719-728)

Scheme 27
  

Scheme 27. A pH-Driven Molecular Shuttle(718)

A relatively low temperature was required to obtain an acceptable distribution bias in favor of
the benzidine station under neutral conditions. This necessitated the development of a
system, which exclusively resided on one station under standard conditions. Positional
discrimination with a ratio of more than 98:2 was obtained in a rotaxane composed of a
dibenzocrown ether macrocycle and ammonium-bipyridinium and ammonium-triazole
stations. This interaction has been exploited in many systems.(729-743) In a structurally
similar rotaxane dimer, pH-induced contraction and expansion was made to function as a
“molecular muscle”.(744-746) The pH-dependent relocation of an amide-based macrocycle
from a succinamide to an hydroxyl-cinnamate station was observed, with enhanced
hydrogen bonding between the deprotonated cinnamate derivative and the macrocycle.(747)
Rotaxanes in which repositioning to another station was driven by macrocycle protonation
have been reported.(748, 749) Shuttling of a metal–macrocycle complex from one ligand to
another on a rotaxane as a result of protonation of one of the ligands has been investigated.
(750) The pH-driven shuttling of cucurbit[7]uril between bipyridinium and carboxylate
stations has recently been reported.(751) Attempts were made to couple this motion to an
oscillating pH background reaction, but the pH oscillation was rapidly damped in this case.
Autonomous translocation is an important concept in molecular machine design, but
requires further research.

4.3.2.2 Redox-Driven Molecular Shuttles


Rotaxanes with electron donor–acceptor units or transition metal complexes can be
controlled by redox chemistry, provided that redox potentials of the interacting units are
carefully chosen.(752-756) After the redox process, the products should be stable on the
shuttling time scale, and rapid charge recombination must be prevented. Redox control
  
processes can be chemical, electrochemical, or photochemical.

An early example of redox-mediated shuttling was reported by Leigh et al. and took
advantage of a redox-active naphthalimide, which was introduced at one end of the thread.
The amide-based macrocycle migrated from a succinimide station to the napthalamide
station upon one-electron reduction of the naphthalamide (Scheme 28).(757-759) Increased
electron density on the imide carbonyl of the naphthalamide station outcompeted the
succinimide station, and the macrocycle rested almost exclusively at the naphthalamide
station. The shuttling was reversible and could be performed through photochemical
excitation accompanied by reduction with an external electron donor.

Scheme 28

Scheme 28. A Redox-Driven Molecular Shuttle(757, 758)

Benzidine, tetrathiafulvalene, viologen, naphthalimide, and dioxynaphthalene derivatives have


been widely used as redox-active stations.(760-777) Redox-induced changes in the
coordination preference of transition metal complexes bound to the macrocycle have also
been used to drive biased Brownian motion toward different stations.(778) A tristable
rotaxane has been reported where the position of the macrocycle over each of the three
stations could be controlled solely using electrochemistry.(779)
4.3.2.3 Ion-Driven Molecular Shuttles
  
The presence of ions usually affects the conformation and stability of the macrocycle on a
station by either interfering with the existing hydrogen bonding, ion–dipole, and dipole–
dipole interactions or by sterically disfavoring the binding. This effect can be strong enough
to drive translocation of the macrocycle to another station on the thread. In crown ether-
based macrocycles, metal ions can be chelated by the macrocycle and thus alter the a nity
of the macrocycle for their binding sites, and in some cases induce shuttling. Lithium was
reported to mediate the shuttling of crown ether-based macrocycles from a naphthalimide
station to a pyromellitic diimide, presumably due to stronger ion–dipole interactions with this
station in the presence of a cation.(780) Shuttling could be reversed by the addition of an
excess of [18]crown-6, which sequestered the lithium cation, returning the macrocycle to its
original equilibrium distribution. Rather than interacting directly with the macrocycle or a
station, ions can bind to another position on the rotaxane and affect the a nity of the
macrocycle with either station. Shuttling under similar allosteric control was observed using
a bis(2-picolyl)amine stopper that chelates cadmium ions.(781, 782) Binding to the nearby
station was disrupted in the presence of cadmium, and the macrocycle repositioned itself to
the distal station. Cadmium could be removed by cyanide complexation to reverse the
process.

Chloride coordination to a palladium complex was reported to drive shuttling from a triazole
station to a pyridinium station (Scheme 29).(783) In the absence of chloride anions, the
palladium metal was chelated by the macrocycle with the fourth coordinate site being
occupied by the triazole unit on the thread. Addition of chloride to the solution as its
tetrabutylammonium salt resulted in displacement of the triazole ligand and the macrocycle
detached from the station. When this occurred, the crown-ether moiety of the macrocycle
was free to interact with the pyridinium station, which promoted translocation. The process
could be reversed by addition of AgPF6. Metal-free, anion-induced shuttling between
naphthalimide and triazolium stations was also reported in which shuttling could be
monitored by UV–vis spectroscopy.(784) Iodide addition has been used to induce shuttling in
a halogen/hydrogen-bonding rotaxane.(785)

Scheme 29
  

Scheme 29. Chloride-Induced Molecular Shuttling in a [2]Rotaxane(783)

In separate work, a shift in uorescence was obtained by chloride-induced displacement of a


macrocycle from a central squaraine station to either of two weakly binding degenerate
stations.(786) In a tristable rotaxane bearing urea, ammonium, and phosphine oxide (listed in
descending binding ability) stations on the thread, macrocycle shuttling could be induced by
the binding of an acetate anion to the urea and subsequently to the ammonium site, which
sequentially blocked these higher a nity stations and forced the macrocycle to bind to the
weakly coordinating phosphine oxide station.(669) Similarly, the binding of a macrocycle to a
guanidinium station was found to be inhibited by PO43– addition.(787) In the presence of
Zn2+, the macrocycle preferentially bound the central 2,2′-bipyridyl station. In the presence of
PO43–, but not Zn2+, the macrocycle moved to the weakly coordinating carbamate station.
Copper removal followed by guest complexation between porphyrins in the two macrocycles
has been used to drive shuttling of a [3]rotaxane.(788)

4.3.2.4 Shuttling Induced by Reversible Covalent


Modi cation
Dynamic covalent bond formation and reversible reactions are important tools in the
synthesis of supramolecular devices and machines. This is due to the ability to control their
stability and lability by varying external conditions. These chemical tools enable shuttles to
do work by “compartmentalizing” the macrocycle on a station and acting as a physical
barrier to prevent random Brownian motion reversing the desired translational motion during
re-equilibration steps. This phenomenon will be discussed in detail in the following sections.
Reaction-based shuttling can be achieved in several ways. A station can be blocked by
introducing a new chemical group, or destabilizing the binding of a macrocycle for steric
 or
electronic reasons. The macrocycle can be chemically trapped on one station. Shuttling
between hydrogen-bonding stations has been controlled by Diels–Alder and retro-Diels–
Alder reactions, which block and unblock the better-binding fumaramide station.(789) In
another example, photoinduced heterolytic cleavage of a C–O bond in a diaryl
cycloheptatriene generated a positively charged tropylium station, and thus repelled the
cationic cyclophane macrocycle, displacing it to a different station.(790, 791) Oxidation or
reduction of sulfoxide-based stations and subsequent changes in the hydrogen bonding to a
benzylic amide macrocycle have also been used to control shuttling.(627)

Scheme 30

Scheme 30. Hydrolytic or Entropically Driven Restriction of Shuttling in a [2]Rotaxane(792)

Kawai et al. reported intrarotaxane imine formation between a diamino macrocycle and two
formyl groups at one of the stations. This reaction locked the macrocycle at the station and
maintained a positional discrimination (Scheme 30).(792) Hydrolysis of the imine in wet
CDCl3 in the presence of acid released the macrocycle from the station and also protonated
the amino moieties, making the polyether station more favorable due to hydrogen bonding
and ion–dipole interactions. Shuttling of the macrocycle could also be promoted by heating
and cooling. This entropy-driven shuttling was possible because of the release of two water
molecules on formation of the imine bonds. The entropic favorability of the imine state could
be more easily overcome at lower temperatures by enthalpic contributions.

Scheme 31
  

Scheme 31. Chemically Driven Shuttling in a [2]Rotaxane(793)

As with fumaramides, trans-azodicarboxamide modules have chemical structures


preorganized for strong hydrogen bonding with benzylic amide-based macrocycles. The
interaction of the macrocycle with the trans-azodicarboxamide station is stronger than that
with the succinamide station. Induced shuttling of a macrocycle to the less favored
succinamide station could be achieved chemically by blocking the favored station by
reaction with triphenylphosphine (Scheme 31).(793) This bulky substituent drove the
macrocycle away from the station. Under subsequent Mitsunobu reaction conditions, in the
presence of a carboxylic acid and an alcohol, the triphenylphosphonium was cleaved from
the station as triphenylphosphine oxide. After oxidation, the macrocycle returned to its initial
position.

4.3.2.5 Photodriven Molecular Shuttles


Photoinduced shuttling is an attractive mode of control over translational motion due to the
ease of stimuli introduction and removal. Moreover, if the back reaction or operation is
spontaneous and does not require an additional input, then the overall process becomes
autonomous; shuttling will occur as long as energy is supplied in the form of light. However,
instead of a continuous light source, ashes of light must be used to prevent the system
reaching and staying at a steady-state distribution. Although photodriven shuttling can be an
e cient approach, the kinetics of translation and photochemical processes have to be
carefully considered. If the photoinduced process involves a redox reaction, then rapid
charge recombination can take place before positional displacement of the macrocycle. To
avoid this, either excited states with a long lifetime are required or external reagents must be
used to reduce or oxidize the species.

In a carefully designed shuttle reported by Stoddart et al., translational switching was


obtained by one-electron reduction of a viologen station by a ruthenium trisbipyridine stopper
(Figure 26).(794) Relatively slow back electron transfer permitted ca. 10% of the macrocycle
to reposition to the dimethyl viologen station on each electron transfer. Continuous
irradiation maintained a 95:5 distribution of the macrocycle in favor of the dimethyl viologen
station. The system equilibrated back to the original viologen position once irradiation
ceased with spontaneous back electron transfer allowing reversion.
Figure 26
  

Figure 26. Photoinitiated redox-driven shuttling in a [2]-rotaxane initiated by (i) irradiation and
(ii) subsequent reduction of viologen station. (iii) Competing back electron transfer from one-
electron reduced viologen to Ru3+. (iv, v) With continuous irradiation the macrocycle shuttles
to the dimethylviologen station. (vi) Ceasing illumination restores the macrocycle to its
original position.(794)

Alongside the photoinduced redox reactions discussed above, improved excited-state


hydrogen-bonding interactions(795) and the photoisomerization of stilbene,(796)
azobenzene,(797) and other ole nic moieties leading to altered hydrogen bonding or steric
bulk creation have also been used to drive shuttling processes.(798) Absorbance and
uorescence has been used to monitor the progress of shuttling along the rotaxane in some
of these systems.(16, 799-801) Likewise, other photochromic compounds such as
spiropyrans have been exploited to control molecular shuttles, by switching between
photoisomers that form hydrogen bonds of different strengths.(802)

4.3.2.6 Molecular Shuttles Driven by Changes in the


Microenvironment
The term “microenvironment” usually refers to the temperature and solvent system. The
easiest way to control shuttling via solvent choice is through the disruption of hydrogen-
bonding interactions between the macrocycle and the station. Several examples of
macrocycle relocation induced by a change in solvent polarity or rotaxane solvation have
been reported.(803) In some cases, solvent molecules facilitate a shuttling process already
driven by other stimuli.(698, 797, 804) For entropy-driven shuttles, the rearrangement of
solvent molecules in or around the macrocycle is an important parameter controlling the
shuttling process.   
Rotaxanes bearing more than two stations and even polymeric rotaxanes driven by these
stimuli have been reported.(797, 805, 806) In these shuttles, work is undone in each reverse
shuttling. Directional shuttling via a ratchet mechanism has been achieved and will be
discussed in the next section.

4.4 Compartmentalized Molecular Machines


Stimuli-responsive molecular shuttles are rotaxanes in which the net position of the
macrocycle on the thread is controlled by external triggers (light, heat, electrons, chemical,
pH, binding events, etc.). In general, these triggers change the properties of the binding sites
and subsequently have an in uence on macrocycle/thread interactions. In this section,
various stimuli-induced mechanisms to create and control nonequilibrium conformations will
be presented. Rotaxane succ1-74 has two identical stations, which are distinguishable due
to the differing stoppers. The thread is divided into two compartments by a bulky silyl ether
barrier, preventing the macrocycle from shuttling between the two stations.(75) By removing
the silyl ether group, dynamic exchange of the macrocycle was enabled and the system
moved toward equilibrium. This resulted in a 50:50 distribution of the macrocycle between
the two stations. The operation is irreversible, because the attachment of the bulky silyl
group again does not change the average position of the macrocycle, and will not restore the
initial distribution of the macrocycle; the position of the macrocycle is not determined by the
state of the machine. The machine starts in a statistically unbalanced state (the bulky silyl
group prevented the macrocycle restoring balance by moving toward an equilibrium
distribution). By “linking” the system, removing the bulky group, the machine could reach its
equilibrium state, because free movement of the macrocycle was no longer prevented by the
barrier. Therefore, the macrocycle had “escaped” to a lower energy distribution (Figure 27).

Figure 27
  

Figure 27. [2]Rotaxane 74 acts as an irreversible mechanical switch. The silyl ether is too
bulky to allow macrocycle shuttling between the two succinamide stations.(75)

Recently, Coutrot and co-workers synthesized a two-compartment molecular machine


consisting of a dibenzo-24-crown-8 macrocycle and a thread with anilinium and
monosubstituted pyridinium amide stations.(807) The rotaxane was synthesized via CuAAC
chemistry. Subsequent N-benzylation of the triazole moiety, which is located in the middle of
the thread, was found to serve as a barrier as well as a third station for the macrocycle.
Depending on which station the macrocycle was located, the N-benzylation of the triazole
allowed trapping of the macrocycle in either the left or the right compartment.

The Leigh group has extended their initial system using the same design described above,
but with two different stations 75 (a fumaramide station and succinamide station) (Figure
28).(75)

Figure 28
  

Figure 28. Operation of a compartmentalized molecular machine, which corresponds to a


two-state Brownian ip- op. Operation steps: (a) Desilylation (80% aqueous acetic acid); (b) E
→ Z photoisomerization (hν at 312 nm); (c) resilylation (TBDMSCl); and (d) Z → E thermal
isomerization (catalytic piperidine).(75)

In this system, the exchange of the macrocycle between stations could be controlled by the
introduction of a barrier in the middle of the thread (kinetic control) as well as by altering the
binding a nity of the macrocycle to the two different stations (thermodynamic control).
Initially, 85% of the macrocycle bound the fumaramide station and 15% the succinamide
station. The system was still unlinked because of the kinetic barrier and therefore not in
equilibrium. A balance-breaking stimulus (photoisomerization at 312 nm) generated a 49:51
E:Z photostationary state. Removal of the barrier, the linking stimulus, allowed the balance to
be restored by the macrocycle equilibration. Restoring the barrier unlinked the system. The
last step was Z → E ole n isomerization, which made the system statistically unbalanced,
unlinked, and not in equilibrium. After one operational cycle of the machine, 56% of the
macrocycles were located on the succinamide station. The thread had therefore changed the
net position of the macrocycle. Because the succinamide station binds the macrocycle less
strongly than the fumaramide station, this process has moved the macrocycle energetically
uphill; that is, it has performed a ratcheting operation, transporting a particle away from
equilibrium. This system and its function is an experimental realization of the transportation
task required in Smoluchowski’s trapdoor and Maxwell’s pressure demon, which are
discussed in the introduction (section 1) and are powered by chemical and light energy.
Figure 29 shows a schematic representation of the four sequential steps performed by
rotaxane 75 to transport the system energetically uphill: balance-breaking 1, linking,
  
unlinking, balance breaking 2 (resetting the machine but not the substrate).

Figure 29

Figure 29. Initially the system is balanced (in proportion to the sizes of the two
compartments), with an equal density of particles in the left and the right compartments. By
changing the volume of the left-hand compartment, the system becomes statistically
unbalanced. Opening the door allows the particle to redistribute according to the new size of
the compartments. Closing the door ratchets the new distribution of the particles. Restoring
the left compartment to its original size results in a concentration gradient of the Brownian
particles across the two compartments. Here, the size of the compartment represents the
energy level of the macrocycle-station system.(75)

There are many different types of compartmentalized molecular machines, some of which
have been described in this section. Three main machine types exist. First is the two-state
(or multistate) Brownian switch, which is a machine that can reversibly change the
distribution of a Brownian substrate between two distinguishable sites as a function of state
of the machine. It does this by biasing the Brownian motion of the substrate (section 4.3).
Second is a two-state (or multistate) Brownian ip- op, a machine that can reversibly change
the distribution of a Brownian substrate between two (or more) distinguishable sites and that
can be reset without restoring the original distribution of the substrate. The distribution of
the substrate cannot be determined from the state of the ip- op, but rather it is determined
by the history of the operation of the machine. Rotaxane 75 is an example of a two-state
Brownian ip- op. Third, a Brownian motor is a machine that can repetitively and
progressively change the distribution of a substrate while the machine can be reset without
restoring the original distribution of the substrate. The rotary motors designed by Feringa
and co-workers are examples of this category of device. Additional examples will follow in
section 4.5.
Brownian ratchet mechanisms fall into two general classes: energy ratchets and information
ratchets (section 1). The above rotaxane is an energy ratchet in which the energy minima
 
and maxima of the potential energy surface are varied without regard to the particle’s
location. In the next section, examples of the second class of Brownian ratchets, information
ratchets, will be explored where a barrier is raised or lowered according to the position of a
Brownian particle on a potential energy surface, resulting in the particle distribution being
directionally driven away from equilibrium.

4.4.1 Molecular Information Ratchets

An information ratchet, such as Maxwell’s pressure demon, requires knowledge of the


position of each particle and is able to open the door if the particle is approaching from a
certain direction. This positional sorting allows the accumulation of particles in one side of
the container and thus results in a pressure gradient. The rst example of a molecular
information ratchet was described by Leigh and co-workers (Scheme 32).(808) Molecular
machine 76 consisted of a dibenzo-24-crown-8-based macrocycle, which was mechanically
locked on a linear thread by bulky 3,5-di-t-butylphenyl stoppers. An α-methyl stilbene unit
divided the thread into two compartments, both with ammonium binding stations. The
rotaxane used photosensitized energy transfer from the macrocycle to the stilbene unit as
the key step in changing the macrocycle distribution. When the stilbene unit adopted the E
form, the macrocycle could move randomly along the full length of the thread by Brownian
motion, while when the Z form is adopted, the macrocycle was trapped in one or the other of
the two compartments. To keep the machine as the Z-isomer, it was operated in the
presence of the photosensitizer PhCOCOPh. The sensitized photostationary state of α-
methyl stilbene is 82:18 Z:E. Selective “gate opening” was achieved by photosensitized
energy transfer from the macrocycle (which has a benzophenone (PhCOPh) moeity
attached) to the stilbene unit. This led to a 55:45 Z:E ratio of the α-methyl stilbene. As energy
transfer was distance dependent, this photosensitized transfer was more likely to happen
when the macrocycle was in the blue compartment (Scheme 32). When the macrocycle was
on the green station, intramolecular energy transfer from the macrocycle was not e cient
and the gate stayed closed, biasing macrocyclic distribution. This is an example of using the
positional information on a Brownian particle to break detailed balance, driving macrocycle
distribution away from equilibrium.

Scheme 32
  

Scheme 32. Structure and Mechanism of the Information Ratcheta

Scheme a (a) Gate closed, but energy transfer from the macrocycle is e cient. (b and c)
Gate is open, allowing free shuttling of the macrocycle. (d) The macrocycle is on the green
station, intramolecular energy transfer (ET) from the macrocycle is ine cient; intermolecular
energy transfer from PhCOCOPh dominates (closing the gate).(808)

An alternate approach to a molecular information ratchet using chemical fuel to directionally


transport a macrocycle was later published by Leigh and co-workers (Scheme 33).(809) In
this example, a phenyl ester attached to a chiral carbon center separated two degenerate
fumaramide binding sites. The close proximity of the stations to the hydroxyl bearing carbon
enabled the chiral center to in uence the position of the macrocycle. At equilibrium, the
macrocycle rested equally at either degenerate station. When a barrier was introduced by
benzoylation with an achiral dimethylamino pyridine (DMAP) catalyst, an equal distribution of
products was obtained. However, when the chiral DMAP-based catalyst (S)-79 was used, a
33:67 distribution in favor of the pro-(S)-fumaramide (S)-80 station was obtained. Using
catalyst (R)-79 resulted in a reverse bias in the position of the macrocycle. Because the
stations were identical, macrocycle relocation resulted in a decrease in entropy, with identical
enthalpic interactions, and thus a ratcheting operation was achieved. Motion away from an
enthalpically favored location was achieved in a similar rotaxane, but with fumaramide and
succinamide stations of different binding strengths. Using the same chiral catalyst, 15% of
the molecules were relocated to the enthalpically unfavored succinamide station.
Scheme 33
  

Scheme 33. Chemically Driven Molecular Information Ratchet(809)

This concept was further developed in the three-compartment, chemically driven, molecular
information ratchet 82.(810) Here, the compartments contained fumaramide groups and
were separated by hydroxyl groups (Scheme 34). The macrocycle could be e ciently trapped
at either end of the track by sequential benzoylation of the hydroxyl groups in the presence of
the chiral catalyst 83 resulting in a 1:21:79 distribution of the macrocycle. The steric barriers
formed prevented the macrocycle from passing, and trapped the ring in a certain
compartment. It was shown that the macrocycle had an in uence on the rate of
benzoylation of the hydroxyl groups on the thread resulting in acylation taking place
preferentially far from the macrocycle. The distributional bias could be reversed with the
enantiomeric catalyst 83, and when both chiral catalysts were used the macrocycle
preferentially resided in the central station with a 10:77:13 distribution.

Scheme 34
  

Scheme 34. Directional Transport of a Macrocycle in a [2]Rotaxane Three-Compartment


Chemical Information Ratchet

Directional translational motion along the axis of pseudorotaxane has also been reported
(Figure 30).(553, 811, 812) Threading and dethreading of the macrocycle could be controlled
by photochemical, chemical, or electrochemical redox reactions. The pseudorotaxane axis
has three different chemical motifs responsible for directional translocation. The rst is a 2-
isopropylphenyl group, a neutral energy barrier at one side of the axle. An electron-rich 1,5-
dioxynaphthaline occupied the center, and at the other end was a positively charged 3,5-
dimethylpyridinium station. A tetracationic electron-poor cyclobis(paraquat-p-phenylene)
macrocycle formed a π-complex with the electron-rich 1,5-dioxynaphthaline group. To form
the complex, the macrocycle had to pass one of the energy barriers, and the neutral 2-
isopropylphenyl moiety presented a lower energy barrier. Hence, threading took place
selectively from this side of the axis. Upon one-electron reduction of the macrocycle,
interaction with the 1,5-dioxynaphthaline station was weakened. Additionally, as a result of
the reduced positive charge on the macrocycle, electrostatic repulsion from the pyridinium
site became less potent, allowing the macrocycle to pass over this barrier with greatly
reduced activation energy. Hence, once reduced, the macrocycle dethreaded over the
pyridinium group. The macrocycle threaded and dethreaded in a directional manner via a
ratcheting mechanism ( ashing ratchet) acting on the potential energy surface of the
pseudorotaxane axle. The same group later reported a ashing ratchet based on similar
principles, in a pseudorotaxane with one stopper, where threading was driven by reduction,
and dethreading slowly occurred after rapid shuttling away from a reoxidized bipyridinium
station.(813) Recently, Credi et al. also reported the photodriven directional   
threading/dethreading of a crown ether macrocycle on an axle using a ashing ratchet
mechanism.(814, 815)

Figure 30

Figure 30. Photodriven directional translational motion in pseudorotaxanes 86–87. Reprinted


with permission from ref 811. Copyright 2013 American Chemical Society.

4.5 Controlling Rotational Motion in Mechanically Bonded


Structures
An alternative way to extract work from mechanically interlocked structures would be to
control the relative rotation of their components. The rate of macrocycle pirouetting can
easily be controlled through temperature, electric eld strength, light, structural freedom, or
by altering its interaction using other chemical moieties or solvents.(564, 698, 816-819)
Nevertheless, controlling the directionality of this movement requires careful design.

4.5.1 Controlling Rotational Motion in Rotaxanes

The pirouetting frequency of a macrocycle around a thread depends strongly on the strength
of the interactions between the macrocycle and the thread (and environment), which must
be broken and reorganized during the movement of the macrocycle. The rate of macrocyclic
pirouetting in some hydrogen-bonding rotaxanes can be controlled by changing the strength
of an electric eld, which alters the strength of the hydrogen bonds.(820-824) In some cases,
direct manipulation of macrocycle/axle interactions can either restrict rotation or increase its
frequency by weakening intracomplex interactions. The formation of weaker hydrogen bonds
with maleamide units as compared to fumaramide units has been discussed in previous
sections. Fumaramide preorganizes two hydrogen-bonding motifs to bind the macrocycle,
leading to an increase in binding strength. On the other hand, the cis-form of this 
ole nic
 
structure can only form hydrogen bonds with one site of the macrocycle. As a consequence,
the station–macrocycle interactions are weaker, and thus rotation was observed to increase
in frequency by 6 orders of magnitude on photoisomerization.(825)

Figure 31

Figure 31. Effect of water on the rate of pirouetting of a macrocycle about an axle. Reprinted
with permission from ref 698. Copyright 2013 Nature Publishing Group.

Like photoisomerization-dependent disruption of hydrogen bonding,(825) solvent choice can


also play a role in controlling these noncovalent interactions and hence in uence the rate of
pirouetting in a rotaxane system.(826) Comparison of the rate of exchange between axial
(E1) and equatorial (E2) protons, which was determined by the pirouetting rate, showed that,
as compared to heating in dry pyridine, the exchange was greatly accelerated when the
pyridine contained 5% D2O (Figure 31).(698, 827) Addition of water sped up redox-driven
shuttling in a different rotaxane system by hydrogen-bonding dependent “lubrication” of the
mobile elements. Interestingly, other protic solvents (alcohols, nitromethane) had less effect
on the pirouetting rate. This was attributed to the ability of water to form 3-D hydrogen-
bonding networks. The effect on macrocycle–thread interactions of the sequential addition
of single methanol molecules has been probed by IR spectroscopy of rotaxane-solvent
clusters.(828, 829)

Besides hydrogen-bonding, cation-induced restraint of motion in a crown ether-based


macrocycle has been reported with a corresponding increase in pirouetting rate upon
demetalation.(817, 818, 830-834) Redox-dependent switching between coordinating ligands
on a macrocycle has been used to control rotational orientation of a macrocycle about a
rotaxane axle.(818) In a hybrid organic–inorganic rotaxane, hydrogen bonding between
uorines on the macrocycle and ammonium stations resulted in slow shuttling but fast
rotation.(835, 836) This rapid rotation was attributed to the array of hydrogen-bonding
interactions formed in which a new bond had already started to form before the original was
completely broken.
4.5.2 Controlling Rotational Motion in Catenanes
  
The points discussed above for rotaxanes also apply to rotational motion in catenanes. In a
catenane, the rate of rotation of one macrocycle with respect to the other depends on the
strength of interactions between the two. Therefore, any stimuli altering this interaction
should result in a change in rotational rate. The solvent dependence of the rate of
macrocycle rotation in catenanes containing benzylic amides has been measured by NMR
analysis and also by AFM-based single molecule measurement techniques.(837) In the
second method, catenanes with different intercomponent hydrogen-bonding ability were
attached to poly(ethylene oxide) polymers and analyzed in dimethylformamide (DMF) and
tetrachloroethane (C2H2Cl4). As a measure of equilibrium conformational entropy, the
persistence length of the polymer was analyzed for two different compounds in two different
solvents to estimate the restriction of motion of the macrocycle in the catenane. In
agreement with the NMR analysis, in a single molecule experiment using the AFM method,
the mobility of the macrocycle was shown to be accelerated by polar solvents or by
disrupting the hydrogen-bonding ability by chemical modi cation.

Ion exchange,(838-840) pH,(841-844) redox,(845-856) demetalation,(842, 857-869) light or


redox-mediated ligand exchange of metals,(853, 855, 870) photochemical switching,
(870-878) photoisomerization-dependent sequestering of macrocycle(879) and solvent,
(826, 880-882) photoisomerization-dependent change in hydrogen-bonding potency,(883)
and the self-assembly of liquid crystal phases(884) have also been reported to control the
rotation of the macrocycle in catenanes. However, directional rotation of one macrocycle
with respect to another in a catenane structure has rarely been investigated. Directional
rotation was achieved with a [3]catenane in which the presence of a third macrocyclic
component helps to restrict the rotational freedom of the molecule (Figure 32).(883) The
large macrocycle on which the two small macrocycles rotate bears four different interaction
sites: two fumaramide motifs with different binding a nities (light green and red, the
methylated station has a lower a nity for steric reasons), one succinic amide ester (orange),
and nally an amide group (dark green). A benzophenone unit is attached close to the rst
fumaramide station to enable selective photosensitized isomerization of this ole nic group
through energy transfer using a higher wavelength than required for nonsensitized
isomerization.

Figure 32
  

Figure 32. Directional circumrotation in a [3]catenane. (i) hν (350 nm), (ii) hν (254 nm), (iii)
heating; or heating with catalytic ethylenediamine; or catalytic Br2, hν (400–670 nm).
Adapted with permission from ref 883. Copyright 2003 Nature Publishing Group.

Initially, in the absence of any stimuli, one of the macrocycles (blue) rests on the most
favored fumaramide station, while the second macrocycle (purple) resides on the second
most favored, the methyl-fumaramide station (Figure 32). Upon photosensitized
isomerization of station A, hydrogen-bonding interactions between the blue macrocycle and
the station weaken, and because the second most favored site is already occupied by the
purple macrocycle, the blue macrocycle moves to the third most favored, succinimide amide
ester station C. Further photoisomerization of the methyl-fumaramide station triggers the
shuttling of the purple macrocycle to the nal binding site, station D. The fumaramide
stations are converted to the cis-isomers either by heating in the presence of a catalytic
amount of ethylenediamine or via irradiation at 400–670 nm in the presence of a catalytic
amount of Br2. This time, however, the relative orientations of the macrocycles force
repositioning to the newly formed fumaramide stations in the opposite, with respect to the
initial, distribution. A second round of stimuli is necessary to fully reset the system and to
bring the macrocycles back to their initial positions. The combination of control of station
binding a nity and macrocycle position allowed directional rotation in this catenane.

Scheme 35
  

Scheme 35. Directional Circumrotation in [2]Catenane 91(885)

A directional [2]catenane rotary motor has also been reported with a simpler
chemical/photoinduced mode of operation and well-de ned ratcheting mechanism.(885) In
this example, the balance breaking operation (to create a new equilibrium distribution) was
photoisomerization of the fumaramide station. However, the macrocycle’s rotation to the
new energy minima was blocked by two orthogonal protecting groups: an acid labile trityl
and base labile tert-butyl-dimethylsilyl. Selective deprotection of one of these protecting
groups allowed the movement of the macrocycle to the succinamide station in only one
direction. For clockwise rotation, trityl deprotection was required. Under equilibrium
conditions, the alcohol had to be protected once more to avoid the work being undone. Full
rotation was attained by deprotection and reprotection of the tert-butyldimethylsilyl moiety
with reisomerization of the fumaramide station. The sequence of deprotection–reprotection
reactions could be reversed to achieve counter clockwise circumrotation. The repositioning
of the macrocycle to its new equilibrium position was via only one of two degenerate
pathways, and thus the system acts as a directional molecular rotor (Scheme 35).

5 Molecular Level Motion Driven by External Jump To

Fields

In most synthetic molecular machines described to date, control over motion arises from the
selective restriction of Brownian uctuations through control of steric and noncovalent
bonding interactions (via manipulation of chemical structure). The application of external
elds can cause the bulk movement or orientation of a molecular species with electric eld-
directed alignment of liquid crystals being the most important technological application.
Several research groups aim to use electromagnetic elds to control submolecular motion,
with most examples concentrating on generating submolecular rotary motion. Molecular
  
rotation involves passage over the minima and maxima of a torsional potential energy
surface. An external eld can either induce an excited state, where the torsional potential
energy surface is altered, or interact with a permanent or induced molecular dipole and so
orient the molecule in a particular direction. The interaction between the eld and the rotor
provides the energy necessary to surmount kinetic barriers and overcome energy dissipation
and thermal randomization. To date, molecular dynamics simulations of molecular rotatory
systems have dominated the eld.(886-930) Only a few examples exist where theoretical
studies have led to the synthesis of rotors and their examination under the in uence of an
electric eld.

A recent example is the simulation of a single-molecule electric revolving door based on a


phenyl-acetylene macrocycle published by Hsu, Li, and Rabitz.(931, 932) The authors found,
based on simulations, that the opened and closed-door states of 92, whose exchange was
accompanied by signi cant conductance variation, could be operated by an external eld.
Furthermore, they proposed that due to the large on–off conductance ratio, the molecular
machine could also serve as an effective switching device (Figure 33).

Figure 33

Figure 33. Schematic representation of an electric revolving door. (a) Door closed-switch on
leading to high conductance. (b) Door open-switch off leading to low conductance.(932)

An earlier example of a computational investigation was published by Ratner and Troisi.


(933, 934) Compound 93 bound between two electrodes exhibited motion under the
in uence of an electric eld (Scheme 36). The authors suggested that the system could nd
applications in molecular electronics and could be used to create switchable molecular
junctions. Theoretical studies revealed that the conformers of 93 (which could be
interchanged by an external eld) also showed differences in conductance across the
junction and could thus be used as a conformational molecular recti er.(935-937)
Scheme 36
  

Scheme 36. A Molecular Recti era

Scheme a (a) High conductance predicted. (b) Low conductance predicted.(934)

Fujimura and co-workers undertook a series of theoretical studies into the mechanism of
rotation in gas-phase directional chiral motors driven by picosecond pulses of a linearly
polarized laser.(938-950) In one study, aldehyde 94 was examined as a chiral molecular
motor with the formyl group as the rotor (Figure 34). They considered both enantiomers in
their studies and came to the conclusion that directional motion originated from the
asymmetric potential-energy surface of the chiral molecules and time-correlated forces
created by laser-permanent dipole interactions.(940) The process is reminiscent of a rocking
Brownian ratchet (section 1.6.2), but here thermal energy is not required to surmount kinetic
barriers.(951)

Figure 34
  

Figure 34. One enantiomer of chiral molecule 94, in which directional rotational motion was
driven by linearly polarized laser pulses and studied by quantum and classical mechanical
simulations.(940)

Several examples of surface-mounted molecular rotors have been reported where motion
could be driven by alternating electric elds or the absorption of pulses of light.(952-962)
Michl and co-workers synthesized, on the basis of computational studies, a series of
surface-mounted rotors.(963-965)Figure 35 shows the structure of polar and nonpolar
versions of their altitudinal rotor. Polar rotor 96 consisted of a 9,10-dihydrophenanthrene
substituted by four uorine atoms on the central ring and had a calculated dipole moment of
3.7 D. The nonpolar rotor was a 4,5,9,10-tetrahydropyrene whose D2 symmetry precluded a
dipole moment. X-ray photoelectron spectroscopy (XPS), scanning tunneling microscopy
(STM), and IR spectroscopy showed that for a fraction of the molecules, the static electric
eld from an STM tip induced a change in the orientation of the polar rotor, but not in the
nonpolar analogue.(966) Several molecular dynamics simulations studying the in uence of
an electric eld or a uid ow on the rotation of a molecule xed on a surface have been
published by the same group.(967, 968) One recent example simulated carborane-based
molecular propellers and showed that they could be successfully driven at GHz rates by an
oscillating electric eld or by a ow of gas.(969)

Figure 35
  

Figure 35. Nonpolar 95 and dipolar 96 altitudinal rotors mounted on an Au(111) surface.
Note that rotor and anking aryl rings are arbitrarily shown perpendicular to the surface for
clarity.(966)

Molecular machines in the solid state and condensed phase will be discussed in more detail
in section 8.6.1. A number of research groups have been working in the eld of crystalline
molecular machines with the goal of creating new materials with interesting properties and
that are responsive to external stimuli such as external electric elds.(970-975)
Amphidynamic crystals are a form of condensed-phase matter with anisotropic molecular
order and controlled dynamics, and they offer a good platform for the design of these new
materials.(976) Molecular rotors are one of the most promising molecular structures for the
synthesis of amphidynamic crystals. Structural designs analogous to macroscopic
gyroscopes and compasses are one possibility in the design of such molecular rotors.
Several examples have been published consisting of a rotating unit linked to a shielding box
or stator by an axle.(977-980) The shielding box generates the local free volume required for
unhindered rotation in an otherwise densely packed environment. The study of the
orientation of and dipole–dipole interactions in dipolar rotor arrays are important in
understanding the dielectric properties of these materials and the dynamics of the dipolar
rotors. Furthermore, materials containing dipolar rotors can be controlled and reoriented by
external electromagnetic elds or optical stimuli.(981, 982) Examples of molecular
gyroscopes have been reported by Gladysz et al., who developed transition metal-based
systems in which trans phosphorus donor atoms are bridged by three methylene chains or
related linkers. For example, the system shown in Figure 36 contains either a Fe(CO)3 or a
Fe(CO)2(NO)+ rotor. The latter possesses a dipole moment, which could be a possible handle
for external control.(983) A series of studies have been carried out by the same group using
gyroscope-like platinum and palladium complexes with trans-spanning bis(pyridine) ligands.
(984)

Figure 36

Figure 36. (a) Molecular structure of transition-metal-based gyroscopes 97 and 98. (b) X-ray
structure of compound 97. X-ray crystal structure reprinted with permission from ref 983.
Copyright 2004 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
  
Molecular gyroscopes with two bulky stators such as triarylmethyl or triptycyl groups,
creating space for a 1,4-phenylene, have been reported.(985-1002) Some of these wheel and
axle motifs (see also section 8.6.4) showed a dielectric response as a function of external
alternating electric elds.(981, 1003)

6 Self-Propelled Nanostructures Jump To

The mechanical self-propelled motion of microscopic objects such as bacteria and the
motion of camphor or soap “boats” driven by interfacial forces have long fascinated
scientists. More recently, focus has shifted to the control of nanoscale motion in synthetic
systems.(1004, 1005) These systems are driven not by thermal energy, but by other
mechanisms more relevant on the nanoscale. Motion may be generated either
autonomously when the energy required for movement is continuously supplied or
nonautonomously.

6.1 Propulsion by Manipulation of Surface Tension


If surface tension is not balanced between two sides of a droplet of liquid or a bubble of gas,
directional transport may occur. This is known as the Marangoni effect and is observed in
several natural systems exhibiting spontaneous ow.(1006, 1007) Temperature gradients
have typically been used in arti cial systems to generate Marangoni ows.(1008-1014) A
droplet containing a species that absorbs to the surface, irreversibly modifying the surface
energy thereof, has also been used to illustrate this effect in a “chemical Marangoni effect”.
(1015-1018)

The autonomous motion of millimeter scale objects on a liquid surface driven by the
catalytic decomposition of hydrogen peroxide was reported by Whitesides et al.(1019) A
small area of platinum near the edge of each disc catalyzes this decomposition and releases
oxygen bubbles, which drive the movement by a recoil force. Others, however, observed the
motion of platinum–gold rods toward the oxygen producing platinum end.(1020-1025) The
contributions of driven and Brownian motion could be distinguished at aqueous/organic
interfaces,(1022) and a rotor based on the same principles has also been synthesized.(1026)
A similar system was reported by Manners and co-workers.(1027) Ozin et al. reported
platinum–silica sphere dimers, which showed quasi-linear motion in bulk solution and
rotational motion at the solution–glass interface.(1028) Rotors of decreased size and
increased rotational rate have since been reported,(1029) both once more based on
hydrogen peroxide decomposition.(1030) The catalytic domain has been scaled down to a
molecular level by Feringa et al.,(1031) who utilized a surface bound dinuclear manganese
catalase analogue.

Recently, Fischer et al. reported a self-propelling nanoparticle of only 30 nm diameter, a size


comparable to certain enzymes, based on a Pt–Au system.(1032) Feringa and co-workers
reported the use of carbon nanotube bound enzymes in autonomous propulsion.(1033) As
an alternative to the ubiquitous use of hydrogen peroxide to propel nanoparticles,hydrogen
 
evolution by a magnesium domain has been exploited to drive autonomous motion.(1034)
Finally, Sen et al. reported the use of ring-opening metathesis polymerization using a
modi ed Grubbs’ catalyst to propel a nanoparticle. Motion was generated by the
consumption of monomer only on one side of the nanoparticle, that where the catalyst is
bound. This creates a concentration gradient across the nanoparticle and thus movement.
(1035)

6.2 Mechanical Self-Propulsion


In contrast to most synthetic, surface tension driven, self-propelled systems,
microorganisms almost exclusively utilize mechanical self-propulsion, akin to a swimming
motion or the corkscrew motion of a boat’s motor. The low Reynolds number at the
nanoscale limits the number of swimming mechanisms that are viable,(1036) as any useful
motion must involve a nonreciprocal motion to break the time reversal symmetry.
(26, 1037-1044) Various solutions have been proposed (Figure 37).(1041, 1045-1050) The
key point in each is that they must possess at least two hinge points; no system with a sole
hinge point can be nonreciprocal. The sequential use of each hinge point allows motion at
low Reynolds number.

Figure 37

Figure 37. (a,b) Proposed designs for swimmers viable at the nanoscale. (c,d) Potential
molecular solutions.(14)
No molecular solutions have yet been reported. However, a microscopic arti cial swimmer
has been reported.(1051) A major di culty in realizing any molecular design would be  
carrying out the driven motion rapidly and frequently enough to be observable against
random Brownian motion. The repetitive, directional, rotation of a chiral molecule is also a
nonreciprocal motion. The eld-driven directional motion of an uncon ned chiral molecular
rotor would result in its propulsion via a screw-propeller type mechanism, such as the
bacterial agella system. Such a system has been theoretically studied as a means to
spatially resolve a racemic mixture of propeller-like molecules.(1052, 1053)

7 Molecular-Level Motion Driven by Atomically Jump To

Sharp Tools

The development of single molecule imaging techniques such as atomic force microscopy
(AFM), scanning tunneling microscopy (STM), and optical and magnetic tweezers has
signi cantly enhanced our understanding of the mechanical properties and working
mechanisms of molecular motor proteins.(1054-1068)

Instead of average statistical information obtained from an ensemble of species, these


techniques allow direct measurement of molecular level forces, mechanical properties and
motions such as rotation, gliding, and translation, pivoting of an individual molecule,
supramolecular host–guest exchange, and coconformational changes within interlocked
molecular assemblies.(1061, 1069-1089) Beside their imaging abilities, the use of such tools
to drive molecular level motion has also been explored.

Molecular scale motion is driven by different forces than in the macroscopic world. Gravity is
essentially irrelevant at low Reynolds number, and frictional forces differ greatly.
(606, 1090-1101) The driving forces/interactions must be su cient to overcome signi cant
thermal uctuations. For the desired motion to be obtained (i.e., translational motion via
rotation of the wheels of a “nanocar” instead of gliding), surface–molecule interactions
should be neither too strong nor too weak to balance dissociation and immobility.

STM-driven positional displacement of xenon atoms at low temperature was reported by


Eigler et al.(1102) The rotation of oxygen and acetylene molecules adsorbed on Pt(111) and
Cu(111) surfaces, respectively, was shown to be induced by the tunneling of electrons from
an STM tip at low temperatures.(1103, 1104) STM- or AFM-driven motions of aromatic
molecules on surfaces have been investigated extensively.(1105-1116) The STM-driven
directional diffusion of 9,10-dithioanthracene over a Cu(111) surface has been reported.
(1117) STM-induced translational motion of porphyrins along the voids of a porphyrin
monolayer on a Cu(100) surface was attributed to the rotation of a t-butyl substituent.(1113)
Displacement to a region with less structural restriction led to more frequent rotation, and
thus the STM probe mediated a switch between a rotating and a nonrotating porphyrin.
Cyclic molecules and uorinated C60 derivatives were also shown to roll across surfaces.
(1118-1120)
Translational motion mediated through rolling structural units (wheels) has been reported in
a number of independent publications. Initial attempts to develop a “molecular wheelbarrow”
  
resulted in overly strong surface–-molecule interactions, and the STM tip caused
fragmentation.(1121-1124) STM-driven rolling of a single molecule deposited on a Cu(111)
surface was rst achieved by Grill et al.(1125) A chemically switchable metalloporphyrin
pinwheel was reported by Lambert et al.(1126) The reversible attachment of two CO2
molecules to a diffusing anthraquinone on a Cu(111) surface was an important example of
the potential of arti cial molecular cargo transporters.(1127) Chiaravalloti et al. designed the
rst molecular system acting as a rack and pinion device, using self-assembled hexa-tert-
butyl-pyrimido-pentaphenylbenzene on Cu(111).(1128) Interlocked arrays formed by self-
assembly allowed the rotation and translocation of a single molecule along the array by
manipulation with an STM tip.

Figure 38

Figure 38. (a) Chemical structure and (b–f) STM micrographs of translational motion of a
four-wheeled “nanocar” on an Au(111) surface at 200 °C. Wheels are shown by yellow spots,
and the path is highlighted with a white arrow. Reprinted with permission from ref 1129.
Copyright 2005 American Chemical Society.

In a four-wheeled “nanocar” with fullerene wheels, thermally induced translational motion on


a Au(111) surface was observed (Figure 38).(1129) Directional motion perpendicular to the
axles of the molecule suggested that the observed movement was generated by rolling of
the fullerene wheels rather than sliding of the entire molecule on the surface. In addition to
thermal activation, an STM tip could be used to drive the motion. Pivoting also took place,
clearly visible as small-angle perturbations to the path of translation in the STM images
(Figure 38e,f). In a structurally related three-wheeled molecule, pivoting was observed to be
the dominant motion on the surface, suggesting that the four-wheeled molecular structure
was important for a rolling translational motion. Similar examples of four- and three-wheeled
uorescent molecules with different wheel sizes (adamantane and p-carboranes as wheels)
have been reported, and their diffusion constants on glass surfaces were determined by
uorescence measurement.(1130, 1131) The photoisomerization kinetics of a similar four-
wheeled molecule functionalized with a rotary motor have also been investigated.(1132)
  
When fullerene wheels were used no photoisomerization took place, whereas with carborane
wheels, e cient isomerization was observed in solution.

The rotation of dialkylthioethers adsorbed on Cu(111) surfaces has been extensively studied
and electrons from an STM tip found to drive a 5% bias in rotational direction.(1133)
Electrical excitation of C–H stretching modes in the molecule contributed to ratcheting in the
rotation. The direction and rate of rotation depended on the chiralities of both the molecule
and the STM tip. In a cerium-centered double-decker molecule self-assembled on a Au(111)
surface, an interesting phenomenon was detected (Figure 39).(1134) The rotational chirality
of individual molecules generated two different orientations on the gold surface, which could
clearly be distinguished by STM. Upon scanning the surface with the STM tip, an irreversible,
abrupt change in the chirality of some of the molecules was observed. The switch resulted
from the rotation of the upper porphyrins, and the irreversibility was attributed to damage
caused to the molecules by the high applied voltage.

Figure 39

Figure 39. (a) Structure of a cerium-centered double-decker molecule. (b) STM micrographs
of the molecules assembled on Au(111) surface. The two distinct isomers, due rotational
chirality, are shown in white and blue crosses. (c) Space- lling model of the two chiral
species. Reprinted with permission from ref 1134. Copyright 2011 American Chemical
Society.

Figure 40
  

Figure 40. (a) Chemical structure of a rotary motor with the groups responsible for double-
bond isomerization (red) and helix inversion (blue) highlighted. (b) Schematic representation
of a full 360° rotation with sequential double-bond isomerization and helix inversion (hexyl
substituents are omitted for clarity). (c) Schematic representation of electron tunneling
exciting the molecule and inducing translational motion on the surface. (d) Cartoon
representation of the motion. Reprinted with permission from ref 1135. Copyright 2011
Nature Publishing Group.

Directional rotary motors undergoing well-de ned structural and/or chemical changes on the
application of an external stimuli were discussed in section 5. Feringa et al. developed a
system utilizing paddle-wheel type directional motion over a Cu(111) surface upon
sequential electronic and vibrational excitation (Figure 40).(1135) Electron tunneling induced
by a STM tip stimulated double bond isomerization, which was followed by helicity inversion.
These sequential con gurational and conformational changes in the molecule propelled it
across the surface. For the molecule to move on the surface linearly, the four rotary wheels
had to rotate in the same direction upon application of the voltage. This is only possible
when the molecule was attached to the surface as the meso-(R,S–R,S) isomer. In other
isomers, a lack of coherent rotation prevented translational motion or induced spinning.

Hao et al. monitored the strength of metal–ligand complexation with an AFM tip.(1136) One
terpyridine ligand was attached to a gold surface while another was tethered to a gold-
coated AFM tip, and the bonding forces between osmium and the ligands were analyzed.
The use of atomic tools to induce relative motion in an interlocked molecular system is of
particular interest in the development of molecular machines, because each component can
be designed to function like the mechanical components of a macroscopic machine.   
(1137, 1138) In a polyrotaxane architecture adsorbed on MoS2, cyclodextrin ring was
translocated by a STM tip along a polyethylene glycol thread.(1139) Such shuttling was
thought to be responsible for the conductance switch observed upon the application of a
voltage via a STM tip in a bistable [2]rotaxane.(1140) The potential use of interlocked
molecular constructs in information storage was highlighted by Leigh et al., who showed
that regular patterns of deformations could be successfully generated on a rotaxane
monolayer with the aid of a STM tip.(1141) These arrays formed due to the relative ease of
intercomponent mobility in these molecular constructs. The effect of electrostatic and steric
forces on the shuttling of a bistable [2]rotaxane tethered to an AFM tip has been investigated
both experimentally and theoretically.(1142) A molecular shuttle made up of fumaramide
and succinamide stations has been attached to a gold surface by thiol linkers (Figure 41).
(1143) The macrocycle was tethered to an AFM probe using poly(ethylene oxide) (PEO).
Strong hydrogen-bonding interactions with the fumaramide station resulted in a 95:5
distribution of macrocycle in favor of this station. As the stretching of the PEO tether
continued, and the force exerted by the PEO linker exceeded the hydrogen-bonding forces
between the macrocycle and the fumaramide station, the ring moved away from the
fumaramide station. Tension in the tether decreased as a result of the shuttling. Further
movement of the cantilever resulted in the detachment of the PEO linker from the probe.

Figure 41

Figure 41. (a) Chemical structure of a rotaxane with fumaramide and succinamide stations
depicted in green and orange, respectively. (b) Schematic representation of macrocycle
movement on a thread attached to a gold surface as a result of the force exerted by an AFM
probe. Application of force by cantilever movement (I,II) was followed by repositioning of the
macrocycle (III) or detachment of the PEO tether (IV) depending on the strength of the force.
Reprinted with permission from ref 1143. Copyright 2011 Nature Publishing Group.  
In addition to molecular level motion generated by atomic tools, modulation of conductivity
by molecular switches attached to nanojunctions is of great interest.(1144-1152) A STM tip
could cause the rotation of selected moieties in the structure of a polyaromatic scaffold, in
the hope of developing switchable nanowires.(1110, 1153-1160) Although most examples
are still far from practical application, signi cant progress has been made in this area.
(1161-1182)

8 Toward Applications of Molecular Machines Jump To

8.1 Current Challenges: Constraining, Communicating,


Correlating
Extracting useful work at the molecular scale requires the restriction of the thermal
movement of submolecular components or the exploitation of thermal motion with
additional ratcheting. Shuttling, switching, and rotation processes in solution can be
modulated externally, and the directionality of each motion can be controlled in single
molecules. Considering an ensemble of such molecules in solution, however, the biased
motions average to give no net directionality. For various applications, the integrity of the
molecular system must be conveyed to the macroscopic world. Although there are possible
applications of molecular machines/devices in solution, such as their use in molecular logic
gate construction or molecular sensing, they are not compatible with typical solid-state
technology. For this reason, molecular machines on solid supports are needed. This
challenge is being addressed with molecular devices and machines being built on surfaces,
interfaces, and polymer matrixes with a variety of electronic, mechanical, or biological
applications.(1183-1209) Controlled drug release through nanovalves,(1210-1226) molecular
electronics,(1227-1234) arti cial molecular muscles,(18, 1224, 1235-1242) information
storage,(1141, 1243, 1244) and modulation of surface properties(1245) are topics of active
research in this area.

Figure 42
  

Figure 42. Communication between molecular devices. Acid generated upon conversion of
merocyanine (MEH+) to spiropyran (SP) protonates a pyridine, and leads to subsequent
complexation of the pyridinium ion (103+) by a calix[6]arene (104).(1246)

Communication between molecular machines is another pertinent challenge in this area. In


biological systems, the work done by one machine can be harvested by another and the
second can then operate. Raymo and Credi have developed a system that addresses this
challenge.(1246, 1247) They used a merocyanine-type photoacid to release protons as the
molecule transforms from its open form to the closed form. The acid protonated a
terpyridine osmium complex and decreased the e ciency of singlet oxygen generation.
(1247) These photogenerated protons were also used to reversibly complex and decomplex
a 1-alkyl-4,4′-bipyridin-1-ium (103+) in a calix[6]arene wheel (104) (Figure 42).(1246) Thus, the
control of one moiety caused a change in another via the transfer of a proton. Recently,
Aprahamian and co-workers successfully coupled a hydrazone switch (section 2.2) with a
merocyanine unit and used the photogeneration of acid to drive the switch reversibly with
high e ciency.(1248)

8.2 Reporting Controlled Motion in Solution


Submolecular movement can be designed to give a detectable output. The output can be
used as a measure of stimulant concentration (sensing) or it can be used for information
storage. In theory, any detectable nondestructive output that provides a reliable distinction
between states of the system is acceptable (NMR, CD response, etc.). However, for most
useful applications, more practical and easily readable optical, electronic, or mechanical
  
outputs are preferred. The response rate can be crucial in certain applications such as
memory devices and molecular sensors. In these systems, fast-responding reporters are
preferred. For a molecular system to be used repeatedly, stability and fatigue resistance are
important.

8.2.1 Conformational Switches in Solution

The restriction of molecular motion or switching between bistable conformations can be


achieved by external stimuli such as ion or organic molecule binding, pH change, or by
photoisomerization of the molecule. Conformational perturbations in polymers or carbon
chains with interesting properties have been reported.(1249-1261) Fluorescence readouts
have been widely used to detect such submolecular motions in part due to their high signal-
to-noise ratios.(16, 1262-1266) One way to obtain a uorescence response is by changing
the electron transfer e ciency between a uorescent component and other moieties in the
system by inducing an electronic change in the acceptor or donor units. Ligand binding or
protonation can effectively change the HOMO and/or LUMO levels of molecules, which in
turn affects electron transfer to or from the uorescent moiety. The change in relative energy
levels upon reduction or oxidization determines the ease of detection of the uorescence
response. A greater response can be obtained by decreasing rotational degrees of freedom
or by energy transfer. Because energy transfer processes are highly distance dependent,
variations in distance between the donor and acceptor units in different conformational
isomers can determine the e ciency of energy transfer. Emission from dimers of certain
chromophores is also widely used as a reporter of molecular processes. This emission
requires the close proximity of similar (excimer) or different (exciplex) chromophores.

Esquema 37
  

Esquema 37. Expansión y contracción controladas alostéricamente por coordinación de


ligandos a

Esquema a El centro de Zn (II) actúa como una unidad catalítica y se utiliza


dietilaminometilantraceno como indicador.(1267)

Mirkin y col. desarrollaron un catalizador supramolecular regulado alostéricamente basado


en la contracción o expansión dependiente del ligando para revelar u ocultar una unidad
catalítica ( Esquema 37 ). Pudieron medir el cambio conformacional indirectamente
utilizando el subproducto de la reacción, ácido acético.(1267)La especie tetrametálica
estaba formada por dos centros de Zn (II) alineados cofacialmente catalíticamente unidos
mediante dos centros Rh (I). La velocidad de catálisis de la reacción de transferencia de acilo
por los iones Zn (II) del núcleo se incrementó cuando la cavidad entre las dos unidades era
mayor. El intercambio de ligandos Rh (I) de tioéteres lábiles a cloruro y monóxido de carbono
condujo a una expansión y, por tanto, a una mejora de la actividad catalítica. A medida que el
subproducto de ácido acético protonaba la amina de dietilaminometilantraceno, se observó
un aumento de la uorescencia provocado por el bloqueo de la transferencia de electrones
fotoinducida del resto amina.

Lee y col. desarrollaron un interruptor conformacional basado en una red cíclica de enlaces
de hidrógeno en un derivado de tris ( N- salicilidenoanilina) ( Figura 43 ).(1268-1271)La
distorsión estructural inducida por los procesos de plegado y desplegado resultó en un
cambio signi cativo en las propiedades electrónicas del sistema. Los cromóforos BODIPY
estaban unidos a la molécula como aceptores de energía uorescente. En la forma plegada,
la molécula adoptó una matriz de enlaces de hidrógeno muy ordenada, y la transferencia
  de
energía tuvo lugar desde el núcleo de tris ( N -salicilidenanilina) a los BODIPY. Cuando se
añadió un aceptor de enlaces de hidrógeno como el uoruro, la molécula se desplegó y se
detectó una disminución sustancial en la uorescencia de los tintes BODIPY. Debido a que
cada elemento móvil cooperó en el proceso de plegado mediante enlaces de hidrógeno, el
cambio conformacional fue muy cooperativo. La intensidad de uorescencia inicial se pudo
recuperar cuando los aniones uoruro fueron capturados por cloruro de trimetilsililo.

Figura 43

Figure 43. (a) Chemical structure of the t-butylphenyl and BODIPY-substituted foldamers, and
(b) X-ray structure of the t-butylphenyl functionalized foldamer. Carbon atoms of the t-butyl
groups and all hydrogen atoms except OH and NH have been omitted for clarity. (c)
Schematic representation of conformational switching. D and A represent energy donor and
acceptor modules, respectively.(1268-1270) Parts (a) and (c) are adapted with permission
from ref 1268. Copyright 2007 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim. Part (b) is
reprinted with permission from ref 1269. Copyright 2006 American Chemical Society.

A number of different uorescent sensors reporting conformational changes have been


synthesized on the basis of energy transfer,(1272-1274) excimer formation,(1275-1282) or
charge transfer mechanisms.(1283-1286) Optical responses have been obtained by simple
stimuli-dependent conformational change,(796, 1262, 1273, 1275, 1287-1289) folding
processes,(1276-1278, 1290) shuttling,(799, 1272, 1283-1285, 1291-1294) switching,(1273)
and transitions between coiled and helical structures.(1295) Optical outputs have been used
to develop molecular logic gates.(1291, 1292, 1296) Recently, Tian et al. used both
phosphorescence and an induced circular dichroism output to develop an INHIBIT molecular
 
logic gate by taking advantage of the photoisomerization-dependent inclusion of an
azobenzene and a bromonaphthalene in the β-CD cavity.(1297)

Figure 44

Figure 44. (a) Ligand-induced variation of the chemical shifts (11B (160 MHz) and 1H NMR
(500 MHz)) of the helix in CD3OD at 298 K. (b) Schematic representation of fast exchange
between two degenerate helical conformers with a single NMR signal and (c) induced bias of
helicity upon ligand binding (yellow), and the anisochronous NMR signal generated. Adapted
with permission from ref 1298. Copyright 2013 Nature Publishing Group.

Techniques other than optical responses have been used to measure conformational
change. Flood et al. used a change in conductance to directly measure a folding process
because unfolding released a bound chloride anion and changed the conductance (Figure 22
in section 3.1).(595) A change in CD response can be a useful indicator of molecular motion
and helicity inversion.(1298-1308) Many shape changes can be easily analyzed by NMR
spectroscopy. Recently, Clayden et al. showed, by 11B, 1H, and 13C NMR, that the equal
distribution of left-handed (M) and right-handed (P) conformers of a helical foldamer could
be perturbed by the addition of a chiral ligand (Figure 44).(1298) Upon ligand complexation
to the boronic ester, chiral information was transferred along the helix, and the isochronous
NMR signals of the diastereotopic nuclei were transformed into anisochronous signals,
indicating a bias in the distribution of helical conformers. When the meso diastereoisomer of
the diol was used to complex the boronic ester, no splitting of the 1H NMR signals was
observed due to the symmetry of the ligand.   
8.2.2 Rotaxane Switches in Solution

Fluorescence spectroscopy has been widely used to detect shuttling processes in


interlocked systems, particularly in rotaxanes. The strong distance dependence of
uorescence quenching by electron transfer has frequently been exploited to measure the
relative position of molecular subcomponents. Photoisomerization-dependent shuttling of a
pyridinium bearing macrocycle along a two station thread was monitored by changes in the
uorescence of an anthracene group tethered near one station.(1285) When the distribution
of macrocycle was biased to this station, charge transfer from the anthracene to the
pyridinium moiety quenched the uorescence, thus providing a detectable uorescence
response due to macrocyclic position. The location of a macrocycle along an alkyl chain in a
rotaxane structure has also been probed using similar techniques.(1272)

Greater rotational and vibrational freedom leads to a greater probability of nonradiative


processes and hence a lower uorescence quantum yield. Any interaction affecting rotation
or vibration can in uence emission intensity. Tian et al. used this effect with a 4-
aminonaphthalimide uorophore in a stilbene bearing rotaxane where the steric hindrance
generated by photoisomerization of the stilbene unit resulted in the relocation of the α-
cyclodextrin macrocycle along the thread.(796) As a result of this movement, uorescence
increased by 46%, which was attributed to the restriction of molecular motion of the nearby
linker moieties. The same group utilized an azobenzene as an additional photoisomerizable
moiety to develop a molecular half adder logic gate (Scheme 38).(1291) This logic gate
performs binary addition using two inputs and two outputs to provide both XOR and AND
gates. When either the azobenzene or the stilbene moiety was isomerized, uorescence
emission of the naphthalimide close to photoisomerized species increased ((Z,E)-112 or
(E,Z)-112). However, if none or both of them were isomerized ((E,E)-112 or (Z,Z)-112),
uorescence intensity decreased due to rapid shuttling of the ring between the stations in
the nonisomerized form or because the ring was trapped in the center of the thread when
both functional moieties were isomerized. These uorescence outputs collectively result in a
XOR gate (output 1 in the truth table). The photoisomerization-dependent decrease in
absorbance at 350 nm and increase at 270 nm were used as an additional output for the
construction of the AND gate of the half adder. The change in absorbance at the isosbestic
point (301 nm) was followed and only breached a predetermined threshold when both units
were isomerized. As such, an interlocked molecular system was successfully used as an all-
photonic logic gate in solution where the inputs induce submolecular motion and outputs are
reporters of their motion. Acid–base and redox switchable bistable rotaxanes with crown
ether macrocycles have been used to create an INHIBIT logic gate.(1292) Room-temperature
phosphorescence and CD responses have been used in a pseudo[1]rotaxane to create a
primitive logic gate.(1297)

Scheme 38
  

Scheme 38. Photoisomerization-Dependent Shuttling of an α-Cyclodextrin on an Azobenzene


and Stilbene Bearing Thread with Two Naphthalimide Derivatives as Fluorescent Stoppersa

Scheme a The percentage of the major isomer in the photostationary state is shown over
the reaction arrows. The truth table for a half-adder logic gate is shown, with the inputs being
380 nm (I1) and 313 nm (I2) irradiation, and outputs being the change in absorbance (O2)
and uorescence (O1) values.(1291)

Indicator displacement assays involving supramolecular complexes and interlocked


molecules are widely used for molecular sensing applications.(1309) In these systems, a
uorescent molecule with complexation-dependent uorescence properties is usually used
as a guest. The optical change generated upon exchange of the uorescent guest with the
analyte is used as the reporting function. Smith et al. used a similar idea to develop a
chloride anion sensor based on a rotaxane (Figure 45).(1310, 1311) A [2]rotaxane (113) was
synthesized using a tetralactam macrocycle and a squaraine dye as building blocks.
Tetralactam macrocycles are known to bind to and quench the emission of red light from
squaraines.(1312) The exposure of this dye, upon chloride-induced displacement of the
macrocycle, restored uorescence. The process could be reversed by chloride precipitation
as the NaCl salt. When the same rotaxane was adsorbed on a C-18 coated reverse-phase
silica gel plate and dipped into different concentrations of aqueous solutions of chloride, the
change in uorescence was large enough to be visually observable. Later, a ratiometric
chloride sensor based on the same squaraine rotaxane was reported.(786) Interlocked
systems are useful in such dye-displacement assays, as the displaced dye typically diffuses
away preventing reversibility, whereas the stable rotaxane structure allows reversibility and
reusability. In a similar example, the sodium ion-dependent shuttling of a [2.2.2]cryptand
away from a squaraine station enhanced squaraine emission.(1312) Interlocked
architectures have been mounted on metal nanoparticles and shown to keep their switching
ability.(1313)   

Figure 45

Figure 45. (a) Chloride-dependent shuttling of tetralactam macrocycle, and (b) subsequent
uorescence enhancement in CHCl3 upon titration with tetrabutylammonium chloride. (c)
Rotaxane solution in the absence (left) or presence (right) of chloride. Adapted with
permission from ref 1310. Copyright 2010 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

Liu et al. have developed a [2]pseudorotaxane with a dual stimulus luminescent lanthanide
switch using a diarylper uorocyclopentene as the photochromic component (Figure 46).
(1314) The diarylper uorocyclopentene (114) bears a benzyl ammonium recognition unit
and reacts reversibly to generate its closed form upon absorption of UV irradiation (365 nm).
The open form can be obtained via irradiation at 614 nm. The Eu3+ complex of
terpyridinyldibenzo-24-crown-8 (115) was used to reversibly complex the ammonium moiety
of 114. This Eu3+ center uoresces at 619 nm through intramolecular energy transfer from
an excited terpyridine ligand. Upon complexation of the two compounds through crown
ether–ammonium interactions, a small amount of quenching of the lanthanide uorescence
at 619 nm was observed (ca. 10%) due to the poor spectral overlap between the donor
emission and acceptor absorbance (which is required for resonant energy transfer (RET)).
Upon UV irradiation to form the closed form of compound 114, spectral overlap increased
and uorescence was quenched by 80%, with an accompanying 3-fold decrease in excited-
state lifetime. The pseudorotaxane could be reversibly disassembled by the addition of
potassium cations and regenerated by the addition of 18-crown-6.

Figure 46
  

Figure 46. (a) Chemical structures of 115, and the open and closed forms of 114. (b)
Schematic representation of light modulated switch and K+/18-crown-6-mediated
complexation. (c) Fluorescence quenching of the Eu3+ complex upon UV irradiation in 1:1
CH3CN/CHCl3. Fluorescence before (I) and after (II) excitation at 390 nm (c inset). Reprinted
with permission from ref 1314. Copyright 2013 American Chemical Society.

The responsive nature of induced circular dichroism makes it an interesting phenomenon to


study with the CD response of one species changing upon its interaction with another.
(387, 1315-1319) The transformation of an achiral carbon center to a chiral one upon
hydrogen-bonding-dependent symmetry breaking provided an early example of the in uence
of supramolecular interactions on chirality.(1320) A number of chiral supramolecular
assemblies,(1315, 1316, 1321) chiral spaces (dissymmetric cavities),(1322) and helices
(1323-1325) have been reported. The solvent-dependent rearrangement of a macrocycle
along a thread has been shown to alter the CD response of a rotaxane (section 4.3.1).(715)
The position of a macrocycle has been used to control the reactivity of a secondary alcohol
toward esteri cation in the presence of a chiral catalyst in a chemical information ratchet
(section 4.4.1).(809, 810) Similarly, a chiraloptic switch has been reported based on a
benzylic amide macrocycle and fumaramide/peptide stations (Figure 47).(1285, 1326)
Photoisomerization-dependent relocation of the macrocycle to the peptide station altered
the CD response due to proximity-dependent macrocycle–leucine residue interactions. A
large reversible difference in elliptical polarization response (>1500 deg cm2 dmol–1) was
obtained upon isomerization of the fumaramide station from E to Z.

Figure 47
  

Figure 47. (a) Chiraloptical switching upon photoinduced shuttling of the macrocycle
between fumaramide (green) and peptide (orange) stations; the chiral center of the peptide
station is highlighted by a black circle. (b) Percentage of E isomer calculated using 1H NMR
(left y axis) and induced CD absorption at 246 nm after alternating irradiation between 254
nm (half integer) and 312 nm (integers) (right y axis). Reprinted with permission from ref
1326. Copyright 2003 American Chemical Society.

Li et al. synthesized rotaxane 117 bearing a tetracyanobutadiene (TCBD) stopper, which


tends to form aggregates via intermolecular dipole–dipole interactions, to attempt to control
this clustering behavior (Figure 48).(1327) The position of macrocycle relative to the TCBD
group determined the strength of intermolecular interactions and hence the structure of the
assemblies formed. In a hexane:chloroform mixture (1/1, v/v), the macrocycle rested on the
peptidic station, and the TCBD units were free to interact to generate nano bers (Figure 48b).
In methanol:chloroform (1/1, v/v), the amphiphilic nature of the molecules generated
perforated nanocapsules (Figure 48c). Finally, when DMSO was used as the solvent, the
macrocycle relocated closer to the TCBD group, interfering with aggregation and generating
a worm-like nanostructure (Figure 48d). A similar phenomenon was reported in which a
fumaramide bearing rotaxane displayed different nanostructures when the macrocycle
bound Zn2+ metal in the presence or absence of irradiation.(1328) Anion and acid/base-
dependent control over the formation of an organogel has been achieved using a [2]rotaxane
architecture.(1329) In this system, stimuli-induced shuttling of the macrocycle between urea
and ammonium stations led to sol–gel phase transitions. Polyrotaxane structures have been
formed in a concentration-dependent manner in solution using the interaction of an electron-
de cient macrocycle with a monoanionic species.(1330) Recently, light and acid/base-
dependent threading and dethreading of pseudorotaxane structures embedded in nano bers
was reported to induce a macroscopic change (more than 1.5-fold enhancement of Young’s
modulus upon dethreading).(1331)

Figure 48
  

Figure 48. (a) Solvent-induced shuttling on a tetracyanobutadiene bearing rotaxane 117. (b)
Proposed assembly of rotaxane and SEM images in CHCl3/n-C6H14 (1/1, v/v), (c) in
CHCl3/MeOH (1/1, v/v), and (d) in DMSO. Reprinted with permission from ref 1327. Copyright
2009 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

The concept of a “molecular muscle”, created from doubly threaded rotaxanes or interlocked
daisy chain molecular structures, was developed by Sauvage et al.(1332) In these systems,
submolecular movement drives a relative motion that either contracts or expands the entire
molecule upon metal exchange,(1332) solvent exchange,(1333, 1334) redox reaction,
(1236, 1238, 1335) or acid/base exchange.(18, 1336-1339) Photoinduced
contraction/expansion has also been reported.(1340) The concept has been used to cause
the macroscopic bending of a microcantilever by means of redox-driven shuttling of the
surface-adsorbed macrocycle along a thread (section 8.6.4).(1238, 1341)

The design of enzyme-responsive molecular machines is an emerging area and could nd


applications in biotechnology. Biodegradable polyrotaxanes have been studied as selective
drug and gene delivery systems.(1342-1349) Leigh and co-workers published two
generations of [2]rotaxane-based propeptides (118, 119) where the peptide axle is protected
from degradation from general peptidases by the macrocycle.(1350, 1351) The introduced
glycosidase-cleavable stopper allows the release of the free peptide in a controlled fashion
through treatment with a speci c glycosidase. The sequence of reactions triggered by the β-
galactosidase leading to the decomposition of the stopper is shown in Scheme 39. This
approach offers a promising alternative delivery system for peptide-based therapies,
because many bioactive peptides suffer from in vivo instability and poor bioavailability.
Similar principles have been used to deliver a deactivated cancer drug (acting as a
stoppering moiety in a rotaxane) via the blood before the macrocycle immolates inside a cell.
This releases the thread, and enzyme-mediated hydrolysis furnishes the active drug
molecule. The rate of hydrolysis in cancerous cells versus benign cells was somewhat
enhanced due to their overexpression of the hydrolyzing enzyme.(1352)

Scheme 39
  

Scheme 39. β-Galactosidase-Triggered Release of Triglycyl Peptide 120 from the


[2]Rotaxane Propeptide 118 and 119(1351)

8.3 Synthetic Molecular Walkers


8.3.1 Introduction

Biological motor proteins have evolved a plethora of functionalities. DNA, RNA, and protein
synthesis machineries (DNA polymerases, RNA polymerases, the ribosome) work
cooperatively with helper proteins to “unzip” molecules (helicases,(1353) poly(ADP-ribose)
polymerases),(1354) release strain in a substrate (gyrase,(1355, 1356) topoisomerase),
(1357, 1358) slide over the substrate, and processively synthesize a product (DNA
polymerases, RNA polymerases, the ribosome).(68-70) The ATPase rotary motor is the
energy factory of the cell producing ATP via complex mechanical processes.(1359) Pumps
(1360) and pores(1361) are essential to maintain communication and transportation across
the membranes of different compartments (functioning at organelle, cell–environment, or
cell–cell boundaries). Among these machines, molecular walkers are attracting increasing
attention as their mechanism of action has been revealed. Dynein, myosin, and kinesin are
walker proteins of the ATPase family and differ in their structure, function, and use of energy.
(1362-1365) Proteins such as collagenase and exonucleases, which are not traditionally
viewed as walker proteins, migrate along their substrate tracks by destroying the track via a
“burnt-bridges” mechanism.(1366-1368)
Biological walkers have some important characteristics.(1369) First, they must be supplied
with energy to do work against random Brownian uctuations. ATP hydrolysis normally   
provides this energy. Second, attachment to a track results in a restriction of their degrees of
freedom facilitating directional 1-D or 2-D walking in solution. Inertia and momentum under
conditions of low Reynold’s number are irrelevant, and work is performed under the in uence
of viscous and thermal motions. Third, to drive the walker away from equilibrium, that is, to
generate directional motion, a ratcheting process (either an energy or an information ratchet)
must take place. In addition to the requirements of a Brownian motor, certain additional
characteristics are necessary for a motor to be de ned as a walker.

(i) Processivity is a measure of the integrity of walker–track interactions during its


operation. A walker should remain attached over multiple operations to extract
useful work.

(ii) Directionality is the exclusive or preferential movement of the walker in one


direction along the track.

(iii) Repetitive operation means the motor should be able to carry out similar
mechanical cycles repetitively.

(iv) Progressive operation is the ability of the motor to culmulatively perform work
with each operating cycle.

(v) Autonomous operation allows the motor to function continuously without


external intervention, as long as an energy source (fuel) is available.

Conventional kinesin (kinesin I) meets all of these criteria and was rst isolated by Vale et al.
in 1985.(1370) It is a homodimeric protein with two identical heads that interact with a
microtubule track. Through cyclical binding, hydrolysis, and then release of ATP, it can walk
along the track by an asymmetric hand-over-hand mechanism.(1371-1373) Because at any
time one of the heads remains attached to the track during the walking process, processivity
is high. On average, this walker takes 100 steps before detaching from the track.(1374-1377)
Although walkers with two attachment points or more are common, this is not a requirement
for processivity. The KIF1A kinesin protein processively walks along microtubules using a
single leg.(1378) In contrast, some 2-legged biological walkers such as myosin-II are
nonprocessive.(1365) Even without being processive, myosin-II can still move directionally,
and in a large ensemble it can generate macroscopic motions such as muscle contraction.

DNA has the notable ability to form predictable hydrogen bonds with a complementary
strand and can be synthesized using machine-assisted technologies. Strand displacement
provides a versatile design for walking processes, and the ability to make complex 3-D
structures can be exploited to perform complex physical tasks. The rst arti cial walkers
were made of DNA. Directional walkers,(1379) autonomous walkers,(1380) walkers using the
burnt-bridges mechanism,(1380) cargo-carrying walkers,(1381) and a light-driven walker
 
(1382, 1383) based on DNA have been reported, and will be examined in section 9.4. Aside
from DNA-based walkers, the preferential diffusion of molecules along the higher symmetry
axis of a metal surface has also been observed.(1117, 1127) Huskens et al. reported the
gradient-driven diffusion of molecules bearing two adamantane legs across α-cyclodextrin-
functionalized surfaces.(1384) Processive crown ether migration on an oligoglycine chain
(1385) and migration of metal atoms on aromatic scaffolds(615, 1386, 1387) have also been
explored. Rearrangement reactions enable the migration of a fragment along a molecule
(such as the Claisen(1388, 1389) and Cope(1390-1393) rearrangements). However,
progressive and repetitive operation using sigmatropic rearrangement reactions is generally
challenging.

For the design of processive small molecule synthetic molecular walkers, mechanically
interlocked architectures are good candidates, because the walker (macrocycle) is
mechanically bonded to the track (thread). However, their interlocked-structure prevents
resetting because each resetting process undoes the walking process, by relocating the
walker to its original position.(1369) Moreover, the macrocycle cannot choose a new path
without bond breaking, which would result in detachment from the thread. A two-leg system
operating under orthogonal conditions could provide a viable synthetic molecular walker.
Orthogonal dynamic covalent exchange reactions can provide a suitable balance of
reversibility and kinetic stability. A rigid track can decrease the possibility of overstepping.

8.3.2 Spontaneous Walking of Small Synthetic Molecular Walkers

Reversible and processive migration of a Michael acceptor along a protein was reported in
1979 by Lawton et al.(1394) More recently, Lehn et al. used dynamic imine exchange to
transport a salicylidene walker 125 along an amine bearing track (Figure 49).(1395)
Deprotection of the amine at one side of the track using methoxyamine initiated the dynamic
exchange reactions. The speed of exchange was shown to be modulated when substitution,
composition of the solvent, or temperature was altered. 1H NMR analysis in CDCl3 proved
that the relative intensity of a characteristic signal corresponding to the transported walker
(H6) increased gradually and in 2 days the transported walker was the dominant product in
solution. Recently, the same group reported a modi ed version of their walker with
thermodynamic sinks at each end of the track. These trap the stochastic walker at one end
in acid as an imine, or as a lactone at the other end under basic conditions.(1396)

Figure 49
  

Figure 49. (a) The structure and operation of walker 125, which uses dynamic imine
exchange chemistry. Amine footholds are highlighted in blue and red. Each molecule is
labeled with one or two numbers in parentheses indicating the amine footholds to which the
walker is attached (foothold amines are assigned with numbers starting from the left). (b) 1H
NMR spectra of indicated protons in CDCl3. H7 corresponds to a side product in which the
walker is detached from the track. Reprinted with permission from ref 1395. Copyright 2012
American Chemical Society.

Figure 50

Figure 50. (a) The chemical structure of the model walker 127, and (b) gradual change in 1H
NMR of the model walker in d6-DMSO upon formation of new positional isomer on the track.
Ratio of (1):(2) isomers reaches 1:0.9 after 15 h of operation. (c) Chemical structure of the
walker on a larger track bearing an anthracene moiety, 128. (d) Fluorescence quenching of
anthracene after 6.5 h of walking. Reprinted with permission from ref 1397. Copyright
  2012
Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

Figure 51

Figure 51. (a) Chemical structure of walker 129, with each positional isomer labeled in a
different color, and amines numbered. (b) Change in the partial 1H NMR spectra over 48 h of
operation. (c) Occupancy of each foothold over time. Reprinted with permission from ref
1398. Copyright 2013 American Chemical Society.

Leigh et al. reported a spontaneous walking process using a Michael addition reaction
between an α-methylene-4-nitrostyrene walker and a polyamine track (Figure 50).(1397) It
was shown that the walker translocated preferentially through successive 1,4-N,N-migration
rather than by overstepping. After 48 h, no out of sequence N,N-migration was observed with
a model walker (Figure 50a,b). 1H NMR analysis showed that the walking was highly solvent
dependent and accelerated in polar solvents such as dimethylformamide and dimethyl
sulfoxide. In d6-DMSO, the walker reached an equilibrium distribution in 15 h. A ratio of 1:0.9
between the occupancy of initial position and the central amine of the track was achieved
over this time period. Intermolecular exchange of less than 6% took place when the model
walker was mixed with a longer free track for 3 days. This indicates a high processivity (on
average 530 steps taken before detachment). Using a longer track modi ed with an
anthracene at the far end, it was possible to monitor walking via an observed decrease in
uorescence as the walker approached the anthracene. As the nitrostyrene walker
approached the edge of the track, it quenched the uorophore (Figure 50d). Leigh et al. have
since reported an extended system with a naphthalene moiety at one end of the track (Figure
51 a).(1398) In the presence of excess base (iPr2NEt), the walker was “trapped” by the
naphthylmethylamine, which was the thermodynamic sink. 1H NMR analysis showed  that 
the steady-state distribution of the walker was biased and 46% of walker molecules had
reached the nal benzylic foothold in 48 h (Figure 51b,c). When a longer track (with 9
footholds) was used, the percentage at the nal station dropped to 19%.

Scheme 40

Scheme 40. Structure and Operation of a Small Molecule Walker, Walking along a Molecular
Tracka

Scheme a Footholds are shown in blue and green; the walker unit is depicted in red.
Numbering shows the footholds to which the walker is attached.(1399)

A walker with two chemically different legs that labilize under orthogonal conditions has
been reported by Leigh et al. (130, Scheme 40).(1399) The walker unit was attached to the
thread by hydrazone and disul de linkages. Hydrazone exchange took place under acidic
conditions, while under these conditions the disul de bond was stable. Disul de exchange
required basic conditions, which did not affect the hydrazone bond. Successive acid–base
cycles led to a 39:36:19 (1,2:2,3:3,4) distribution of walker, with only 6% forming the
overstepped (1,4) isomer. Changing the length of the walker alkyl chain from 6 carbons to 5
and using oxidation instead of base exchange at the nal step, the distribution could be
biased in favor of the (3,4) isomer, which was attributed to strain in the (2,3) isomer.(1400)

Recently, Bayley et al. monitored an organoarsenic molecule walking along a cysteine


bearing track in a protein pore, by measuring changes in ionic current. The walker “stepped”
by thiol exchange reactions and displayed a limited degree of processivity and directionality
(the latter due to the thermodynamic sink at the track terminus).(1401) An inchworm walker
able to walk on a copper surface has been reported. It could be constrained to one
dimension of motion by the use of oligomeric “fences”.(1402)

8.3.3 Directional Synthetic Small Molecule Walkers

The biased migration of molecules along a track requires either a preference for one isomer
over others under the operating conditions or that strain should be released upon stepping.
An e cient ratcheting step is needed. Recently, the use of metal complexes in a molecular
biped was explored by Leigh et al. (Figure 52).(1403) The track contained three different
pyridine derivatives: 2,6-dialkyl-4-N,N-dimethylaminopyridine (foothold 1, green), 3,5-
  
dialkylpyridine (foothold 2, red), and 2,6-dialkylpyridine (foothold 3, blue). The walker
consisted of pincer ligands able to complex the pyridine foothold via Pd(II) or Pt(II) centers.
Initially, the walker was attached to the track by complexation of Pd(II) with foothold 1, and
Pt(II) with foothold 2. Upon protonation of the free 2,6-dialkylpyridine foothold (foothold 3)
and heating, exchange of the Pd(II) complex between foothold 1 and foothold 3 took place. A
distribution of 15:85 in favor of foothold 3 was observed. The proton on foothold 3 was
captured by the more strongly basic foothold 1 in this process. Migration could be reversed
when the rst foothold was deprotonated and the solution was heated. A ratio of 95:5 in
favor of the initial position was obtained. Even though this example is not truly a walker, the
signi cant directional bias in the stepping process of the metal-based biped could be used to
inform more advanced designs.

Figure 52

Figure 52. Toward directional molecular walkers. (a) Chemical structure and schematic
representation of the walker attached to the track, and (b) operation of walking through
successive acid–base addition and heating cycles. Reprinted with permission from ref 1403.
Copyright 2014 American Chemical Society.

Finally, a small molecule walker displaying all of the desired properties of an arti cial walker,
save autonomy, has been reported by Leigh et al. (Scheme 41).(1404) The walker was based
on the previously published 130 (Scheme 40), but a stilbene moiety now linked footholds 2
and 3 (Scheme 41). The walker operated by the same hydrazone and disul de exchange
reactions, but with photoisomerization of the stilbene moiety driving directionality.
Isomerization led to greater proximity between footholds 2 and 3 (over 1 and 2) in the cis
form of the stilbene, resulted in a steady-state distribution of 40:60 in favor of the (2,3)
positional isomer. During the second hydrazone exchange step, reisomerization to trans
stilbene moved footholds 2 and 3 apart, which led to almost exclusive (5:95) formation of the
(3,4) isomer.

Scheme 41
  

Scheme 41. Light-Driven Transport of a Molecular Walker along a Molecular Tracka

Scheme a Footholds are shown in blue and green; the walker unit is depicted in red. Each
molecule is labeled with two numbers in parentheses indicating the two footholds to which
the walker is attached. (a) (i) hν (365 nm), CD2Cl2, (ii) DBU, DTT, CHCl3; (b) (i) I2, hν (500 nm),
CD2Cl2, (ii) TFA, CHCl3.(1404)

8.3.4 Challenges Yet To Be Overcome

Signi cant progress has been achieved in the synthesis of synthetic molecular walkers with
highly processive, progressive, and in some cases directional walkers being reported. The
exploration of additional, reversible, walker–track interactions may lead to greater control
over the directionality of motion. Autonomous walking remains a challenge to be addressed.
Polymeric tracks could be designed for cargo transport over large distances. The ability to
immobilize walkers on surfaces could lead to future technological applications.

8.4 Switchable Catalysts


In the living cell, many processes and reactions occur in parallel. To make sure that these
reactions and their products do not interfere with each other, these operations must be
rigorously controlled and switched on or off when necessary. In nature, enzymes control the
outcome and rate of the large majority of reactions taking place in a cell. Recently, a number
of biologically inspired systems have been published using molecular machines to control
the outcome of reactions by switching on and off catalytic units or by controlling the
enantioselectivity of the process.(1405-1412) In this section, switchable catalysts controlled
by stimuli such as light, pH change, ion in ux, and redox chemistry will be explored.

8.4.1 Photoswitchable Catalysts

Several examples of photoswitchable catalytic systems have been reported.


(1407, 1413-1415) The reactivity of the catalytic unit is typically controlled through alteration
of the steric shielding of the active site or by bringing the catalytic units into closer proximity.
In these examples, switching between the cis and trans forms of a photochromic group such
as an azobenzene or a stilbene, or the electrocyclization of a diarylethene, provides the
requisite control. An early example of a photoswitchable catalyst was published by Rebek Jr.
and co-workers.(1416) Two adenine receptors capable of forming complexes with purine
bases were linked through a trans azobenzene moiety. When the system was isomerized to
its cis conformation, the rate of reaction between an amine and p-nitrophenyl ester was
markedly increased due to greater proximity between the reactive groups, which increased
the effective molarity of these groups and thus increased the rate of reaction. A switchable
catalyst based on a light-responsive cavitand was published by the same group. The
cavitand was functionalized with an azobenzene switch and the cavity of the trans state 
accessible to a guest molecule, while the cis conformation could not bind guests.
Piperidinium acetate was used as the guest, and it could be shown that the host–guest
complex signi cantly accelerated a Knoevenagel condensation between malononitrile and
an aromatic aldehyde.(1417) Hecht and co-workers described the photoswitchable tertiary
piperidine derivative 133, which could function as a general base catalyst in a Henry reaction
between p-nitrobenzaldehyde and nitroethane (Figure 53). Catalysis was regulated through
the shielding/deshielding of the catalytic site upon irradiation of an azobenzene unit.(1418)
The reversible shielding and deshielding of a catalytic site by photoisomerization has also
been used to control the organocatalytic activity of a thiourea and a guanidine catalyst.
(1419, 1420)

Figure 53

Figure 53. (a) Molecular structure of the photoswitchable piperidine base 133, and (b) X-ray
structure of the photoswitchable piperidine base 133. X-ray crystal structure reprinted with
permission from ref 1418. Copyright 2008 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

The photoswitchable organocatalyst 134 consisting of a photochromic diarylethene (DAE)


unit and an imidazolium salt was reported by Bielawski and co-workers.(1421, 1422) Under
ambient light and in the presence of base, the imidazolium species catalyzed
transesteri cation and amidation. These reactions were considerably slowed upon
photoinduced formation of the ring-closed isomer (Scheme 42).(1421)
Scheme 42
  

Scheme 42. Ring-Opened and Ring-Closed Isomers of the DAE-Modulated Photoswitchable


Organocatalyst 134(1421)

Exploiting the ability of dithienylethene to switch between ring-closed and ring-opened


isomers upon irradiation, Branda et al. designed a photo responsive system mimicking
enzyme cofactor pyridoxal 5′-phosphate 135 (PLP) (Scheme 43). PLP is responsible for
amino acid metabolism and participates in a diverse range of enzymatic reactions such as
transamination, racemization, decarboxylation, and various elimination processes. The
structural features responsible for the action of PLP are the conjugated aldehyde and
pyridinium functional groups. Branda and co-workers replaced the core ring of PLP with a
dithienylethene where, in the ring-open form, the pyridinium and aldehyde units were
electronically isolated from each other preventing catalytic activity (138-inactive). However,
photoirradiation to form the ring-closed (138-active) isomer resulted in a fully conjugated
structure, which restored the connectivity between the pyridinium and the aldehyde groups
and therefore led to catalytic activity.(1423)

Scheme 43

Scheme 43. (a) Reaction of an Amino Acid with PLP Showing the Aldimine 136 Initially
Produced, and the Molecular Structure of the Quinonoid 137 Formed after Removal of the
Amino Acid’s α-Hydrogen, Which Leads to Racemization; and (b) Photoresponsive PLP
Mimic 138(1423)
Feringa et al. used one of their rotary molecular motors based on a chiral overcrowded
alkene, which can perform a directional 360° rotary cycle fueled by light and heat to create
 a
photoswitchable catalyst. During this rotation, the helicity of the motor changes, which
changes the overall chirality of the system. A DMAP (dimethylaminopyridine) Brønsted base
(Figure 54) and a thiourea hydrogen-bonding donor group, which together comprise a
bifunctional organocatalyst, perform Michael additions when in close proximity. As a
consequence, the catalytic activity as well the stereoselectivity can be controlled. Michael
addition of 2-methoxy thiophenol to cyclohexanone was used as a test system. The (P,P)-
trans-140 isomer resulted in a racemic (R,S)-thiol adduct in low yield (7%) after a lengthy
reaction. When the (M,M)-cis-140 isomer was used, the Michael addition proceeded
signi cantly faster furnishing a 50% yield of product, with an er of 75/25 (S/R). Finally, the
(P,P)-cis-140 isomer led an 83% yield with an inversion in enantioselectivity providing an er of
23/77 (S/R).(1424) Recently, the same group reported a bisphosphine derivative of the same
photoswitchable rotary motor. Both product enantiomers of a palladium-catalyzed
desymmetrization reaction could be formed, although in situ switching experiments were
complicated by the photosensitivity of the active palladium complex.(1425)

Figure 54

Figure 54. Schematic representation and molecular structure of a bifunctional


organocatalyst integrated in a directional light-driven molecular motor. A, DMAP; B, thiourea
hydrogen-bonding donor group. (a) A and B are remote. (b) A and B are in close proximity
with M helicity in the (M,M)-cis-140 isomer, preferentially providing the (S) enantiomer of the
reaction product. (c) A and B are in close proximity with P helicity in the (P,P)-cis-140 isomer,
generating (R) enriched product. Step 1: irradiation at 312 nm at 20 °C. Step 2: heating at 70
°C. Step 3: irradiation at 312 nm at −60 °C. Step 4: temperature −10 °C.(1424)
8.4.2 pH-Dependent Switchable Catalysts
  
Several rotaxanes containing a catalytic unit have been reported. However, these cannot
typically be switched on/off.(1318, 1426-1430) Leigh et al. described the rotaxane-based
switchable organocatalyst 141, in which catalytic activity could be controlled by pH-
dependent macrocycle shuttling (Scheme 44). The design consisted of a dibenzo[24]crown-8
macrocycle and an axle containing both triazolium rings and a dibenzylamine/ammonium
moiety. The secondary amine/ammonium group was able to carry out iminium catalysis.
However, when the rotaxane was protonated, the ammonium group was a better binding site
for the macrocycle than the triazolium ring, and so the macrocycle blocked the catalytic
center, preventing the reaction. When the secondary amine of the rotaxane was not
protonated, the triazolium ring provided a better binding site for the macrocycle and the
dibenzylamine group was exposed, and could therefore participate in catalysis. As a model
reaction, the Michael addition of an aliphatic thiol to trans-cinnamaldehyde was performed.
The free thread was found to e ciently catalyze the reaction in both its protonated and its
deprotonated states.(1431) Recently, Leigh and co-workers explored the activation modes of
rotaxane catalyst 141 and published a chiral version of this design, with a chiral center next
to the secondary amine.(1432, 1433) This chiral organocatalyst was able to perform an
asymmetric Michael addition with good stereoselectivity.(1433) The pH-driven shuttling of a
pyridyl-2,6-dicarboxyamide macrocycle between squaramide (hydrogen bond catalyst) and
ammonium (iminium catalyst) stations has been reported.(1434) The macrocycle blocks the
site it is bound to, allowing the selective exposure of catalytic sites and the generation of
different products.

Scheme 44
  

Scheme 44. Acid–Base Switching of the Position of the Macrocycle, Which Conceals or
Exposes the Catalytic Site on the Rotaxane(1431)

8.4.3 Allosteric Regulation of Switchable Catalysts through Ion Addition

A different approach to the regulation of catalytic activity is via the addition of small
molecules such as ions, which can change the supramolecular structure of the catalyst, as is
observed in many enzymatic processes. Mirkin and co-workers reported the synthesis of
triple-layer complex 142 composed of two transition metal hinges, two chemically inert
blocking exterior layers, and a single catalytically active interior Al (III)-salen complex, which
can act as a ring-opening polymerization catalyst for ε-caprolactone (Figure 55). Polymer
growth and molecular weight could be controlled by the addition of Cl– (catalytic layer
exposed) or of the Cl– abstracting agent NaB(ArF)4, which results in the fully closed complex
142.(1435-1437) Recently, the same research group reported an allosterically regulated
photoredox catalyst based on a similar switching mechanism.(1438)

Figure 55
  

Figure 55. (a) Allosteric supramolecular triple-layer complex 142, which regulates the
catalytic living polymerization of ε-caprolactone. (b) Molecular structures of the
components.(1435)

Another ion triggered switch was the self-locking system published by Schmittel and co-
workers (Scheme 45).(1439-1441) Their design consisted of a zinc porphyrin and a 4-aza-
2,2′-bipyridine tether, which either coordinated to the zinc porphyrin or formed a complex
with copper and a shielded phenanthroline. In the presence of copper, the tether was
removed from the zinc porphyrin center, and piperidine was bound at the coordination site.
When copper was removed, piperidine was liberated and able to catalyze Knoevenagel
reactions.(1439)

Scheme 45

Scheme 45. Schematic Representation of the Reversible Locking and Unlocking of


Switchable Catalyst 143(1439)

8.4.4 Redox-Driven Switchable Catalysts


Canary and co-workers have reported a redox-switchable chiral catalyst capable of delivering
either enantiomer of a nitro-Michael addition product depending on the oxidation state
of a 
single copper atom. The design, inspired by previous work carried out in their group and by
others,(1442-1444) is based on complexes derived from methionine, which were shown to
undergo inner sphere ligand rearrangement upon one-electron oxidation or reduction of
copper. The rearrangement is coupled to the orientation of quinolone rings to afford
enantiomeric con gurations. Urea catalysis produced the (S) enantiomer in 72% ee and 55%
yield, and the in situ reduced complex produced the (R) enantiomer in 71% ee and 43% yield.
(1445)

8.5 Synthesis Using Arti cial Molecular Machines


Nature provides us with many examples of biological molecular machines that engage in
sequence-speci c chemical synthesis. For example, the ribosome utilizes sequence
information stored in mRNA to synthesize proteins with excellent delity.(68-70) The various
DNA polymerases provide further examples of sequence-speci c polymer synthesis by
biological molecular machines.(71) The inherently programmable nature of DNA has been
used by synthetic chemists to accomplish sequential oligomer synthesis. A common
strategy is exempli ed by the seminal work of Liu et al.,(1446) where complementary
sequences of DNA capped with different, mutually reactive end groups are annealed to
create an extremely high local concentration of the two reactive species ensuring
intracomplex bond formation predominates. After the rst reaction is complete, a biotin tag
on one strand allows its selective removal, after which a new reactive partner can be
annealed to the existing strand and a second reaction conducted. Peptide,(1446-1448)
carbon–carbon,(1449-1452) carbon–heteroatom bond forming reactions,(1453) and even
cycloaddition reactions have all been reported on the basis of this type of methodology.
(1454) More elaborate DNA systems have been used by Seeman and co-workers to create a
“molecular assembly line” where three gold nanoparticles could be combined to form eight
differently composed products, illustrating the complexity that can be achieved with DNA-
based molecular machines.(1455)

In a seminal example of a synthetic system mimicking some of the properties of an enzyme,


Nolte, Rowan et al. reported a rotaxane-based catalytic system.(1456) A glycoluril clip-based
macrocycle containing a manganese porphyrin catalyst catalyzed the epoxidation of a
butadiene polymer, which formed the thread of the rotaxane. A greatly enhanced proportion
of cis-epoxide was observed in the product, consistent with previous work showing that
when the reaction occurred in the macrocyclic cavity steric constraints favored the cis
product.(1457) This system has been further optimized by substitution of the porphyrin
macrocycle with protective ligand groups, which increase turnover number (TON) and
provide even greater cis selectivity by preventing reaction on the exterior face of the
macrocycle (Figure 56).(1458) The rate of threading of their macrocycle onto polymeric
materials was also studied and found to be highly dependent on interactions between the
thread and the outside of the macrocycle preassociating the two moieties.(1459)
Figure 56
  

Figure 56. Porphyrin-catalyzed epoxidation of a butadiene polymer by 144, utilizing a


rotaxane architecture. Reprinted with permission from ref 1456. Copyright 2003 Nature
Publishing Group.

A cyclodextrin dimer has been used as both a catalyst (for the polymerization of δ-
valerolactone) and a molecular “clamp”, which guides the output of the polymer.(1460) In this
system, an α-cyclodextrin acts as the active site while a β-cyclodextrin acts as the clamp
guiding the growing chain away from the active site (Scheme 46). An additional example was
provided by Takata et al.,(1430) where the cyclization of propargyl or allyl urethane groups in
the rotaxane backbone to oxazolidinones was catalyzed by a palladium center, bound to the
macrocycle.

Scheme 46

Scheme 46. Cyclodextrin Dimer 145 Polymerization Catalyst(1460),a

Scheme a Reprinted with permission from ref 1460. Copyright 2011 Wiley-VCH Verlag GmbH
& Co. KGaA, Weinheim.

Recently, Leigh et al. reported the rst small molecule arti cial molecular machine (146)
capable of sequence-speci c peptide synthesis.(666) The machine is based on a rotaxane
architecture, which is used to ensure processivity in a manner reminiscent of both the
ribosome-mRNA structure and the DNA clamp of DNA polymerases.(68-71) Sequence
information contained in the track is directly converted into the sequence of the peptide
synthesized, and thus the track plays some aspects of the role of mRNA in the ribosomal
system. Reactive phenolic ester groups take the place of tRNA-bound amino acids andare 
sequentially “picked up” by the catalytic arm of the macrocycle. The catalytic arm bears a
cysteine group and operates by native chemical ligation (NCL) chemistry where the thiolate
of the cysteine unit reacts rst with the ester group before transferring the activated amino
acid to the end of the growing chain.(1461) The steric bulk of the loaded amino acid prevents
the macrocycle passing over the barrier unit before this reaction has taken place but allows
free passage of the macrocycle after being cleaved. This prevents the peptide sequence
from being scrambled and allows sequential reaction. Rigid spacing units between each
loaded amino acid minimize the likelihood of the catalytic arm encountering an out of
sequence amino acid. The catalytic unit consists of a cysteine-glycine-glycine motif with the
second two amino acids present to prevent an unfavorable 1,8-S,N-acyl shift transition state
during the second ligation (Figure 57).

Figure 57

Figure 57. Leigh’s small molecule peptide synthesizer, 146.(666)

After the third and nal amino acid has been cleaved from the thread, the macrocycle can
dethread and be isolated. The delity of sequence transfer in this process was demonstrated
by tandem mass spectrometry of the macrocyclic product, identical to an authentic sample
synthesized by conventional methods. No out of sequence or abbreviated products were
observed by HPLC–MS underlining the extremely high level of control afforded by this
system. This molecular machine showed processivity, sequence speci city, and autonomy
and an exceptional level of control at the molecular level. However, several problems remain
with this design. The rate of reaction is vastly slower than that of the ribosome; one peptide
bond takes an average of 12 h to form, whereas the ribosome makes approximately 20
amide bonds every second.(68) Additionally, information is read in a destructive manner;
once the macrocycle has cleaved the ester bonds, there is no way to reload the machine for
a second run. Finally, the NCL reaction used to catalyze peptide formation limits the scope of
amino acids that can be incorporated into the growing peptide chain.

Recently, Leigh et al. reported an extension to this system, creating a four barrier machine
whose nal product is a macrocycle-bound heptapeptide.(661) The sequence speci city of
the product was again con rmed by tandem mass spectrometry and compared to an
authentic sample. Although no problems were encountered in this system, whose nal bond
formation involves a 20-membered ring S,N-acyl transfer, the ever-expanding size of the
required cyclic transition state will at some point limit the length of potential peptide
  
products formed by the current design (Scheme 47).

Scheme 47

Scheme 47. Operation of Leigh’s Peptide Synthesizer(666),a

Scheme a Reprinted with permission from ref 666. Copyright 2013 American Association for
the Advancement of Science (AAAS).

8.6 Controlled Motion on Surfaces, in Solids, and Other


Condensed Phases
8.6.1 Controlled Motion in Solids and Condensed Phases

A major challenge is to use molecular machines for practical applications by utilizing


molecular scale changes to create macroscopic effects. Using molecular machines in the
solid state or other condensed-phase matter could lead to new materials with a higher level
of complexity with controlled cooperative molecular motion leading to changes in the
properties and function of the material on a macroscopic scale. In recent years, several
examples of molecular motion in condensed-phase matter have been described, some of
which show promise for applications in the elds of electronics and optoelectronics.
  
Garcia-Garibay and co-workers have investigated the rotational dynamics and photophysical
properties of the crystalline, linearly conjugated, phenyleneethynylene molecular dirotor 147
(Figure 58). A pentiptycene unit was incorporated as a central stator about which the two
anking ethynylphenylene rotators could adopt various torsion angles. X-ray diffraction
studies showed that in the solid-state structure of molecular dirotor 147 all of the
phenyleneethynylene chromophors are arranged parallel to one another and therefore shared
the same rotation axis. The chromophore displayed signi cant uorescence changes as a
function of interphenylene torsion angles. The authors suggest that with this system external
control of the rotation could be achieved by the application of an electric eld, which would
allow a rapid shift of solid-state uorescence emission and optical properties upon the
application of appropriate stimuli.(1462-1470)

Figure 58

Figure 58. (a) Molecular structure of linear conjugated phenylethynylene molecular dirotor
147. Rotation of the phenylene rotor (shown with an arrow) creates rotamers with varying
degrees of π-conjugation and so wavelengths of emission. (b) X-ray structure of the dirotor
147. X-ray crystal structure reprinted with permission from ref 1462. Copyright 2013
American Chemical Society.
Sozzani and co-workers have published several examples of molecular rotors in crystals with
open porosity, as well as molecular rotors embedded in porous frameworks such as  
aromatic or organosilica frameworks. In some of these systems, the rotational motion can
be actively regulated in response to guest molecules such as CO2, I2, tetraethylammonium
chloride, and water. These responsive materials may nd applications spanning from
sensors to actuators, which could capture and release chemicals on command.(1471-1474)
Recently, Schurko and Loeb published a metal–organic framework (MOF) material
containing dynamically interlocked components.(1475) They used [2]rotaxane 148 as the
organic linker and binuclear Cu(II) units as the nodes (Figure 59). Void spaces inside the rigid
framework allowed the macrocyclic ring of the rotaxane to rotate rapidly. Initially the
macrocycle is locked in place through hydrogen bonding from an ether oxygen atom of the
macrocycle to a copper-bound H2O. Dynamic motion occurs upon removal of the water
molecules by heating to 150 °C under vacuum, destroying hydrogen-bond interactions and
allowing rapid circumrotation about the rotaxane axle. Rotation can be quenched by the
addition of water. This type of material could be useful for the creation of solid-state
molecular switches and machines.(1475) A rotaxane-based molecular shuttle incorporated
into the structure of a MOF has recently been reported.(1476)

Figure 59

Figure 59. (a) Structure of [2]rotaxane 148. (b) Schematic representation of the structural
design components used to create the metal–organic framework. (c) X-ray structure of the
tetra-ester precursor to [2]rotaxane 148. (d) X-ray structure of a single unit of the
mechanically interlocked molecule, coordinated to four Cu(II) paddlewheel clusters. X-ray
crystal structure reprinted with permission from ref 1475. Copyright 2012 Nature Publishing
Group.
8.6.2 Solid-State Molecular Electronic Devices
  
In a series of ground-breaking but controversial experiments interfacing switchable
rotaxanes and catenanes with silicon-based electronics, molecular shuttles have been
employed in solid-state molecular electronic devices.(1146, 1477-1485) Bistable [2]rotaxanes
and [2]catenanes have been the subject of numerous experimental investigations in the
course of the development of such molecular electronic devices.(1233, 1478, 1486-1495)
Here, the bistable [2]catenanes and [2]rotaxanes feature a cyclobis(paraquat-p-phenylene)
(CBPQT4+) macrocycle and two stations, often a tetrathiafulvalene (TTF) site and a
dioxynaphtalene site (DNP). Initially the macrocycle preferentially resides over the TTF site
due to strong aromatic charge-transfer interactions between the components; this is referred
to as the ground-state coconformation (GSCC). Electrochemical oxidation of the TTF station
to form TTF2+ generates Coulombic repulsion between CBPQT4+ and TTF2+, and drives the
translation of the macrocycle to the DNP station, to give the metastable state
coconformation (MSCC). The process can be reversed on reduction of TTF2+ to TTF
followed by either thermal relaxation of the macrocycle to the TTF station, or reduction of the
bipyridinium units in the cyclophane ring to the corresponding radical cations, which reduces
the activation barrier to shuttling, restoring the system to the GSCC. These two mechanically
distinguishable states exhibit different characteristic tunneling currents. On the basis of
quantum mechanical computational studies, the MSCC state is predicted to be the more
highly conducting state. The switching cycle can be detected by a number of experimental
techniques including time- and temperature-dependent electrochemistry and spectroscopy.
Studies have shown that current levels on switching are in uenced by temperature, the
structure of the rotaxane/catenane, and the environment in which the molecular machines
are embedded.(1496-1498) Different environments, including Langmuir–Blodgett (LB) lms,
(1499-1501) self-assembled monolayers (SAMs),(1502-1506) and solid-state molecular-
switch tunnel junctions (MSTJs), have been extensively studied.
(1150, 1151, 1488, 1489, 1507-1512) In one particular MSTJ, a monolayer of switchable
rotaxane 149 was embedded between two conducting electrodes (Figure 60). This MSTJ
acts as a gate, which can be opened or closed in response to an applied voltage by changes
in conductivity and resistance and could be used in molecular logic gate designs. The
reported design showed stable switching voltages of −2 and +2 V, with reasonable on/off
ratios and low switch-closed currents. Nanometer-scale devices have been built using this
approach and connected to form 2-D crossbar circuit architectures.(1144) As a next step, the
authors published the design of a 160-kilobit molecular electronic memory circuit consisting
of 400 silicon-nanowire electrodes (16 nm wide) and crossed by 400 Ti electrodes
sandwiching a monolayer of bistable [2]rotaxanes.(1513) Despite the interesting ndings,
many remain skeptical about the utility of rotaxanes in electronics, and an array of scienti c
and engineering challenges remain to be addressed such as device robustness, improved
etching tools, and improved switching speed.(1514-1516)

Figure 60
  

Figure 60. (a) Rotaxane 149-based molecular switch tunnel junctions and proposed
mechanism for the operation. (i) In the ground state, the tetracationic cyclophane (dark blue)
mainly encircles the TTF station (green) and the junction exhibits low conductance. (ii)
Application of a positive bias results in one- or two-electron oxidation of the TTF units (green
→ pink), and increases electrostatic repulsion causing (iii) shuttling of the macrocycle to the
DNP station (red). (iv) Returning the bias to near −0 V provides a high conductance state, in
which the TTF units have been regenerated, but translocation of the cyclophane has not yet
occurred due to a signi cant activation barrier to movement. Thermally activated decay of
this metastable state may occur slowly ((iv) → (i), in a temperature-dependent manner) or
can be triggered by the application of a negative voltage (v), which temporarily reduces the
cyclophane to its diradical dication form (dark blue → orange), allowing facile recovery of the
thermodynamically favored coconformation (vi). (b) Example of one design of a molecular
switch. The coloring of the functional units corresponds to that used for the structural
diagrams.(1479, 1482, 1483) Reprinted with permission from ref 14. Copyright 2007 Wiley-
VCH Verlag GmbH & Co. KGaA, Weinheim.

8.6.3 Using Mechanical Switches To Affect the Optical Properties of Materials

Molecular machines can in uence the optical properties of materials in the solid state as
well as in solution (section 8.2.2). Numerous molecular-machine-functionalized materials
where changes can be visually monitored have been reported. Some pertinent examples are
discussed below.

Liquid crystals have become widespread in numerous technological applications.


(1517-1520) The photoalignment of liquid crystals has considerable potential for display and
other optoelectronic applications. The reorienting effect of light on nematic liquid crystals is
well-known.(1521-1525) However, the effect can be greatly enhanced by the presence of a
dopant dichroic dye. Studies have suggested that the mechanism of this effect involves the
photoexcited dye molecules acting as Brownian rotors in the nematic liquid crystal.
(1526, 1527) The change in orientation was attributed to a ratchet mechanism operating via
the generation of torque. Feringa and co-workers reported a chiral nematic liquid crystal lm
doped with a previously reported chiral light-driven molecular motor. Irradiation of the  lm 
resulted in directional motion of the molecular motor, which induced a rotation of the rod-like
liquid crystal molecules. This rotation led to the alteration of the color of the lm over a large
part of the visible spectrum (Figure 61).(1528, 1529)

Figure 61

Figure 61. Color changes in a liquid-crystal lm doped with a light-driven rotor. Reprinted with
permission from ref 1528. Copyright 2002 American National Academy of Sciences.

Yamaguchi et al. reported crystalline molecular gyrotop, which showed temperature-


dependent changes in optical properties as a result of structural expansion upon rotor
acceleration (Figure 62).(1530-1532) Utilizing a phenylene moiety as the rotor, they observed
that the orientation of the rotor in the crystal was ordered below 270 K, but became
disordered above this temperature generating a slight deformation of the crystal lattice.
Using the temperature-dependent features of the crystal, the birefringence (Δn) of the crystal
could be controlled. At 280 K, the phenylene moiety undergoes a 180° ipping between two
equilibrium states, which provides an almost constant Δn. Above this temperature, Δn
decreases as the phenylene moiety rapidly rotates, and the cage expands. The dynamic and
optical properties are reversible.(1530)

Figure 62

Figure 62. (a) Molecular structure of gyrotops 150 and 151. The bulkier 151 does not exhibit
rotation. (b) X-ray structure of 150. (c) Single crystal of 150 irradiated with polarized white
light. (d) X-ray structure of 151. (e) Single crystal irradiated with polarized white light. For
150, a continuous change in color was observed, due to thermal expansion. Reprinted with
permission from ref 1530. Copyright 2012 American National Academy of Sciences.
A pseudorotaxane that acts as a thermally driven molecular switch in a crystalline state was
recently published by Horie and co-workers.(1533-1535) Crystals of the pseudorotaxane
  
underwent phase transitions upon heating with accompanying changes in optical properties.

Leigh et al. used [2]rotaxanes to control uorescence by distance-dependent intercomponent


electron transfer both in solution and on a polymer lm. The thread of these rotaxanes
included an anthracene uorophore as a stopper attached to a glycylglycine hydrogen-
bonding station and a C11 alkyl chain that could act as a second station. The macrocycles in
152 and 153·2H2+ contained nitrophenyl and pyridinium moieties, respectively, which are
known to quench the uorescence by distance-dependent electron transfer. Macrocyclic
shuttling could be induced by changes in solvent with strongly hydrogen-bonding solvents
displacing the macrocycle from the glycylglycine station to the alkyl thread. In non-hydrogen-
bonding solvents (e.g., benzene, CH2Cl2, CH3CN), the macrocycle was located on the
glycylglycine station and uorescence was completely quenched. In strongly hydrogen-
bonding solvents (e.g., DMSO and NH2CHO), the macrocycle encapsulated the alkyl chain
and uorescence was restored. Polymer/rotaxane hybrids were used to prepare transparent
lms on quartz slides. Initially, no uorescence was detected when the slides coated with a
154 containing lm were illuminated with UV light (254–350 nm). However, after the slides
were exposed to DMSO vapor, prior to illumination, a characteristic blue uorescence was
observed showing that similar shuttling processes were occurring on the polymer lm as
were observed in solution. Masking regions of the lm from the DMSO vapor allowed the
transitory etching of patterns on the polymeric lms (Figure 63).(1536)

Figure 63
  

Figure 63. (a) Chemical structures of rotaxane initiators 152 and 153 and the corresponding
PMMA-based polymers 154, 155, and 155·2H2+. (b) Images obtained by casting lms of
polymer 154 on quartz slides, then covering the lms with aluminum masks and exposing
the unmasked area to DMSO vapor for 5 min. The photographs were taken while illuminating
the slides with an 8-W UV lamp (254–350 nm). Reprinted with permission from ref 1536.
Copyright 2005 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

Polymeric 155·2H2+ lms responded to two different stimuli. Protonation of the macrocycle
by CF3CO2H vapor caused quenching of uorescence, while DMSO vapor induced shuttling
of the macrocycle and subsequently restored uorescence emission. The response of
155·2H2+ to the different combinations of two stimuli (DMSO and protons) corresponds to
an INHIBIT Boolean logic gate (Figure 64).(1536)

Figure 64
  

Figure 64. (a) Aluminum grid used in the experiment. (b) Pattern generated when lms of 155
were exposed to tri uoroacetic acid vapor for 5 min through the aluminum-grid mask. (c)
Mesh pattern obtained by rotation of the aluminum grid by 90° and exposure of the lm
shown in (b) to DMSO vapor for a further 5 min; only regions exposed to tri uoroacetic acid
but not to DMSO were quenched as shown in the magni ed view. Inset: Truth table for an
INHIBIT logic gate. The photographs were taken while illuminating the slides with an 8-W UV
lamp (254–350 nm). Reprinted with permission from ref 1536. Copyright 2005 Wiley-VCH
Verlag GmbH & Co. KGaA, Weinheim.

Hydrosol–gel systems represent another example of a nonsolution media. The effect of


doping these systems with rotaxanes has been studied by various groups.(1208, 1537-1541)
Light-driven rotaxanes based on α-cyclodextrin and cucurbit[7]uril have been dispersed in
thermoreversible hydrogel systems by Tian and co-workers. They observed that when the
rotaxanes were embedded in the hydrogel, their optical performance ( uorescence and
induced circular dichroism) improved. They attributed this observation to the restriction of
movement in the hydrogel system.(1539, 1542)

8.6.4 Using Mechanical Switches To Affect the Mechanical Properties of Materials

Piezoelectric materials provide the most common example of an external stimulus being
converted into a mechanical change in the material. Materials incorporating molecular
machines could provide a far greater level of control over macroscopic properties than is
currently possible, and their development has been the focus of great interest.(1543, 1544)
Conformational changes in polymers leading to macroscopic mechanical changes have
been extensively studied. Hydrogels, consisting of water-swollen cross-linked networks of
neutral or ionic amphiphilic polymers, have been shown to undergo reversible phase   
transitions in response to changes in temperature, solvent, pH, electric eld, or irradiation
with an up to 100-fold expansion on phase change.(1545-1558) Katchalsky et al. rst
exploited these properties to develop devices that converted chemical potential energy into
macroscopic mechanical work.(1559, 1560) A large variety of devices based on the
characteristics of these materials have been proposed(1561-1581) including guest binding
and release,(1582) control of enzyme activity,(1583) and the macroscopic motion of a
polymer gel by selective contraction of alternate sides of a gel.(1571) Finally, guest binding
by built-in recognition sites can lead to large-scale changes in the volume of the host gel.
(1584-1598)

Shape memory materials have also been designed that utilize control at a molecular level.
(1599, 1600) While most of these systems rely on a temperature variation-induced shape
change, a system utilizing a photoinitiated radical reaction to rearrange covalent cross-links
and recover the “memorized” shape has been reported.(1601) A reversible cross-linking [2 +
2] cycloaddition has also been used to achieve this control.(1602)

Conducting polymers have been used to create stimuli responsive mechanical changes.
(1603-1605) Electrochemical oxidation can induce counterion expulsion or inclusion and
thus volume change, although these changes tend to be slow in aqueous solution.
(1606-1609) Solid state,(1610) gel,(1611) and ionic liquid electrolytes have been shown to
accelerate this process.(1612) Conformationally exible calix[4]arenes have been used as
“hinges” to form conducting oligothiophenes where the cone (but not the kite) form of the
calix[4]arene maximizes the desired π–π interactions and switching could be utilized to
provide macroscopic mechanical motion.(1613, 1614)

In the previous examples, a macroscopic change was induced by the sum of multiple,
relatively uncontrolled conformational changes in a polymeric network; that is, the observed
change is not an inbuilt feature of the molecular components. The use of particular
submolecular conformational/con gurational switches to directly trigger macroscopic
changes has numerous bene ts such as overcoming problems caused by slow diffusion of
guests or solvents and ensuring a uniform response in the material. The use of
con gurational switches to control long-range order in liquid crystalline phases has been
discussed elsewhere (section 8.6.3), but a similar application in synthetic polymers often
furnishes a useful result.

Azobenzenes are often used to control reversible contraction in polymers, a process that has
been observed at a molecular level using AFM,(1615) as well as properties such as viscosity
and solubility.(1616, 1617) In the aforementioned AFM experiment, the application of a “load”
to the AFM tip and subsequent isomerization toward the cis isomer provided the rst direct
measurement of the conversion of light to mechanical energy. Chiral azobenzene side chains
can allow control over the screw sense of arti cial helical polymers with isomerization
leading to interconversion of the P and M forms.(1618-1620) Several liquid crystal polymeric
systems have been reported.(209, 1621-1635) Macroscopic motion was induced in liquid
crystal polymer springs devised by Fletcher, Katsonis et al.(1636) Here, a chiral dopant,
 a 
polymerizable liquid crystal, and a polymerizable azobenzene switch were used to generate a
liquid crystal polymeric spring. Irradiation with UV light caused cis–trans isomerization of the
azobenzene switch, which in turn led to coiling/uncoiling or helical inversion of the spring.
More complex behavior such as side to side bending could be induced by careful
manipulation of the initial helical state and directed illumination.(1636)

Light-driven directional rotor 156 has been used to drive the macroscopic motion of a glass
rod.(1637, 1638) When doped at 1 wt % into a liquid crystal, the helicity of the rotor induced a
helical arrangement in the liquid crystal, giving rise to a characteristic “ ngerprint” texture in
the liquid crystal’s surface. Irradiation altered the distribution of isomers of the rotor and thus
caused the liquid crystal’s helical nature to rearrange. This process could be observed by the
clockwise rotation it caused of a glass rod laid on the surface and of the “ ngerprint” pattern.
Eventually, however, rotation stopped as the system reached a photostationary state with the
distribution of molecules between isomers reaching equilibrium. When the light source was
removed, the isomer distribution decayed back to its initial state, with concomitant reverse
rotation. The use of the opposite isomer of the rotor led to inversion of the direction of
rotation (Figure 65).

Figure 65

Figure 65. (a) Rotor 156. (b) Rotation of micrometer scale glass rod on doped liquid crystal
lm. Photos taken at 15 s intervals. Reprinted with permission from ref 1637. Copyright 2001
Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.

Shuttling in [3]rotaxane 157 has been used to generate a macroscopic mechanical response,
deformation of a microcantilever beam, which had been coated with a monolayer of ca. 6
billion rotaxanes.(1238, 1341) Oxidation of the tetrathiafulvalene (TTF, green) station
decreased the a nity of the macrocycle for this station, and induced shuttling to the
naphthalene (red) station. A deformation of 550 nm was recorded, and the process could be
reversed and repeated (with diminishing amplitude) over several cycles (Scheme  48).  
(1238, 1341) Microcantilever deformation by guest recognition at the surface has been used
by several groups as a “read-out” from molecular detectors.(1639-1646) A polymer has been
reported that contracts upon illumination based on a directional light-driven rotor (Scheme
48).(1647) A photodriven rotary motor has also been used to induce the disassembly of self-
assembled nanotubes.(1648)

Scheme 48

Scheme 48. Molecular Motion Generating a Macroscopic Mechanical Response


(1238, 1341, 1647),a

Scheme a Reprinted with permission from ref 1647. Copyright 2015 Nature Publishing
Group.

8.6.5 Using Mechanical Switches To Affect Interfacial Properties

The ability to easily modify the properties of surfaces in a reversible manner would be
extremely valuable in the design and synthesis of technologically useful devices. Stimuli-
responsive polymer modi cation of surfaces can lead to control over wettability, adhesive
ability, porosity, patterning, and interfacial interactions; this topic has been reviewed
elsewhere.(1649-1663)
Interlocked architectures such as bistable rotaxane 158 have been used to restrict or allow
access to the pores in materials such as mesoporous silica particles.(1664) The  positively
 
charged macrocyclic ring initially prefers the TTF station, and in this conformation, diffusion
into and out of the nanopores of the silica is allowed. Oxidation of the TTF moiety causes
shuttling of the ring to the dioxynaphthalene station, and closure of the nanopores, trapping
a portion of the solvent/solute mixture. Solvent exchange and reduction of the oxidized TTF
station leads to guest release (Figure 66). The aggregation and deaggregation of a polymeric
material has been caused by the formation of pseudorotaxanes between strands. This has
been used to transiently trap nanoparticles in the pores formed upon aggregation.(1665)

Figure 66

Figure 66. (a) Structure of nanopore gate, and (b) controlled release of guest from
nanopores.(1664) Reprinted with permission from ref 14. Copyright 2007 Wiley-VCH Verlag
GmbH & Co. KGaA, Weinheim.

The translation of the macrocyclic ring of 159 has been converted into an electric signal.
(1666-1668) A gold electrode was used in place of one stopper of the rotaxane, and a
cyclodextrin ring with an attached ferrocene was used as the macrocycle. On photoactivated
shuttling, the change in rate of distance-dependent oxidation of the ferrocene moiety allowed
detection of the shuttling motion.(1669) When a glucose oxidase enzyme was attached via a
rotaxane architecture to a gold electrode, the CBPQT4+ macrocycle of the rotaxane acted as
a two-electron shuttle and allowed enzyme operation, which was prevented when the
enzyme was attached solely by the thread.(1670) A similar effect was seen in a system
involving a CdS nanoparticle in place of the enzyme in which the observed photocurrent was
ampli ed 8-fold in the rotaxane over the thread.(1671) In a simpler analogue of these
systems involving a CBPQT4+ macrocycle and a diiminobenzene station (159), electrode-
  
induced shuttling could be directly observed.(1672) The reduction in macrocycle charge on
shuttling was shown to decrease the hydrophilicity of the surface, and thus increased the
contact angle of a drop of water on the surface (Scheme 49).

Scheme 49

Scheme 49. Electrode Controlled Macrocycle Shuttling, Leading to Control of the


Hydrophilicity of the Surface(1672)

This control of wettability has been exploited by many groups to form a wide range of
interesting systems, especially in the development of “self-cleaning” surfaces.(1673-1679)
Surface wettability has been controlled by surface mounted photochromic switches.(1680)
Self-assembled monolayers of carboxylate-terminated alkanethiols on gold have been
switched between hydrophilic (carboxylate exposed, gold electrode at a negative potential)
and hydrophobic (carboxylate surface bound, gold electrode at a positive potential) states by
varying the electric potential.(1681) Bipyridinium groups have been used in place of
carboxylates to similar effect.(1682)

Several methods for the photoswitchable change of surface wettability have been developed.
Systems based on the reversible exposure of hydrophobic groups upon shuttling of a
cyclodextrin ring,(1683) or the release of a hydrophilic guest,(1684) have been successfully
exploited. Feringa and co-workers exploited a tripodal stator to x a molecular directional
rotor to a surface, which upon irradiation exposed or hid a hydrophobic uorinated chain
providing control over surface wettability (Figure 67).(1685)

Figure 67
  

Figure 67. Feringa’s tripodal wettability switch 160.(1685)

Perhaps the most powerful demonstration of the usefulness of control over wettability is in
the macroscopic transport of a liquid across a surface. The earliest example came from
Whitesides and Chaudhury who created a hydrophilicity gradient on a silica wafer by reaction
with a trichloroalkylsilane vapor and observed the motion of a droplet of water up a 15°
incline.(1677) Numerous examples of droplet motion driven by surface energy
gradients/steps have since been reported.(1686-1690) However, remotely controlling
wettability to generate droplet motion has remained challenging.(1691) This remote control
would be particularly useful in “lab-on-a-chip” type applications where it could provide a
gentle and cheap alternative to expensive microscopic pumps and strong electric elds.
(1692)

Azobenzene containing calix[4]resorcinarene 161, when deposited as a monolayer, provided


the rst example of remote control of surface energy and therefore droplet motion by
irradiation.(1693) The monolayer initially consisted of all cis isomers. However, a droplet of
olive oil has more favorable interactions with the extended trans form, where interactions
with the alkyl chain of the calix[4]arene can be maximized. As such, when an asymmetric
light source irradiated the droplet, forming more trans-calix[4]arene on one side of the droplet
than the other, a surface energy gradient was created and the droplet moved away from the
trans enriched area. This movement could be continued if the light source followed the rear
edge of the droplet (Figure 68).
Figure 68
  

Figure 68. Calix[4]arene 161, used to control surface wettability.(1693)

Photoresponsive control of droplet motion has also been achieved using a system based on
bistable rotaxane 162, adsorbed onto a gold surface using a thiol linker.(1245)
Diiodomethane drops could be transported on a millimeter scale across a surface using the
macrocycle of 162 to either hide or expose the uoroalkane region of the rotaxane and thus
modify surface energy. Using this strategy, a microliter droplet could be moved up a 12°
incline. Roughly 50% of the absorbed photon energy was converted into gravitational
potential energy of the drop (Figure 69).

Figure 69
  

Figure 69. Light switchable rotaxane 162, and transport of a microliter drop of CH2I2 across a
at surface (a–d) and up a 12° incline (e–h). Reprinted with permission from ref 1245.
Copyright 2005 Nature Publishing Group.

The above examples of macroscopic control utilizing molecular motion underscore the
potential importance of this concept for technological applications. Similar control of motion
has been achieved with surfaces microscopically patterned with thermoresponsive polymers
(1694-1696) and by amplifying the effect of the photochemical switching of a spiropyran-
functionalized surface.(1697)

9 Arti cial Biomolecular Machines Jump To

9.1 Hybrid Biomolecular Systems


Biological systems make extensive use of motor proteins such as ATPase, kinesin, myosin,
and dynein. These biological motors provide both examples of machines viable at the
nanoscale and useful building blocks for the creation of hybrid systems. The availability of
preformed nanoscale machines has greatly expedited the creation of complex
nanotechnological systems. Although a thorough exploration of this topic is beyond the
purview of this Review and has been covered elsewhere,(15, 47, 1364, 1698-1702) a brief
overview is useful. The initial purpose of these biohybrid experiments was to further probe
the complex mechanisms by which biological machines operate. An early example was
provided by Kinosita et al., who reported a series of experiments exploring the F1-ATPase
rotary motor.(1703-1707) They were able to observe the 360° rotation of the motor, powered
by the hydrolysis of ATP, directly, via microscopy, for individual motors mounted on glass.
Mechanical rotation has also been observed in an F0F1-ATPase, where the rotor domain is
coupled to a proton transporting domain.(1708) Even more remarkably, ATP synthesis has
been driven by rotating a magnetic bead attached to an F1-ATPase with electromagnets.
(1709) This is a stunning example of chemical synthesis driven by an external mechanical
force. ATPase systems have been less exploited in recent years due to the di culties in
using them in synthetic systems.(1710) However, a recent example showed the insertion of
ATP synthases into an arti cial lipid microcapsule and their subsequent use to generate
 ATP 
inside the capsule.(1711)

Kinesin, myosin, and dynein systems have been much more extensively utilized.(1712-1722)
They have been shown to transport cargos in arti cial systems, although cargos must
typically be large to outcompete diffusive forces.(1375, 1723-1727) These biological walkers
have been controlled by the design of appropriate tracks,(1728-1730) or by the application of
external forces. Kinesin has been used to stretch a length of DNA,(1715) and to enable the
detection of nanomolar concentrations of a targeted protein.(1731) Cross-linked actin and
myosin gels were found to move over each other in the presence of ATP.(1732) Microtubule
containing gels combined with kinesin and ATP have been shown to spontaneously generate
autonomous motility.(1733, 1734) As a further bene t to using biological motors, ATPase
motors have been shown to operate at a near 100% e ciency,(1707, 1735-1737) and kinesin
at >50%,(1738) whereas in a typically used polymeric system where phase transitions are
used to generate molecular level forces, an e ciency of 0.0001% was observed. The eld of
biomolecular electronics often uses molecular motion as a switching mechanism,
particularly in systems utilizing photoactive proteins (be they native or engineered).
(1739-1750)

9.2 Hybrid Membrane-Bound Machines


Ionophores and ion channels, both synthetic(1751-1755) and biological,(1756-1766) have
been extensively explored and exploited. Biological ion channels have been inserted into
arti cial membranes and used to sequence single strands of DNA,(1756) to open the pore of
a nonselective e ux channel,(406, 1767, 1768) and to effect the light-driven production of
ATP by F0F1-ATPase.(1769) Feringa et al. reported the switching of the “mechanosensitive
channel of large conductance’ (MscL) of E. coli, which can be opened in response to the
introduction of charged entities at a certain location in the protein channel.(406) Initially, a
light-cleavable group was used to release acetate anions and thus open the channel, with a
reversible version, utilizing a spiropyran switch, also reported (Scheme 50). Ion channel
proteins modi ed postsynthetically to be light switchable have been inserted into living cells,
and successfully opened and closed.(1770-1777) Unnatural, light switchable amino acids
have also been incorporated into ion channel proteins.(400, 1778-1780)

Scheme 50
  

Scheme 50. (a) Irreversible Photocleavage of 163 Leading to Pore Opening; and (b)
Reversible Photoswitching of 164, Leading to Pore Openinga

Scheme a MscL = mechanosensitive channel.(406)

Active transport between aqueous phases across an organic phase,(1781-1783) and lipid
bilayers,(1784) has been observed, driven by differing redox potentials in each aqueous
phase. An electron donor–acceptor molecule has been directionally inserted into a lipid
bilayer, which on irradiation created a redox potential gradient and ferried protons across the
bilayer establishing a pH gradient. However, the quantum yield for the process was only
0.004.(1785) Calcium ions have been transported across a membrane in a similar system.
(1786) Work to establish longer lived photoinduced charge separated states, which might
lead to greater quantum yields and further applications, has continued.(627, 1787-1807)
Alternating currents have been induced in a protein-based photoelectrochemical cell by
irradiation with intermittent light.(1808) An interesting recent example utilized the light-
induced ring-opening of spiropyrans under ultraviolet light and their ring closure under visible
light illumination. When a membrane doped with spiropyran 165 was illuminated with
ultraviolet light on one side and visible light on the other, the differing proton a nities of ring-
opened and ring-closed spiropyran led to proton shuttling across the membrane and the
generation of an electric potential of ca. 210 mV and pH gradient of ca. 3.6 pH units (Figure
70).(1809)

Figure 70
  

Figure 70. Proton gradient established by spiropyran (165) shuttling upon differential
illumination of the two sides of a membrane. Reprinted with permission from ref 1809.
Copyright 2014 Nature Publishing Group.

9.3 DNA-Based Motors and Switches


Biological building blocks have been used to design and create molecular machines by many
research groups.(1810-1828) A large number of DNA-based molecular motors,(1829-1832)
walkers,(1382, 1833-1838) tweezers,(1839) gears,(1840) springs,(1841) robots,(1842, 1843)
transporters,(1844) and interlocked structures such as DNA rotaxanes and catenanes can be
found in the literature.(1845-1850) All are built by self-assembly, exploiting the sequence-
speci c interactions that bind complementary oligonucleotides together to form double helix
or triplex structures.(1851-1870) Beside base-pairing, other structural motifs can be formed
and used in the design of molecular machines such as the pH-induced self-assembly of C-
rich sequences into i-motif con gurations,(1871, 1872) the ion-induced self-organization of
G-rich sequences into G-quadruplexes,(1873-1876) and the metal–ion bridging of duplex
DNA by T–Hg2+–T or C–Ag+–C complexes.(1877, 1878) Various fuels such as single-
stranded oligonucleotide fragments, pH variation, metal ions, and light have been used to
trigger these DNA devices.(1879-1884) These systems have been used in molecular sensing,
drug delivery and other medical applications, the construction of logic gates, the control of
chemical transformations, and for many other purposes.(1874, 1885-1911)
DNA tweezers represent a simple class of DNA machines.(1884, 1912) They are two-armed
constructs bridged by a DNA linker that can undergo transitions between open and closed
 
states in response to external triggers such as the addition of single-stranded DNA or metal
ions, or a change in pH. Willner et al. reported a biomolecular logic gate based on three
different DNA tweezers A, B, and C. These were activated by different inputs: protons, Hg2+
ions, and nucleic acid strands (Figure 71). The output delivered by this machine depends
both on the inputs provided and its initial internal state. Depending on the input, there are
eight possible con gurations of the three tweezers (open or closed for each). The output
could be studied by measuring the Förster resonance energy transfer (FRET) between
different pairs of uorophore and quenching molecules attached to the arms of each of the
three tweezers. The linker unit is common to all three tweezers, meaning that tweezers A and
B can also be opened by the complementary antilinker. Thus, for any pair of tweezers, there
are two different inputs that cause a change in the state of the device. In total, the device can
adopt 16 different states and can furthermore be used as a memory storage system
because each state and output is dependent not only on the most recent input but also on
past states and inputs.(1913, 1914)

Figure 71

Figure 71. (a) Tweezer A: in the closed form the arms are bound to the linker unit (blue) by
Hg2+ ions through T–Hg2+–T bonds. To open the molecular tweezer, Hg2+ is sequestered by
the addition of cysteine. (b) Tweezer B: in acid the arms form an i-motif, thus releasing the
linker unit, whereas at pH = 7.2, the i-motif is destroyed resulting in the stabilization
of 
the 
closed structure. (c) Tweezer C: the linker unit can be released by a complementary strand,
the antilinker that opens the tweezers. Reprinted with permission from ref 1914. Copyright
2010 American National Academy of Sciences.

DNA machines have been used to regulate enzyme cascade reactions.(1890, 1891) Liu and
co-workers reported a machine containing DNA double crossover (DX) motifs, which formed
two rigid arms, joined by an immobile four-way junction (Figure 72).(1915) A DNA motor that
could switch between stem-loop and double-helix structures, driven by a strand
displacement reaction, was incorporated at the center of the machine to cycle between open
and closed states. This design ampli es the small motion generated by the DNA motor into a
much greater change in separation between the ends of the two arms where glucose
oxidase (GOx) and horseradish peroxidase (HRP) were attached. In this biochemical cascade
system, GOx rst catalyzed the oxidation of glucose to generate gluconic acid and hydrogen
peroxide. Hydrogen peroxide is catalytically reduced by HRP into H2O. At the same time, HRP
turns ABTS2– into ABTS–, which allowed the kinetics of the peroxidase to be monitored. HRP
has a much higher turnover rate then GOx, so the distance hydrogen peroxide must diffuse
has a crucial in uence on the rate of reaction. Therefore, when the two enzymes are
attached to the two arms, the diffusion distance of hydrogen peroxide can be changed from
6 to 18 nm by operation of the DNA motor, regulating the rate of enzymatic reaction.
Sequential addition of fuel and antifuel strands showed that this regulation was reversible.

Figure 72

Figure 72. DNA machine reported by Liu and co-workers, which could be used to regulate an
enzyme cascade reaction. Reprinted with permission from ref 1915. Copyright 2013 Wiley-
VCH Verlag GmbH & Co. KGaA, Weinheim.

Willner and co-workers have used catenated DNA machines as carriers of Au nanoparticles.
(1916) Ordered assemblies of nanoparticles with de ned geometries have recently attracted
a great deal of interest as these engineered nanoparticle systems are anticipated to show
unique plasmonic properties.(1917, 1918) The synthesis and structural characterization of
DNA catenanes has been reported previously,(1847, 1849, 1919, 1920) as has a three-station
directional DNA catenane rotary motor.(1921) A three-ring catenated nanostructure has been
used to program the assembly of two Au nanoparticles (of 5 and 10 nm diameter, Figure 73).
One of the nanoparticles was attached to ring α and the other to ring γ (L1). Ring α included
 
two equivalent domains, I and II, which are complementary to domain III on ring γ. While the
initial duplex between rings β and γ in region III is energetically favored, treatment of the
linear catenated system L1 with a fuel strand (Fu) displaces ring γ, which then translocates
through hybridization to either site I or II of ring α (structures M1 or M2). In the presence of
the appropriate blocker units (B), the directional translocation of ring γ proceeds below or
above the central ring β to yield the con gurations M3 or M4, respectively. The changes in
proximity of the nanoparticles led to changes in plasmonic coupling, which was theoretically
modeled. Such systems could have uses in nanomedicine and intracellular diagnostics.
(1916)

Figure 73

Figure 73. (a) Programmed migration of two Au nanoparticles. (b) STEM images
corresponding to the different structures; the bar corresponds to 20 nm. Reprinted with
permission from ref 1916. Copyright 2013 Nature Publishing Group.

9.4 DNA Walkers


The controlled and predictable strand displacement of DNA and RNA has been used to
construct complex devices.(1369, 1869, 1922-1932) Among them, walkers represent a major
subclass. The basic principle behind 1-D walking is illustrated in Figure 74. A relatively rigid
and stable DNA track functionalized with protruding single-stranded anchorage points
(footholds) is typically used.(1369, 1933) Unlike the direct interactions of biological walker
proteins with polymeric tracks, the footholds are attached to the track and bind the walker
unit by base paring. The walker unit consists of regions with different nucleotide sequences
(feet), which are attached to the track via another DNA fragment (the anchor strand), that can
hybridize with both the walker and one of the footholds leaving a “toehold region” unpaired.
Addition of the fuel strand results in base-pairing with an anchor strand, starting from the
protruding toehold, and forms a more thermodynamically favored waste duplex. The removal
of the anchor strand liberates the foothold and one foot of the walker. Walking is processive
because the other foot is still attached to the track by a holding strand, which provides
additional stabilization. The walker can be reattached to the track via the addition of another
anchor strand.   

Figure 74

Figure 74. Schematic representation of a DNA-based walker and track system. Footholds
protrude from the track as single-stranded DNA fragments. Anchor and holding strands
enable the walker unit to bind to the footholds by hybridization. Areas with functional
importance are labeled, and complementary strands are depicted in the same color for
clarity.

Stimuli-dependent positional DNA switches(1934, 1935) and DNA-based devices performing


multistep organic synthesis while migrating along a track have been reported.(1447) An in
silico “tumbleweed” walker design has also been proposed.(1936) The rst DNA-based
walker was nonautonomous and bipedal and was reported by Sherman and Seeman.(1834)
High dilution conditions were used to prevent the walker from scrambling betwen different
tracks. Sequential addition of two different anchor and fuel strands in an aqueous buffer at
16 °C led to the desired walking motion being obtained. The products were characterized by
polyacrylamide gel electrophoresis (PAGE). The energy required for directional walking was
provided by the additional base-pairing in the waste duplex (red toehold region, Figure 74).
Mechanistically, this walker is an inchworm walker because the leading foot remains the
same throughout the operation. A hand-over-hand DNA walker (the mechanism of operation
of kinesin) has been published by Shin and Pierce who used a similar design.(1833)
Transport of a cargo over a DNA origami tile and the synthesis of nanoparticle sequences
have been reported. It used a DNA origami walker unit with four “feet” for controlled
movement, and three “arms” for picking up cargo, each consisting of single strands of DNA
(Figures 75 and 76).(1455, 1937, 1938) Fuel strands (Fi) were used to drive the motion and to
remove the anchor strands (Ai). Each station was loaded with a distinct gold nanoparticle
cargo and could be switched between states where cargo delivery was possible and where it
was not. By manipulating the selective release of each foot from complementary strands of
DNA on the DNA tile, the movement of the walker could be controlled. When coupled to the
ability of the stations to be switched “on” or “off”, this allowed the formation of eight,
differently composed, noncovalently bound products from the full operational sequence. This
remarkable level of control on the nanoscale shows that the forces of Brownian motion can
be exploited to great effect in the synthesis of complex supramolecular products.
Transportation of a DNA cargo on a DNA origami tile over 16 consecutive steps has
 also
 
been reported.(1939)

Figure 75

Figure 75. (a) Structure of the DNA walker with four “feet” (F1–4) and three “hands” (H1–3).
(b) Movement of walker across the DNA origami tile driven by sequentially added DNA “fuel”
strands labeled FA. (c) Loading of cargo onto DNA walker. Reprinted with permission from
ref 1455. Copyright 2010 Nature Publishing Group.

Figure 76
  

Figure 76. (a) Operation of DNA-based walker. (b) AFM images of walker. Reprinted with
permission from ref 1455. Copyright 2010 Nature Publishing Group.

Turber eld, Reif, and Yan have reported autonomous walking in a DNA-based system.(1379)
Three enzymes, the PfIM I and BstAP I restriction enzymes, and T4 ligase, were required for
autonomous operation. Successive cleavage and ligation of the walker-foothold duplex
resulted in directional walking along the track. Two autonomous burnt-bridges walkers have
been reported using either a restriction endonuclease as an additive(1940) or a walker unit
with intrinsic DNAzyme activity.(1380, 1941, 1942) An autonomous DNA motor whose
propulsion was driven by random polymerization has been synthesized.(1829) DNA strands
were propeled at the growing end of the polymer by the energy gained from hybridization. A
number of enzyme-free autonomous DNA walkers have since been published.
(1836, 1943-1945) Autonomous multipedal walkers(1842, 1946) and walkers performing
autonomous and progressive acylation reactions have been disclosed.(1447) Autonomous
DNA walkers that are driven by pyrene-mediated photocleavage of disul de bonds,(1947) or
photoisomerization-dependent hindrance/exposure of a complementary strand,(1383) have
been designed.

A vital requirement for useful cargo transport is the ability to choose the correct path at a
junction point. The walker should be capable of crossing the junction, and the path choice
should be programmable. Turber eld et al. were able to autonomously regulate the transport
of a DNA cargo on a branched track (Figure 77).(1843) Successive Holliday junctions formed
between the selected fuel, foothold, and the cargo as the walker migrated. These led to
strand exchange (equivalent to junction migration), followed by loop opening, which allowed
the controlled migration of the cargo from one foothold to another. The same group have
since used a single nucleoside, adenosine, to control the route taken at the track junction.
(1381)   

Figure 77

Figure 77. Walker with a choice of path at a junction. (a) Initially cargo resides on position W,
and (b) the use of different set of fuels (F) leads to transport of cargo to either position. A
displacing strand (Dz) was used for the fragmentation of the molecular ensemble and
subsequent analysis of the cargo position. Reprinted with permission from ref 1843.
Copyright 2011 American Chemical Society.

10 Conclusion and Outlook Jump To

Perhaps the best way to appreciate the technological potential of controlled molecular-level
motion is to recognize that molecular machines lie at the heart of every signi cant biological
process. Over billions of years of evolution, nature has not repeatedly chosen this solution for
achieving complex task performance without good reason. In stark contrast to biology, none
of mankind’s fantastic myriad of present-day technologies exploit controlled molecular-level
motion in any way at all: every catalyst, every material, every polymer, every pharmaceutical,
and every reagent all function through their static or equilibrium dynamic properties. When
we learn how to build arti cial structures that can control and exploit molecular-level motion,
and interface their effects directly with other molecular-level substructures and the outside
world, it will potentially impact on every aspect of functional molecule and materials
 design.
 
An improved understanding of physics and biology will surely follow.

As indicated by the many examples that appear in this Review, the future for the eld of
arti cial molecular machines is very bright.(1948) After a somewhat di cult period in the
1990s, when chemists struggled to understand the basic differences between machines at
the macroscopic and nanometer length scales, scientists now have the know-how and
synthetic tools available to enable them to make suitable machine architectures (e.g.,
catenanes, rotaxanes, overcrowded alkenes, molecules that walk upon tracks, etc.). They can
switch the position of components (often by clever manipulation of noncovalent interactions
between the various parts), they understand how to use ratchet mechanisms to create
motor-mechanisms, and they are learning how to introduce them into more complex
molecular machine systems.

Yet there are still several basic challenges to overcome for molecular machines to be able to
ful ll their potential:

(i) In contrast to motor proteins, powered by ATP hydrolysis or proton gradients, there are as
yet no chemically driven synthetic small-molecule motors that can operate autonomously
(i.e., move or rotate directionally as long as a chemical fuel is present), the closest counter-
examples being the Feringa overcrowded alkene motors that rotate continuously under
irradiation with light.(429-453)

(ii) Most of the synthetic molecular machines reported to date are based on only a single
functioning part that moves, stops, uoresces, catalyzes a reaction, etc. To perform tasks
that cannot be accomplished by conventional chemical means, it will be necessary to design
systems with multiple integrated parts, each component performing a dedicated role within
the machine ensemble. This will not be straightforward because unlike a watch where the
second hand, say, does not interfere with the components in the escapement mechanism,
the components of a chemical machine are not easily isolated from each other (or the
environment) and interference from one reactive part of a machine with another will be a
signi cant issue as complexity increases beyond the current rather trivial systems.

(iii) The machines we are familiar with in the macroscopic world are generally stable,
operating unchanged through many cycles, and by and large they do not make “mistakes”.
This contrasts with those in the biological world where the stochastic nature of molecular
dynamics mean that motors stall, or step backward, or detach from tracks, or make errors in
synthesis that are spotted or corrected by other machines. These are intrinsic differences
that require fundamentally different philosophies in terms of the way machines carry out
tasks that chemists have not yet started to tackle.

(iv) The “reading” of a sequence of functional groups on a polymer stand (mRNA or DNA) is
how biological molecular machines are programmed to carry out synthesis operations in the
correct sequence. Although it is possible to use biological polymers to do this, molecular
machines made from DNA are obviously far more limited in terms of operating conditions,
chemical stability, and functionality than wholly synthetic systems. As yet there is
no small-
 
molecule “Turing machine”,(1949) that is, a molecular machine that can read information
from a symbol-encoded molecular strand. Programming of small-molecule machines with
an information-rich non-DNA strand (rather than the sequential addition of chemical stimuli,
say) would be a game-changing development.

There are also issues on which protagonists in the eld have differing points-of-view. The
debate continues over whether it will ultimately prove more productive to build molecular
machines based on macroscopic objects (e.g., Stoddart’s “molecular elevator”(730) and
“molecular pistons”(527) and Tour’s “nanocars”,(1129-1132) etc.) or based on biological
machines (e.g., arti cial systems that seek to reproduce aspects of the behavior of kinesin
(1398-1400) or the ribosome(666)). Certainly mimicking biology is not the only way to
achieve complex functionality: computer chips are manufactured from silicon wafers rather
than being wet and carbon-based like our brains. Yet equally machines have to be designed
according to the environment they are intended to operate in, and there may be reasons why
biology uses motors and not switches, and tracks and not wheels, to transport molecular
cargoes. Or it may be that evolution just did not discover these solutions to such problems
and that mankind, with the whole of the periodic table and known synthetic chemistry to
work with, can. Perhaps the most productive approach will ultimately be found by following
neither of these lines of investigation too closely, for example, by using chemical principles
for “molecular robotics” in which ratcheted motions of molecular components (i.e.,
biologically inspired mechanisms) are used to perform tasks that have their origins in
innovations introduced to advance developments in macroscopic technology (e.g., factory
assembly lines).

The environments that molecular machines can most productively operate in also remain an
open question. Efforts to incorporate molecular machines into regular arrays within
crystalline solids (e.g., MOFs) are progressing apace and should allow tethered molecular
machines to interact effectively with solution- or gas-phase substrates. It will be interesting
to see what applications such systems are being tailored for. Molecular machines mounted
on surfaces should allow for surface properties to be changed in a facile and effective
manner using a minimal amount of the high-cost molecular structures. However, whether
monolayers of molecular machines can be su ciently robust will no doubt be an issue for
practical applications. Most biological molecular machines operate while dissolved in
solution, a situation that has no equivalent for macroscopic mechanical machines, and so
chemists will have to consider carefully how to effectively address and utilize such machines
in that environment. Molecular machines have the potential to act productively under each of
these conditions; what is important is that the application is appropriate for which to use
synthetic molecular machines. In most cases, the cost of synthesis will mean that this will
only be the case if there is no way to carry out the task without using a molecular machine.
So rotaxane-based “molecular muscles” have to compete with shape-memory polymers (or
even stilbene-containing polymers) to be effective, and rotaxane-based switches, which work
through the physical displacement of submolecular components, will have to compete with
electron movements in silicon for applications. Chemical synthesis is one area where
arti cial molecular machines may one day be able to perform tasks that are not possible
  to
achieve using conventional chemical methods, a potential “killer app”!

A quarter-of-a-century after chemists made the rst edging molecular protomachines, the
tantalizing prospect of arti cial molecular machines that can perform useful tasks is moving
ever closer to becoming a reality. Molecular machines with multiple integrated parts are in
design terms a fusion of the familiar with the strange. Examples of mechanical engineering
from the world around us provide a conceptual framework for what we want tiny machines
to achieve. Biology shows us how machines cope with the nature of the environment in
performing tasks at nanometer length scales, and physics explains the often counterintuitive
ways that such small objects must behave. Yet ultimately, while drawing on all of these
disciplines for ideas, guidance, and inspiration, such machines have to be designed, built, and
operated through chemistry, the central science.

Special Issue
This paper is an additional review for Chem. Rev. 2015, 115, 15, “Supramolecular Chemistry”.

The authors declare no competing nancial interest.

Acknowledgment Jump To

We thank the ERC and EPSRC for nancial support and the European Union Seventh
Framework Marie Curie Intra-European Fellowship Program and the Swiss National Science
Foundation for Postdoctoral Fellowships (to S.E.-C. and A.L.N., respectively).

Abbreviations

AB azobenzene

ABTS 2,2′-azino-bis(3-ethylbenzothiazoline-6-sulfonic acid)

AFM atomic force microscope

ATP adenosine triphosphate

BODIPY boron-dipyrromethene
Bpy bipyridine
  
CB cucurbituril

CBPQT cyclobis(paraquat-p-phenylene)

CBS Corey–Bakshi–Shibata

CD circular dichroism or cyclodextrin

CuAAC copper-catalyzed azide–alkyne cycloaddition

Cys cysteine

D displacing strand

DAE diarylethene

DBU 1,8-diazabicycloundec-7-ene

DMAP 4-dimethylaminopyridine

DMF dimethylformamide

DMSO dimethyl sulfoxide

DNA DNA

DNP dioxynaphtalene

DTT dithiothreitol

ee enantiomeric excess

ESIPT excited-state intramolecular proton transfer


EXSY exchange spectroscopy
  
F fuel strand

FRET Förster resonance energy transfer

GOx glucose oxidase

GSCC ground-state coconformation

HBT 2-(2-hydroxyphenyl)-benzothiazole

HOMO highest occupied molecular orbital

HPLC high-performance liquid chromatography

HRP horseradish peroxidase

iPr2NEt N,N-diisopropylethylamine

IR infrared

LB Langmuir–Blodgett

LUMO lowest unoccupied molecular orbital

MOF metal organic framework

Mc merocyanine

MS mass spectrometry

MSCC metastable state coconformation

MscL large-conductance mechanosensitive channel


MSTJ molecular switch tunnel junction
  
MV methyl viologen

NCL native chemical ligation

NMR nuclear magnetic resonance

XPS X-ray photoelectron spectroscopy

PAGE polyacrylamide gel electrophoresis

PEO poly(ethylene oxide)

PLP pyridoxal 5′-phosphate

PMB p-methoxybenzyl

PMMA poly(methyl methacrylate)

R Reynolds number

RET resonance energy transfer

RNA ribonucleic acid

SAM self-assembled monolayer

SEM scanning electron microscope

Sp spiropyran

STM scanning tunneling microscope

T temperature
TBDMSCl tert-butyldimethylsilyl chloride
  
TCBD tetracyanobutadiene

TFA tri uoroacetic acid

TNF 2,4,7-trinitro-9- uorenone

TON turnover number

TTF tetrathiafulvalene

UV ultraviolet

References Jump To

This article references 1949 other publications.

1. Brown, R. A Brief Account of Microscopical Observations Made on the Particles Contained in the Pollen of
Plants Philos. Mag. 1828, 4, 171– 173 DOI: 10.1080/14786442808674769 [Crossref], [Google Scholar]

2. Brown, R. On the Particles Contained in the Pollen of Plants; and on the General Existence of Active Molecules
in Organic and Inorganic Bodies Edinb. New Philos. J. 1828, 5, 358– 371 [Google Scholar]

3. Perrin, J. In Atoms (English Translation), 2nd ed.; Hammick, D. L., Ed.; Constable and Co.: London, 1923.
[Google Scholar]

4. Einstein, A. Über die von der Molekularkinetischen Theorie der Wärme Geforderte Bewegung von in Ruhenden
Flüssigkeiten Suspendierten Teilchen Ann. Phys. 1905, 17, 549– 560 DOI: 10.1002/andp.19053220806
[Crossref], [CAS], [Google Scholar]

5. Feynman, R. P.; Leighton, R. B.; Sands, M. The Feynman Lectures on Physics; Addison-Wesley: Reading, MA,
1963; Vol. 1, Chapter 46. [Google Scholar]

6. Smalley, R. E.; Drexler, K. E. Nanotechnology Chem. Eng. News 2003, 81, 37– 42 [Google Scholar]
7. Balzani, V.; Credi, A.; Raymo, F. M.; Stoddart, J. F. Arti cial Molecular Machines Angew. Chem., Int. Ed. 2000,
39, 3348– 3391 DOI: 10.1002/1521-3773(20001002)39:19<3348::AID-ANIE3348>3.0.CO;2-X [Crossref],
  
[PubMed], [CAS], [Google Scholar]

8. Credi, A. Arti cial Molecular Motors Powered by Light Aust. J. Chem. 2006, 59, 157– 169
 DOI: 10.1071/CH06025 [Crossref], [CAS], [Google Scholar]

9. Paxton, W. F.; Sundararajan, S.; Mallouk, T. E.; Sen, A. Chemical Locomotion Angew. Chem., Int. Ed. 2006, 45,
5420– 5429 DOI: 10.1002/anie.200600060 [Crossref], [PubMed], [CAS], [Google Scholar]

10. Tian, H.; Wang, Q.-C. Recent Progress on Switchable Rotaxanes Chem. Soc. Rev. 2006, 35, 361– 374
 DOI: 10.1039/b512178g [Crossref], [PubMed], [CAS], [Google Scholar]

11. Leigh, D.; Pérez, E. Dynamic Chirality: Molecular Shuttles and Motors. In Supramolecular Chirality; Crego-
Calama, M.; Reinhoudt, D., Eds.; Springer: Berlin, Heidelberg, 2006; Vol. 265, pp 185– 208. [Crossref], [Google
Scholar]

12. Hess, H. Self-assembly Driven by Molecular Motors Soft Matter 2006, 2, 669– 677 DOI: 10.1039/b518281f
[Crossref], [CAS], [Google Scholar]

13. Browne, W. R.; Feringa, B. L. Making Molecular Machines Work Nat. Nanotechnol. 2006, 1, 25– 35
 DOI: 10.1038/nnano.2006.45 [Crossref], [PubMed], [CAS], [Google Scholar]

14. Kay, E. R.; Leigh, D. A.; Zerbetto, F. Synthetic Molecular Motors and Mechanical Machines Angew. Chem., Int.
Ed. 2007, 46, 72– 191 DOI: 10.1002/anie.200504313 [Crossref], [PubMed], [CAS], [Google Scholar]

15. Balzani, V.; Credi, A.; Venturi, M. Molecular Machines Working on Surfaces and at Interfaces
ChemPhysChem 2008, 9, 202– 220 DOI: 10.1002/cphc.200700528 [Crossref], [PubMed], [CAS], [Google Scholar]

16. Ma, X.; Tian, H. Bright Functional Rotaxanes Chem. Soc. Rev. 2010, 39, 70– 80 DOI: 10.1039/B901710K
[Crossref], [PubMed], [CAS], [Google Scholar]

17. Wang, J.; Manesh, K. M. Motion Control at the Nanoscale Small 2010, 6, 338– 345
 DOI: 10.1002/smll.200901746 [Crossref], [PubMed], [CAS], [Google Scholar]
18. Coskun, A.; Banaszak, M.; Astumian, R. D.; Stoddart, J. F.; Grzybowski, B. A. Great Expectations: Can Arti cial
 
Molecular Machines Deliver on Their Promise? Chem. Soc. Rev. 2012, 41, 19– 30 DOI: 10.1039/C1CS15262A 
[Crossref], [PubMed], [CAS], [Google Scholar]

19. Yang, W.; Li, Y.; Liu, H.; Chi, L.; Li, Y. Design and Assembly of Rotaxane-Based Molecular Switches and
Machines Small 2012, 8, 504– 516 DOI: 10.1002/smll.201101738 [Crossref], [PubMed], [CAS], [Google Scholar]

20. Neal, E. A.; Goldup, S. M. Chemical Consequences of Mechanical Bonding in Catenanes and Rotaxanes:
Isomerism, Modi cation, Catalysis and Molecular Machines for Synthesis Chem. Commun. 2014, 50, 5128–
5142 DOI: 10.1039/c3cc47842d [Crossref], [PubMed], [CAS], [Google Scholar]

21. van Dongen, S. F. M.; Cantekin, S.; Elemans, J. A. A. W.; Rowan, A. E.; Nolte, R. J. M. Functional Interlocked
Systems Chem. Soc. Rev. 2014, 43, 99– 122 DOI: 10.1039/C3CS60178A [Crossref], [PubMed], [CAS], [Google
Scholar]

22. Zhang, M. M.; Yan, X. Z.; Huang, F. H.; Niu, Z. B.; Gibson, H. W. Stimuli-Responsive Host-Guest Systems
Based on the Recognition of Cryptands by Organic Guests Acc. Chem. Res. 2014, 47, 1995– 2005
 DOI: 10.1021/ar500046r [ACS Full Text ], [CAS], [Google Scholar]

23. Xue, M.; Yang, Y.; Chi, X. D.; Zhang, Z. B.; Huang, F. H. New Class of Macrocycles for Supramolecular
Chemistry Acc. Chem. Res. 2012, 45, 1294– 1308 DOI: 10.1021/ar2003418 [ACS Full Text ], [CAS], [Google
Scholar]

24. Gil-Ramirez, G.; Leigh, D. A.; Stephens, A. J. Catenanes Fifty Years of Molecular Links Angew. Chem., Int. Ed.
2015, 54, 6110– 6150 DOI: 10.1002/anie.201411619 [Crossref], [PubMed], [CAS], [Google Scholar]

25. Colinvaux, P. Life at Low Reynolds-Number Nature 1979, 277, 353– 354 DOI: 10.1038/277353a0
[Crossref], [Google Scholar]

26. Purcell, E. M. Life at Low Reynolds-Number Am. J. Phys. 1977, 45, 3– 11 DOI: 10.1119/1.10903
[Crossref], [Google Scholar]

27. Shenker, O. R. Maxwell’s Demon 2: Entropy, Classical and Quantum Information, Computing Stud. Hist.
Philos. M. P. 2004, 35B, 537– 540 DOI: 10.1016/j.shpsb.2004.04.001 [Crossref], [Google Scholar]

28. Maxwell, J. C. Theory of Heat; Longmans, Green and Co.: London, 1871; Chapter 22. [Google Scholar]
173 
29. Smoluchowski, M. S. On Opalescence of Gases in the Critical State Philos. Mag. 1912, 23, 165–
 DOI: 10.1080/14786440108637209 [Crossref], [Google Scholar]

30. Feynman, R. P.; Vernon, F. L. The Theory of a General Quantum System Interacting with a Linear Dissipative
System Ann. Phys. 1963, 24, 118– 173 DOI: 10.1016/0003-4916(63)90068-X [Crossref], [Google Scholar]

31. Ehrenberg, W. Maxwell’s Demon Sci. Am. 1967, 217, 103– 110 DOI: 10.1038/scienti camerican1167-103
[Crossref], [Google Scholar]

32. For the rst private written discussion of the “temperature demon”, see:
Maxwell, J. C. Letter to P. G. Tait, 11 December 1867; quoted in C. G. Knott, Life and Scienti c Work of Peter
Guthrie Tait; Cambridge University Press: London, 1911; p 213; [Google Scholar]
and reproduced in: The Scienti c Letters and Papers of James Clerk Maxwell 1862; Harman, P. M., Ed.;
Cambridge University Press: Cambridge, 1995; Vol. II, p 331. [Google Scholar]

33. Szilard, L. On the Minimization of Entropy in a Thermodynamic System with Interferences of Intelligent
Beings Eur. Phys. J. A 1929, 53, 840– 856 DOI: 10.1007/BF01341281 [Crossref], [Google Scholar]

34. Bennett, C. The Thermodynamics of Computation—A Review Int. J. Theor. Phys. 1982, 21, 905– 940
 DOI: 10.1007/BF02084158 [Crossref], [CAS], [Google Scholar]

35. Landauer, R. Irreversibility and Heat Generation in the Computing Process IBM J. Res. Dev. 1961, 5, 183–
191 DOI: 10.1147/rd.53.0183 [Crossref], [Google Scholar]

36. Berut, A.; Arakelyan, A.; Petrosyan, A.; Ciliberto, S.; Dillenschneider, R.; Lutz, E. Experimental Veri cation of
Landauer’s Principle Linking Information and Thermodynamics Nature 2012, 483, 187– 189
 DOI: 10.1038/nature10872 [Crossref], [PubMed], [CAS], [Google Scholar]

37. The idea of a “pressure demon” was introduced by Maxwell in a later letter to Tait (believed to date from
early 1875); quoted in: Knott, C. G. Life and Scienti c Work of Peter Guthrie Tait; Cambridge University Press:
London, Cambridge, 1911; p 214; and reproduced in: The Scienti c Letters and Papers of James Clerk Maxwell
1874; Harman, P. M., Ed.; Cambridge University Press: London, Cambridge, 2002; Vol. III, p 185. [Google Scholar]

38. Zheng, J. Z.; Zheng, X.; Zhao, Y.; Xie, Y.; Yam, C. Y.; Chen, G. H.; Jiang, Q.; Chwang, A. T. Maxwell’s Demon and
Smoluchowski’s Trap Door (vol 75, art no 041109, 2007) Phys. Rev. E 2007, 75, 041109
 DOI: 10.1103/PhysRevE.75.041109 [Crossref], [CAS], [Google Scholar]
 of
39. Skordos, P. A.; Zurek, W. H. Maxwell’s Demon, Recti ers, and the Second Law: Computer Simulation  
Smoluchowski’s Trapdoor Am. J. Phys. 1992, 60, 876– 882 DOI: 10.1119/1.17007 [Crossref], [Google Scholar]

40. Parrondo, J. M. R.; Espanol, P. Criticism of Feynman’s Analysis of the Ratchet as an Engine Am. J. Phys.
1996, 64, 1125– 1130 DOI: 10.1119/1.18393 [Crossref], [Google Scholar]

41. Sakaguchi, H. A Langevin Simulation for the Feynman Ratchet Model J. Phys. Soc. Jpn. 1998, 67, 709– 712
 DOI: 10.1143/JPSJ.67.709 [Crossref], [CAS], [Google Scholar]

42. Hondou, T.; Takagi, F. Irreversible Operation in a Stalled State of Feynman’s Ratchet J. Phys. Soc. Jpn. 1998,
67, 2974– 2976 DOI: 10.1143/JPSJ.67.2974 [Crossref], [CAS], [Google Scholar]

43. Magnasco, M. O.; Stolovitzky, G. Feynman’s Ratchet and Pawl J. Stat. Phys. 1998, 93, 615– 632
 DOI: 10.1023/B:JOSS.0000033245.43421.14 [Crossref], [Google Scholar]

44. Jarzynski, C.; Mazonka, O. Feynman’s Ratchet and Pawl: An Exactly Solvable Model Phys. Rev. E: Stat. Phys.,
Plasmas, Fluids, Relat. Interdiscip. Top. 1999, 59, 6448– 6459 DOI: 10.1103/PhysRevE.59.6448 [Crossref],
[PubMed], [CAS], [Google Scholar]

45. Jahn, R.; Fasshauer, D. Molecular Machines Governing Exocytosis of Synaptic Vesicles Nature 2012, 490,
201– 207 DOI: 10.1038/nature11320 [Crossref], [PubMed], [CAS], [Google Scholar]

46. Piccolino, M. Biological Machines: From Mills to Molecules Nat. Rev. Mol. Cell Biol. 2000, 1, 149– 153
 DOI: 10.1038/35040097 [Crossref], [PubMed], [CAS], [Google Scholar]

47. Kinbara, K.; Aida, T. Toward Intelligent Molecular Machines: Directed Motions of Biological and Arti cial
Molecules and Assemblies Chem. Rev. 2005, 105, 1377– 1400 DOI: 10.1021/cr030071r [ACS Full Text ],
[CAS], [Google Scholar]

48. Pomorski, T.; Hrafnsdóttir, S.; Devaux, P. F.; Meer, G. V. Lipid Distribution and Transport Across Cellular
Membranes Semin. Cell Dev. Biol. 2001, 12, 139– 148 DOI: 10.1006/scdb.2000.0231 [Crossref], [PubMed],
[CAS], [Google Scholar]

49. Haydon, D. A.; Hladky, S. B. Ion Transport Across Thin Lipid Membranes: A Critical Discussion of
Mechanisms in Selected Systems Q. Rev. Biophys. 1972, 5, 187– 282 DOI: 10.1017/S0033583500000883
[Crossref], [PubMed], [CAS], [Google Scholar]

  
50. Whittam, R.; Wheeler, K. P. Transport Across Cell Membranes Annu. Rev. Physiol. 1970, 32, 21– 60
 DOI: 10.1146/annurev.ph.32.030170.000321 [Crossref], [PubMed], [CAS], [Google Scholar]

51. Gouaux, E.; MacKinnon, R. Principles of Selective Ion Transport in Channels and Pumps Science 2005, 310,
1461– 1465 DOI: 10.1126/science.1113666 [Crossref], [PubMed], [CAS], [Google Scholar]

52. Jorgensen, P. L.; Hakansson, K. O.; Karlish, S. J. D. Structure and Mechanism of Na,K-ATPase: Functional
Sites and Their Interactions Annu. Rev. Physiol. 2003, 65, 817– 849
 DOI: 10.1146/annurev.physiol.65.092101.142558 [Crossref], [PubMed], [CAS], [Google Scholar]

53. Hayashi, S.; Ueno, H.; Shaikh, A. R.; Umemura, M.; Kamiya, M.; Ito, Y.; Ikeguchi, M.; Komoriya, Y.; Iino, R.; Noji,
H. Molecular Mechanism of ATP Hydrolysis in F1-ATPase Revealed by Molecular Simulations and Single-
Molecule Observations J. Am. Chem. Soc. 2012, 134, 8447– 8454 DOI: 10.1021/ja211027m [ACS Full Text ],
[CAS], [Google Scholar]

54. Nakanishi-Matsui, M.; Sekiya, M.; Nakamoto, R. K.; Futai, M. The Mechanism of Rotating Proton Pumping
ATPases Biochim. Biophys. Acta, Bioenerg. 2010, 1797, 1343– 1352 DOI: 10.1016/j.bbabio.2010.02.014
[Crossref], [PubMed], [CAS], [Google Scholar]

55. von Ballmoos, C.; Wiedenmann, A.; Dimroth, P. Essentials for ATP Synthesis by F1F0 ATP Synthases Annu.
Rev. Biochem. 2009, 78, 649– 672 DOI: 10.1146/annurev.biochem.78.081307.104803 [Crossref], [PubMed],
[CAS], [Google Scholar]

56. Nyblom, M.; Poulsen, H.; Gourdon, P.; Reinhard, L.; Andersson, M.; Lindahl, E.; Fedosova, N.; Nissen, P. Crystal
Structure of Na+, K+-ATPase in the Na+-Bound State Science 2013, 342, 123– 127 DOI: 10.1126/science.1243352
[Crossref], [PubMed], [CAS], [Google Scholar]

57. Nakamoto, R. K.; Scanlon, J. A. B.; Al-Shawi, M. K. The Rotary Mechanism of the ATP Synthase Arch.
Biochem. Biophys. 2008, 476, 43– 50 DOI: 10.1016/j.abb.2008.05.004 [Crossref], [PubMed], [CAS], [Google
Scholar]

58. Toyoshima, C.; Inesi, G. Structural Basis of Ion Pumping by Ca2+-ATPase of the Sarcoplasmic Reticulum
Annu. Rev. Biochem. 2004, 73, 269– 292 DOI: 10.1146/annurev.biochem.73.011303.073700 [Crossref],
[PubMed], [CAS], [Google Scholar]
59. Kastritis, P. L.; Bonvin, A. M. J. J. On the Binding of Macromolecular Interactions: Daring to Ask Why Proteins
 
Interact J. R. Soc., Interface 2012, 10, 20120835 DOI: 10.1098/rsif.2012.0835 [Crossref], [PubMed], [Google 
Scholar]

60. Maloney, P. C.; Kashket, E. R.; Wilson, T. H. A Protonmotive Force Drives ATP Synthesis in Bacteria Proc. Natl.
Acad. Sci. U. S. A. 1974, 71, 3896– 3900 DOI: 10.1073/pnas.71.10.3896 [Crossref], [PubMed], [CAS], [Google
Scholar]

61. Møller, J. V.; Nissen, P.; Sørensen, T. L. M.; Maire, M. l. Transport Mechanism of the Sarcoplasmic Reticulum
Ca2+-ATPase Pump Curr. Opin. Struct. Biol. 2005, 15, 387– 393 DOI: 10.1016/j.sbi.2005.06.005 [Crossref],
[PubMed], [CAS], [Google Scholar]

62. Birge, R. R. Nature of the Primary Photochemical Events in Rhodopsin and Bacteriorhodopsin Biochim.
Biophys. Acta, Bioenerg. 1990, 1016, 293– 327 DOI: 10.1016/0005-2728(90)90163-X [Crossref], [PubMed],
[CAS], [Google Scholar]

63. Neutze, R.; Pebay-Peyroula, E.; Edman, K.; Royant, A.; Navarro, J.; Landau, E. M. Bacteriorhodopsin: A High-
resolution Structural View of Vectorial Proton Transport Biochim. Biophys. Acta, Biomembr. 2002, 1565, 144–
167 DOI: 10.1016/S0005-2736(02)00566-7 [Crossref], [PubMed], [CAS], [Google Scholar]

64. Lanyi, J. K.; Luecke, H. Bacteriorhodopsin Curr. Opin. Struct. Biol. 2001, 11, 415– 419
 DOI: 10.1016/S0959-440X(00)00226-8 [Crossref], [PubMed], [CAS], [Google Scholar]

65. Shibata, M.; Yamashita, H.; Uchihashi, T.; Kandori, H.; Ando, T. High-speed Atomic Force Microscopy Shows
Dynamic Molecular Processes in Photoactivated Bacteriorhodopsin Nat. Nanotechnol. 2010, 5, 208– 212
 DOI: 10.1038/nnano.2010.7 [Crossref], [PubMed], [CAS], [Google Scholar]

66. Hirai, T.; Subramaniam, S. Protein Conformational Changes in the Bacteriorhodopsin Photocycle:
Comparison of Findings from Electron and X-Ray Crystallographic Analyses PLoS One 2009, 4, e5769
 DOI: 10.1371/journal.pone.0005769 [Crossref], [PubMed], [CAS], [Google Scholar]

67. Haupts, U.; Tittor, J.; Oesterhelt, D. Closing in on Bactereorhodopsin: Progress in Understanding the Molecule
Annu. Rev. Biophys. Biomol. Struct. 1999, 28, 367– 399 DOI: 10.1146/annurev.biophys.28.1.367 [Crossref],
[PubMed], [CAS], [Google Scholar]

68. Yonath, A. Hibernating Bears, Antibiotics, and the Evolving Ribosome (Nobel Lecture) Angew. Chem., Int. Ed.
2010, 49, 4340– 4354 DOI: 10.1002/anie.201001297 [Crossref], [CAS], [Google Scholar]
69. Ramakrishnan, V. Unraveling the Structure of the Ribosome (Nobel Lecture) Angew. Chem., Int.  
Ed. 2010, 49,
4355– 4380 DOI: 10.1002/anie.201001436 [Crossref], [PubMed], [CAS], [Google Scholar]

70. Steitz, T. A. From the Structure and Function of the Ribosome to New Antibiotics (Nobel Lecture) Angew.
Chem., Int. Ed. 2010, 49, 4381– 4398 DOI: 10.1002/anie.201000708 [Crossref], [PubMed], [CAS], [Google Scholar]

71. Bruck, I.; O’Donnell, M. The Ring-type Polymerase Sliding Clamp Family Genome Biol. 2001, 2,
reviews3001.1– reviews3001.3 DOI: 10.1186/gb-2001-2-1-reviews3001 [Crossref], [Google Scholar]

72. Astumian, R. D. Comment: Detailed Balance Revisited Phys. Chem. Chem. Phys. 2009, 11, 9592– 9594
 DOI: 10.1039/b911608g [Crossref], [PubMed], [CAS], [Google Scholar]

73. Astumian, R. D. Adiabatic Operation of a Molecular Machine Proc. Natl. Acad. Sci. U. S. A. 2007, 104,
19715– 19718 DOI: 10.1073/pnas.0708040104 [Crossref], [CAS], [Google Scholar]

74. Mattia, E.; Otto, S. Supramolecular Systems Chemistry Nat. Nanotechnol. 2015, 10, 111– 119
 DOI: 10.1038/nnano.2014.337 [Crossref], [PubMed], [CAS], [Google Scholar]

75. Chatterjee, M. N.; Kay, E. R.; Leigh, D. A. Beyond Switches: Ratcheting a Particle Energetically Uphill with a
Compartmentalized Molecular Machine J. Am. Chem. Soc. 2006, 128, 4058– 4073 DOI: 10.1021/ja057664z
[ACS Full Text ], [CAS], [Google Scholar]

76. Astumian, R. D. Stochastic Conformational Pumping: A Mechanism for Free-energy Transduction by


Molecules Annu. Rev. Biophys. 2011, 40, 289– 313 DOI: 10.1146/annurev-biophys-042910-155355 [Crossref],
[PubMed], [CAS], [Google Scholar]

77. Astumian, R. D. Microscopic Reversibility as the Organizing Principle of Molecular Machines Nat.
Nanotechnol. 2012, 7, 684– 688 DOI: 10.1038/nnano.2012.188 [Crossref], [PubMed], [CAS], [Google Scholar]

78. Simon, M. S.; Peskin, C. S.; Oster, G. F. What Drives the Translocation of Proteins? Proc. Natl. Acad. Sci. U. S.
A. 1992, 89, 3770– 3774 DOI: 10.1073/pnas.89.9.3770 [Crossref], [PubMed], [CAS], [Google Scholar]

79. Astumian, R. D. Thermodynamics and Kinetics of a Brownian Motor Science 1997, 276, 917– 922
 DOI: 10.1126/science.276.5314.917 [Crossref], [PubMed], [CAS], [Google Scholar]
80. Bier, M. Brownian Ratchets in Physics and Biology Contemp. Phys. 1997, 38, 371– 379
 DOI: 10.1080/001075197182180 [Crossref], [CAS], [Google Scholar]   

81. Juelicher, F.; Ajdari, A.; Prost, J. Modeling Molecular Motors Rev. Mod. Phys. 1997, 69, 1269– 1281
 DOI: 10.1103/RevModPhys.69.1269 [Crossref], [Google Scholar]

82. Astumian, R. D. Making Molecules into Motors Sci. Am. 2001, 285, 56– 64
 DOI: 10.1038/scienti camerican0701-56 [Crossref], [PubMed], [CAS], [Google Scholar]

83. Lipowsky, R.; Jaster, N. Molecular Motor Cycles: From Ratchets to Networks J. Stat. Phys. 2003, 110, 1141–
1167 DOI: 10.1023/A:1022101011650 [Crossref], [Google Scholar]

84. Parrondo, J. M. R.; Dinís, L. Brownian Motion and Gambling: From Ratchets to Paradoxical Games Contemp.
Phys. 2004, 45, 147– 157 DOI: 10.1080/00107510310001644836 [Crossref], [Google Scholar]

85. Astumian, R. D. Design Principles for Brownian Molecular Machines: How to Swim in Molasses and Walk in
a Hurricane Phys. Chem. Chem. Phys. 2007, 9, 5067– 5083 DOI: 10.1039/b708995c [Crossref], [PubMed],
[CAS], [Google Scholar]

86. Parrondo, J. M. R.; de Cisneros, B. J. Energetics of Brownian Motors: A Review Appl. Phys. A: Mater. Sci.
Process. 2002, 75, 179– 191 DOI: 10.1007/s003390201332 [Crossref], [CAS], [Google Scholar]

87. Reimann, P.; Hänggi, P. Introduction to the Physics of Brownian Motors Appl. Phys. A: Mater. Sci. Process.
2002, 75, 169– 178 DOI: 10.1007/s003390201331 [Crossref], [CAS], [Google Scholar]

88. Landauer, R. Inadequacy of Entropy and Entropy Derivatives in Characterizing the Steady State Phys. Rev. A:
At., Mol., Opt. Phys. 1975, 12, 636– 638 DOI: 10.1103/PhysRevA.12.636 [Crossref], [Google Scholar]

89. Büttiker, M. Transport as a Consequence of State-dependent Diffusion Z. Phys. B: Condens. Matter 1987, 68,
161– 167 DOI: 10.1007/BF01304221 [Crossref], [Google Scholar]

90. Landauer, R. Motion out of Noisy States J. Stat. Phys. 1988, 53, 233– 248 DOI: 10.1007/BF01011555
[Crossref], [Google Scholar]
91. Van Kampen, N. G. Relative Stability in Nonuniform Temperature IBM J. Res. Dev. 1988, 32, 107– 111
 DOI: 10.1147/rd.321.0107 [Crossref], [Google Scholar]   

92. Sinha, K.; Moss, F. Analog Simulation of a Simple System with State-dependent Diffusion J. Stat. Phys.
1989, 54, 1411– 1423 DOI: 10.1007/BF01044724 [Crossref], [Google Scholar]

93. Rousselet, J.; Salome, L.; Ajdari, A.; Prostt, J. Directional Motion of Brownian Particles Induced by a Periodic
Asymmetric Potential Nature 1994, 370, 446– 447 DOI: 10.1038/370446a0 [Crossref], [PubMed], [CAS], [Google
Scholar]

94. Faucheux, L. P.; Bourdieu, L. S.; Kaplan, P. D.; Libchaber, A. J. Optical Thermal Ratchet Phys. Rev. Lett. 1995,
74, 1504– 1507 DOI: 10.1103/PhysRevLett.74.1504 [Crossref], [PubMed], [CAS], [Google Scholar]

95. Faucheux, L. P.; Libchaber, A. Selection of Brownian Particles J. Chem. Soc., Faraday Trans. 1995, 91, 3163–
3166 DOI: 10.1039/ft9959103163 [Crossref], [CAS], [Google Scholar]

96. Faucheux, L. P.; Stolovitzky, G.; Libchaber, A. Periodic Forcing of a Brownian Particle Phys. Rev. E: Stat. Phys.,
Plasmas, Fluids, Relat. Interdiscip. Top. 1995, 51, 5239– 5250 DOI: 10.1103/PhysRevE.51.5239 [Crossref],
[PubMed], [CAS], [Google Scholar]

97. Astumian, R. D.; Bier, M. Mechanochemical Coupling of the Motion of Molecular Motors to ATP Hydrolysis
Biophys. J. 1996, 70, 637– 653 DOI: 10.1016/S0006-3495(96)79605-4 [Crossref], [PubMed], [CAS], [Google
Scholar]

98. Gorre, L.; Ioannidis, E.; Silberzan, P. Recti ed Motion of a Mercury Drop in an Asymmetric Structure
Europhys. Lett. 1996, 33, 267– 272 DOI: 10.1209/epl/i1996-00331-2 [Crossref], [CAS], [Google Scholar]

99. Zhou, H.-X.; Chen, Y.-D. Chemically Driven Motility of Brownian Particles Phys. Rev. Lett. 1996, 77, 194– 197
 DOI: 10.1103/PhysRevLett.77.194 [Crossref], [PubMed], [CAS], [Google Scholar]

100. Gorre-Talini, L.; Jeanjean, S.; Silberzan, P. Sorting of Brownian Particles by the Pulsed Application of an
Asymmetric Potential Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 1997, 56, 2025– 2034
 DOI: 10.1103/PhysRevE.56.2025 [Crossref], [CAS], [Google Scholar]

101. Gorre-Talini, L.; Silberzan, P. Force-Free Motion of a Mercury Drop Alternatively Submitted to Shifted
Asymmetric Potentials J. Phys. I (France) 1997, 7, 1475– 1485 [Crossref], [Google Scholar]
102. Astumian, R. D.; Derényi, I. Fluctuation Driven Transport and Models of Molecular Motors and Eur.
Pumps
Biophys. J. 1998, 27, 474– 489 DOI: 10.1007/s002490050158 [Crossref], [PubMed], [CAS], [Google Scholar]

103. Gorre-Talini, L.; Spatz, J. P.; Silberzan, P. Dielectrophoretic Ratchets Chaos 1998, 8, 650– 656
 DOI: 10.1063/1.166347 [Crossref], [PubMed], [CAS], [Google Scholar]

104. Linke, H.; Sheng, W.; Löfgren, A.; Hongqi, X.; Omling, P.; Lindelof, P. E. A Quantum Dot Ratchet: Experiment
and Theory Europhys. Lett. 1998, 44, 341– 347 DOI: 10.1209/epl/i1998-00562-1 [Crossref], [CAS], [Google
Scholar]

105. Lorke, A.; Wimmer, S.; Jager, B.; Kotthaus, J. P.; Wegscheider, W.; Bichler, M. Far-infrared and Transport
Properties of Antidot Arrays with Broken Symmetry Phys. B 1998, 249–251, 312– 316
 DOI: 10.1016/S0921-4526(98)00121-5 [Crossref], [CAS], [Google Scholar]

106. Bekele, M.; Rajesh, S.; Ananthakrishna, G.; Kumar, N. Effect of Landauer’s Blow Torch on the Equilibration
Rate in a Bistable Potential Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 1999, 59, 143– 149
 DOI: 10.1103/PhysRevE.59.143 [Crossref], [CAS], [Google Scholar]

107. Derényi, I.; Bier, M.; Astumian, R. D. Generalized E ciency and its Application to Microscopic Engines Phys.
Rev. Lett. 1999, 83, 903– 906 DOI: 10.1103/PhysRevLett.83.903 [Crossref], [CAS], [Google Scholar]

108. Linke, H.; Humphrey, T. E.; Löfgren, A.; Sushkov, A. O.; Newbury, R.; Taylor, R. P.; Omling, P. Experimental
Tunneling Ratchets Science 1999, 286, 2314– 2317 DOI: 10.1126/science.286.5448.2314 [Crossref], [PubMed],
[CAS], [Google Scholar]

109. Mennerat-Robilliard, C.; Lucas, D.; Guibal, S.; Tabosa, J.; Jurczak, C.; Courtois, J. Y.; Grynberg, G. Ratchet for
Cold Rubidium Atoms: The Asymmetric Optical Lattice Phys. Rev. Lett. 1999, 82, 851– 854
 DOI: 10.1103/PhysRevLett.82.851 [Crossref], [CAS], [Google Scholar]

110. Parmeggiani, A.; Jülicher, F.; Ajdari, A.; Prost, J. Energy Transduction of Isothermal Ratchets: Generic
Aspects and Speci c Examples Close to and Far from Equilibrium Phys. Rev. E: Stat. Phys., Plasmas, Fluids,
Relat. Interdiscip. Top. 1999, 60, 2127– 2140 DOI: 10.1103/PhysRevE.60.2127 [Crossref], [PubMed],
[CAS], [Google Scholar]

111. Switkes, M.; Marcus, C. M.; Campman, K.; Gossard, A. C. An Adiabatic Quantum Electron Pump Science
1999, 283, 1905– 1908 DOI: 10.1126/science.283.5409.1905 [Crossref], [PubMed], [CAS], [Google Scholar]
  
112. van Oudenaarden, A.; Boxer, S. G. Brownian Ratchets: Molecular Separations in Lipid Bilayers Supported on
Patterned Arrays Science 1999, 285, 1046– 1048 DOI: 10.1126/science.285.5430.1046 [Crossref], [PubMed],
[CAS], [Google Scholar]

113. Linke, H.; Sheng, W. D.; Svensson, A.; Löfgren, A.; Christensson, L.; Xu, H. Q.; Omling, P.; Lindelof, P. E.
Asymmetric Nonlinear Conductance of Quantum Dots with Broken Inversion Symmetry Phys. Rev. B: Condens.
Matter Mater. Phys. 2000, 61, 15914– 15926 DOI: 10.1103/PhysRevB.61.15914 [Crossref], [CAS], [Google
Scholar]

114. Lipowsky, R. Molecular Motors and Stochastic Models. In Stochastic Processes in Physics, Chemistry, and
Biology; Freund, J.; Pöschel, T., Eds.; Springer: Berlin Heidelberg, 2000; Vol. 557, pp 21– 31. [Crossref], [Google
Scholar]

115. Mahadevan, L.; Matsudaira, P. Motility Powered by Supramolecular Springs and Ratchets Science 2000,
288, 95– 99 DOI: 10.1126/science.288.5463.95 [Crossref], [PubMed], [CAS], [Google Scholar]

116. Bustamante, C.; Keller, D.; Oster, G. The Physics of Molecular Motors Acc. Chem. Res. 2001, 34, 412– 420
 DOI: 10.1021/ar0001719 [ACS Full Text ], [CAS], [Google Scholar]

117. Höhberger, E. M.; Lorke, A.; Wegscheider, W.; Bichler, M. Adiabatic Pumping of Two-dimensional Electrons
in a Ratchet-type Lateral Superlattice Appl. Phys. Lett. 2001, 78, 2905– 2907 DOI: 10.1063/1.1355672 [Crossref],
[CAS], [Google Scholar]

118. Astumian, R. D. Protein Conformational Fluctuations and Free-energy Transduction Appl. Phys. A: Mater.
Sci. Process. 2002, 75, 193– 206 DOI: 10.1007/s003390201406 [Crossref], [CAS], [Google Scholar]

119. Linke, H.; Humphrey, T. E.; Lindelof, P. E.; Löfgren, A.; Newbury, R.; Omling, P.; Sushkov, A. O.; Taylor, R. P.; Xu,
H. Quantum Ratchets and Quantum Heat Pumps Appl. Phys. A: Mater. Sci. Process. 2002, 75, 237– 246
 DOI: 10.1007/s003390201335 [Crossref], [CAS], [Google Scholar]

120. Marquet, C.; Buguin, A.; Talini, L.; Silberzan, P. Recti ed Motion of Colloids in Asymmetrically Structured
Channels Phys. Rev. Lett. 2002, 88, 168301 DOI: 10.1103/PhysRevLett.88.168301 [Crossref], [PubMed],
[CAS], [Google Scholar]
121. Huang, L. R.; Cox, E. C.; Austin, R. H.; Sturm, J. C. Tilted Brownian Ratchet for DNA Analysis Anal. Chem.
2003, 75, 6963– 6967 DOI: 10.1021/ac0348524 [ACS Full Text ], [CAS], [Google Scholar]   

122. Matthias, S.; Muller, F. Asymmetric Pores in a Silicon Membrane Acting as Massively Parallel Brownian
Ratchets Nature 2003, 424, 53– 57 DOI: 10.1038/nature01736 [Crossref], [PubMed], [CAS], [Google Scholar]

123. Mogilner, A.; Oster, G. Polymer Motors: Pushing out the Front and Pulling up the Back Curr. Biol. 2003, 13,
R721– R733 DOI: 10.1016/j.cub.2003.08.050 [Crossref], [PubMed], [CAS], [Google Scholar]

124. Oster, G.; Wang, H. Rotary Protein Motors Trends Cell Biol. 2003, 13, 114– 121
 DOI: 10.1016/S0962-8924(03)00004-7 [Crossref], [PubMed], [CAS], [Google Scholar]

125. Kurzyński, M.; Chełminiak, P. Stochastic Action of Actomyosin Motor Phys. A 2004, 336, 123– 132
 DOI: 10.1016/j.physa.2004.01.017 [Crossref], [CAS], [Google Scholar]

126. de Souza Silva, C. C.; Van de Vondel, J.; Morelle, M.; Moshchalkov, V. V. Controlled Multiple Reversals of a
Ratchet Effect Nature 2006, 440, 651– 654 DOI: 10.1038/nature04595 [Crossref], [PubMed], [CAS], [Google
Scholar]

127. Linke, H.; Downton, M. T.; Zuckermann, M. J. Performance Characteristics of Brownian Motors Chaos
2005, 15, 26111 DOI: 10.1063/1.1871432 [Crossref], [PubMed], [CAS], [Google Scholar]

128. Reimann, P. Brownian Motors: Noisy Transport Far from Equilibrium Phys. Rep. 2002, 361, 57– 265
 DOI: 10.1016/S0370-1573(01)00081-3 [Crossref], [CAS], [Google Scholar]

129. Gabryś, B. J.; Pesz, K.; Bartkiewicz, S. J. Brownian Motion, Molecular Motors and Ratchets Phys. A 2004,
336, 112– 122 DOI: 10.1016/j.physa.2004.01.016 [Crossref], [Google Scholar]

130. Mislow, K. Stereochemical Consequences of Correlated Rotation in Molecular Propellers Acc. Chem. Res.
1976, 9, 26– 33 DOI: 10.1021/ar50097a005 [ACS Full Text ], [CAS], [Google Scholar]

131. Rappoport, Z.; Biali, S. E. Sterically Crowded Stable Simple Enols Acc. Chem. Res. 1988, 21, 442– 449
 DOI: 10.1021/ar00156a002 [ACS Full Text ], [CAS], [Google Scholar]
132. Rappoport, Z.; Biali, S. E. Threshold Rotational Mechanisms and Enantiomerization Barriers of Polyarylvinyl
Propellers Acc. Chem. Res. 1997, 30, 307– 314 DOI: 10.1021/ar960216z [ACS Full Text  
], [CAS], [Google 
Scholar]

133. Wolf, C. Dynamic Stereochemistry of Chiral Compounds; RSC Publishing: UK, 2008; Chapter 8, pp 399–
443. [Google Scholar]

134. Gust, D.; Mislow, K. Analysis of Isomerization in Compounds Displaying Restricted Rotation of Aryl Groups
J. Am. Chem. Soc. 1973, 95, 1535– 1547 DOI: 10.1021/ja00786a031 [ACS Full Text ], [CAS], [Google Scholar]

135. Mislow, K.; Gust, D.; Finocchiaro, P.; Boettcher, R. Stereochemical Correspondence Among Molecular
Propellers. Stereochemistry I; Springer: Berlin, Heidelberg, 1974; Vol. 47, pp 1– 28. [Crossref], [Google Scholar]

136. Katoono, R.; Kawai, H.; Fujiwara, K.; Suzuki, T. Dynamic Molecular Propeller: Supramolecular Chirality
Sensing by Enhanced Chiroptical Response through the Transmission of Point Chirality to Mobile Helicity J. Am.
Chem. Soc. 2009, 131, 16896– 16904 DOI: 10.1021/ja906810b [ACS Full Text ], [CAS], [Google Scholar]

137. Driesschaert, B.; Robiette, R.; Le Duff, C. S.; Collard, L.; Robeyns, K.; Gallez, B.; Marchand-Brynaert, J.
Con gurationally Stable Tris(tetrathioaryl)methyl Molecular Propellers Eur. J. Org. Chem. 2012, 6517– 6525
 DOI: 10.1002/ejoc.201200801 [Crossref], [CAS], [Google Scholar]

138. Katoono, R.; Kawai, H.; Ohkita, M.; Fujiwara, K.; Suzuki, T. A C(3)-symmetric Chiroptical Molecular Propeller
Based on Hexakis(phenylethynyl)benzene with a Threefold Terephthalamide: Stereospeci c Propeller Generation
Through the Cooperative Transmission of Point Chiralities on the Host and Guest upon Complexation Chem.
Commun. 2013, 49, 10352– 10354 DOI: 10.1039/c3cc43571g [Crossref], [PubMed], [CAS], [Google Scholar]

139. Finocchiaro, P.; Gust, D.; Mislow, K. Structure and Dynamic Stereochemistry of Trimesitylmethane. I.
Synthesis and Nuclear Magnetic Resonance Studies J. Am. Chem. Soc. 1974, 96, 2165– 2167
 DOI: 10.1021/ja00814a028 [ACS Full Text ], [CAS], [Google Scholar]

140. O̅ki, M. Unusually High Barriers to Rotation Involving the Tetrahedral Carbon Atom Angew. Chem., Int. Ed.
Engl. 1976, 15, 87– 93 DOI: 10.1002/anie.197600871 [Crossref], [Google Scholar]

141. Yamamoto, G.; O̅ki, M. Dual Mechanisms of the Aryl Group Rotation in 9-(3,5-dimethylbenzyl) Triptycene
Derivatives Chem. Lett. 1979, 1251– 1254 DOI: 10.1246/cl.1979.1251 [Crossref], [CAS], [Google Scholar]
142. Yamamoto, G.; O̅ki, M. Two Consecutive Gear Motions in Conformational Interconversion in 9-(2-
 
methylbenzyl)triptycene Derivatives Chem. Lett. 1979, 1255– 1258 DOI: 10.1246/cl.1979.1255 [Crossref], 
[CAS], [Google Scholar]

143. Yamaoto, G.; O̅ki, M. Restricted Rotation Involving the Tetrahedral Carbon. XXXV. Stereodynamics of 9-(3,5-
Dimethylbenzyl)triptycene Derivatives Bull. Chem. Soc. Jpn. 1981, 54, 473– 480 DOI: 10.1246/bcsj.54.473
[Crossref], [Google Scholar]

144. Yamamoto, G.; O̅ki, M. Restricted Rotation Involving the Tetrahedral Carbon. XXXVI. Stereodynamics of 9-
(2-MethylbenzyI)triptycene Derivatives Bull. Chem. Soc. Jpn. 1981, 54, 481– 487 DOI: 10.1246/bcsj.54.481
[Crossref], [CAS], [Google Scholar]

145. Yamamoto, G.; O̅ki, M. Restricted Rotation Involving the Tetrahedral Carbon. Part 46. Correlated Rotation
in 9-(2,4,6-trimethylbenzyl)triptycenes. Direct and Roundabout Enantiomerization-diastereomerization Processes
J. Org. Chem. 1983, 48, 1233– 1236 DOI: 10.1021/jo00156a017 [ACS Full Text ], [CAS], [Google Scholar]

146. Yamamoto, G. Molecular Gears with Two-toothed and Three-toothed Wheels J. Mol. Struct. 1985, 126,
413– 420 DOI: 10.1016/0022-2860(85)80130-7 [Crossref], [CAS], [Google Scholar]

147. Yamamoto, G.; O̅ki, M. Restricted Rotation Involving the Tetrahedral Carbon. LVII. Stereodynamics of 9-(2-
Alkylphenoxy)-1,4-dimethyltriptycenes Bull. Chem. Soc. Jpn. 1985, 58, 1953– 1961 DOI: 10.1246/bcsj.58.1953
[Crossref], [CAS], [Google Scholar]

148. Yamamoto, G.; O̅ki, M. Restricted Rotation Involving the Tetrahedral Carbon. LIX. Stereodynamics of Singly
peri-Substituted 9-(3,5-Dimethylphenoxy)triptycene Derivatives Bull. Chem. Soc. Jpn. 1986, 59, 3597– 3603
 DOI: 10.1246/bcsj.59.3597 [Crossref], [CAS], [Google Scholar]

149. Yamamoto, G. Detailed Dynamic NMR Study of a Molecular Gear, 1-Methoxy-9-(3,5-


dimethylbenzyl)triptycene Bull. Chem. Soc. Jpn. 1989, 62, 4058– 4060 DOI: 10.1246/bcsj.62.4058 [Crossref],
[CAS], [Google Scholar]

150. Hiizu Iwamura, K. M. Stereochemical Consequences of Dynamic Gearing Acc. Chem. Res. 1988, 21, 175–
182 DOI: 10.1021/ar00148a007 [ACS Full Text ], [Google Scholar]

151. Kawada, Y.; Iwamura, H. Unconventional Synthesis and Conformational Flexibility of Bis(1-triptycyl) ether J.
Org. Chem. 1980, 45, 2548– 2550 DOI: 10.1021/jo01300a069 [ACS Full Text ], [Google Scholar]
152. Kawada, Y.; Iwamura, H. Bis(4-chloro-1-triptycyl) ether. Separation of a Pair of Phase Isomers of Labeled
Bevel Gears J. Am. Chem. Soc. 1981, 103, 958– 960 DOI: 10.1021/ja00394a049 [ACS Full Text ],  
[CAS], [Google
Scholar]

153. Kawada, Y.; Iwamura, H. Bis(4-chloro-1-triptycyl) ether. Separation of a Pair of Phase Isomers of Labeled
Bevel Gears Tetrahedron Lett. 1981, 22, 1533– 1536 DOI: 10.1016/S0040-4039(01)90370-3 [Crossref],
[CAS], [Google Scholar]

154. Kawada, Y.; Iwamura, H. Correlated Rotation in Bis(9-triptycyl)methanes and Bis(9-triptycyl) Ethers.
Separation and Interconversion of the Phase Isomers of Labeled Bevel Gears J. Am. Chem. Soc. 1983, 105,
1449– 1459 DOI: 10.1021/ja00344a006 [ACS Full Text ], [CAS], [Google Scholar]

155. Kawada, Y.; Okamoto, Y.; Iwamura, H. Correlated Internal Rotation in Bis(2,6-dichloro-9-triptycyl)methane.
To what Extent can Phase Isomers be Separated and Identi ed? Tetrahedron Lett. 1983, 24, 5359– 5362
 DOI: 10.1016/S0040-4039(00)87868-5 [Crossref], [CAS], [Google Scholar]

156. Iwamura, H.; Ito, T.; Ito, H.; Toriumi, K.; Kawada, Y.; Osawa, E.; Fujiyoshi, T.; Jaime, C. Crystal and Molecular
Structure of Bis(9-triptycyl) Ether J. Am. Chem. Soc. 1984, 106, 4712– 4717 DOI: 10.1021/ja00329a013 [ACS
Full Text ], [CAS], [Google Scholar]

157. Iwamura, H. Molecular Design of Correlated Internal Rotation J. Mol. Struct. 1985, 126, 401– 412
 DOI: 10.1016/0022-2860(85)80129-0 [Crossref], [CAS], [Google Scholar]

158. Koga, N.; Iwamura, H. Barrier to Coupled Internal Rotation in Bis(9-triptycyl) Ether. Kinetics of
Intramolecular Exciplex Formation in Racemic 2,3-Benzo-9-triptycyl 2-(N,N-dimethylaminomethyl)-9-triptycyl
ether J. Am. Chem. Soc. 1985, 107, 1426– 1427 DOI: 10.1021/ja00291a061 [ACS Full Text ], [CAS], [Google
Scholar]

159. Koga, N.; Iwamura, H. A Kinetic Study of Coupled Internal Rotation in Racemic 2,3-Benzo-9-triptycyl 2-
(dimethylaminomethyl)-9-triptycyl ether by Means of Exciplex Fluorescence Dynamics Chem. Lett. 1986, 247–
250 DOI: 10.1246/cl.1986.247 [Crossref], [CAS], [Google Scholar]

160. Kawada, Y.; Yamazaki, H.; Koga, G.; Murata, S.; Iwamura, H. Bis(9-triptycyl)amines, A Missing Link Between
the Corresponding Methanes and Ethers. An Unconventional Synthesis and In uence of Nitrogen Con gurational
Inversion on the Coupled Disrotatory Trajectory J. Org. Chem. 1986, 51, 1472– 1477 DOI: 10.1021/jo00359a016
[ACS Full Text ], [CAS], [Google Scholar]
161. Kawada, Y.; Ishikawa, J.; Yamazaki, H.; Koga, G.; Murata, S.; Iwamura, H. Bis(9-triptycyl)amines, A Missing
Link Between the Corresponding Methanes and Ethers. An Unconventional Synthesis and In uence  
of Nitrogen
Con gurational Inversion on the Coupled Disrotatory Trajectory Tetrahedron Lett. 1987, 28, 445– 448
 DOI: 10.1016/S0040-4039(00)95752-6 [Crossref], [CAS], [Google Scholar]

162. Hounshell, W. D.; Johnson, C. A.; Guenzi, A.; Cozzi, F.; Mislow, K. Sterochemical Consequences of Dynamic
Gearing in Substituted bis (9-triptycyl) Methanes and Related Molecules Proc. Natl. Acad. Sci. U. S. A. 1980, 77,
6961– 6964 DOI: 10.1073/pnas.77.12.6961 [Crossref], [PubMed], [CAS], [Google Scholar]

163. Cozzi, F.; Guenzi, A.; Johnson, C. A.; Mislow, K. Stereoisomerism and Correlated Rotation in Molecular Gear
Systems. Residual Diastereomers of Bis(2,3-dimethyl-9-triptycyl)methane J. Am. Chem. Soc. 1981, 103, 957–
958 DOI: 10.1021/ja00394a048 [ACS Full Text ], [CAS], [Google Scholar]

164. Johnson, C. A.; Guenzi, A.; Mislow, K. Restricted Gearing and Residual Stereoisomerism in Bis(1,4-dimethyl-
9-triptycyl)methane J. Am. Chem. Soc. 1981, 103, 6240– 6242 DOI: 10.1021/ja00410a054 [ACS Full Text ],
[CAS], [Google Scholar]

165. Johnson, C. A.; Guenzi, A.; Nachbar, R. B., Jr.; Blount, J. F.; Wennerstroem, O.; Mislow, K. Crystal and
Molecular Structure of Bis(9-triptycyl) ketone and Bis(9-triptycyl)methane J. Am. Chem. Soc. 1982, 104, 5163–
5168 DOI: 10.1021/ja00383a028 [ACS Full Text ], [CAS], [Google Scholar]

166. Buergi, H. B.; Hounshell, W. D.; Nachbar, J. R. B.; Mislow, K. Conformational Dynamics of Propane, di-tert-
Butylmethane, and Bis(9-triptycyl)methane. An Analysis of the Symmetry of Two Threefold Rotors on a Rigid
Frame in Terms of Nonrigid Molecular Structure and Energy Hypersurfaces J. Am. Chem. Soc. 1983, 105, 1427–
1438 DOI: 10.1021/ja00344a004 [ACS Full Text ], [CAS], [Google Scholar]

167. Guenzi, A.; Johnson, C. A.; Cozzi, F.; Mislow, K. Dynamic Gearing and Residual Stereoisomerism in Labeled
Bis(9-triptycyl)methane and Related Molecules. Synthesis and Stereochemistry of Bis(2,3-dimethyl-9-
triptycyl)methane, Bis(2,3-dimethyl-9-triptycyl)carbinol, and Bis(1,4-dimethyl-9-triptycyl) Methane J. Am. Chem.
Soc. 1983, 105, 1438– 1448 DOI: 10.1021/ja00344a005 [ACS Full Text ], [CAS], [Google Scholar]

168. Koga, N.; Kawada, Y.; Iwamura, H. Recognition of the Phase Relationship between Remote Substituents in
9,10-Bis(3-chloro-9-triptycycloxy)triptycene Molecules Undergoing Rapid Internal Rotation Cooperatively J. Am.
Chem. Soc. 1983, 105, 5498– 5499 DOI: 10.1021/ja00354a063 [ACS Full Text ], [CAS], [Google Scholar]

169. Yamamoto, G. Molecular Gears with Two-toothed and Three-thoothed Wheels J. Mol. Struct. 1985, 126,
413– 420 DOI: 10.1016/0022-2860(85)80130-7 [Crossref], [CAS], [Google Scholar]
170. Yamamoto, G. Dynamic Stereochemistry of Molecular Geras, 9-Benzyltriptycene and 9-Phenoxytriptycene,
Studied by 13C Dynamic NMR Spectroscopy and Molecular Mechanics Calculations Tetrahedron 1990,
 46, 
2761– 2772 DOI: 10.1016/S0040-4020(01)88370-8 [Crossref], [CAS], [Google Scholar]

171. Chance, J. M.; Geiger, J. H.; Okamoto, Y.; Aburatani, R.; Mislow, K. Stereochemical Consequences of a
Parity Restriction on Dynamic Gearing in Tris(9-triptycyl)germanium Chloride and Tris(9-triptycyl)cyclopropenium
Perchlorate J. Am. Chem. Soc. 1990, 112, 3540– 3547 DOI: 10.1021/ja00165a044 [ACS Full Text ],
[CAS], [Google Scholar]

172. Kawada, Y.; Sakai, H.; Oguri, M.; Koga, G. Preparation of an Dynamic Gearing in cis-1,2-Bis(9-
triptycyl)ethylene Tetrahedron Lett. 1994, 35, 139– 142 DOI: 10.1016/0040-4039(94)88184-7 [Crossref],
[CAS], [Google Scholar]

173. Nikitin, K.; Müller-Bunz, H.; Ortin, Y.; Risse, W.; McGlinchey, M. J. Twin Triptycyl Spinning Tops: A Simple
Case of Molecular Gearing with Dynamic C2 Symmetry Eur. J. Org. Chem. 2008, 2008, 3079– 3084
 DOI: 10.1002/ejoc.200800202 [Crossref], [Google Scholar]

174. Frantz, D. K.; Linden, A.; Baldridge, K. K.; Siegel, J. S. Molecular Spur Gears Comprising Triptycene Rotators
and Bibenzimidazole-based Stators J. Am. Chem. Soc. 2012, 134, 1528– 1535 DOI: 10.1021/ja2063346 [ACS
Full Text ], [CAS], [Google Scholar]

175. Stevens, A. M.; Richards, C. J. A Metallocene Molecular Gear Tetrahedron Lett. 1997, 38, 7805– 7808
 DOI: 10.1016/S0040-4039(97)10042-9 [Crossref], [CAS], [Google Scholar]

176. Brydges, S.; Harrington, L. E.; McGlinchey, M. J. Sterically Hindered Organometallics: Multi-n-rotor (n = 5, 6
and 7) Molecular Propellers and the Search for Correlated Rotations Coord. Chem. Rev. 2002, 233, 75– 105
 DOI: 10.1016/S0010-8545(02)00098-X [Crossref], [CAS], [Google Scholar]

177. Kuttenberger, M.; Frieser, M.; Hofweber, M.; Mannschreck, A. Axially Chiral Thioamides of Acrylic Acid:
Correlated and Uncorrelated Internal Rotations Tetrahedron: Asymmetry 1998, 9, 3629– 3645
 DOI: 10.1016/S0957-4166(98)00373-5 [Crossref], [CAS], [Google Scholar]

178. Clayden, J.; Pink, J. H. Concerted Rotation in a Tertiary Aromatic Amide: Towards a Simple Molecular Gear
Angew. Chem., Int. Ed. 1998, 37, 1937– 1939
 DOI: 10.1002/(SICI)1521-3773(19980803)37:13/14<1937::AID-ANIE1937>3.0.CO;2-4 [Crossref], [CAS], [Google
Scholar]
179. Johnston, E. R.; Fortt, R.; Barborak, J. C. Correlated Rotation in a Conformationally Restricted Amide Magn.
 
Reson. Chem. 2000, 38, 932– 936 DOI: 10.1002/1097-458X(200011)38:11<932::AID-MRC752>3.0.CO;2-6 
[Crossref], [CAS], [Google Scholar]

180. Bragg, R. A.; Clayden, J. Using Symmetry to Monitor Geared Bond Rotation in Aromatic Amides by
Dynamic NMR Org. Lett. 2000, 2, 3351– 3354 DOI: 10.1021/ol0064462 [ACS Full Text ], [CAS], [Google Scholar]

181. Bragg, R. A.; Clayden, J.; Morris, G. A.; Pink, J. H. Stereodynamics of Bond Rotation in Tertiary Aromatic
Amides Chem. - Eur. J. 2002, 8, 1279– 1289
 DOI: 10.1002/1521-3765(20020315)8:6<1279::AID-CHEM1279>3.0.CO;2-7 [Crossref], [PubMed], [CAS], [Google
Scholar]

182. Hellwinkel, D.; Melan, M.; Degel, C. R. Die Stereochemie ortho-Substituierter Triphenylamin-derivate
Tetrahedron 1973, 29, 1895– 1907 DOI: 10.1016/S0040-4020(01)83217-8 [Crossref], [CAS], [Google Scholar]

183. Hummel, J. P.; Gust, D.; Mislow, K. Mechanism of Stereoisomerization in Triarylboranes J. Am. Chem. Soc.
1974, 96, 3679– 3681 DOI: 10.1021/ja00818a068 [ACS Full Text ], [CAS], [Google Scholar]

184. Wille, E. E.; Stephenson, D. S.; Capriel, P.; Binsch, G. Iterative Analysis of Exchange-Broadened NMR Band
Shapes. The Mechanism of Correlated Rotations in Triaryl Derivatives of Phosphorus and Arsenic J. Am. Chem.
Soc. 1982, 104, 405– 415 DOI: 10.1021/ja00366a006 [ACS Full Text ], [CAS], [Google Scholar]

185. Clegg, W.; Lockhart, J. C.; McDonnell, M. B. Comparison of the Steric Barriers in Three- and Two-bladed
Propeller Crowns J. Chem. Soc., Perkin Trans. 1 1985, 1019– 1023 DOI: 10.1039/p19850001019 [Crossref],
[CAS], [Google Scholar]

186. Berg, U.; Liljefors, T.; Roussel, C.; Sandstroem, J. Steric Interplay Between Alkyl Groups Bonded to Planar
Frameworks Acc. Chem. Res. 1985, 18, 80– 86 DOI: 10.1021/ar00111a003 [ACS Full Text ], [CAS], [Google
Scholar]

187. Biali, S. E.; Rappoport, Z. Stable Simple Enols. 8.′ Synthesis and Keto Enol Equilibria of the Elusive 2,2-
Dimesitylethanal and 1,2,2-Trimesitylethanone. Conformations of 1,2,2-Trimesitylethanone and 1,2,2-
Trimesitylethanol J. Am. Chem. Soc. 1985, 107, 1007– 1015 DOI: 10.1021/ja00290a043 [ACS Full Text ],
[CAS], [Google Scholar]

188. Hansen, P. E.; Spanget-Lansen, J.; Laali, K. K. Conformational Studies of Phenyl and (1-Pyrenyl)
Triarylmethylcarbenium Ions: Semiempirical Calculations and NMR Investigations under Stable Ion Conditions J.
Org. Chem. 1998, 63, 1827– 1835 DOI: 10.1021/jo9715480 [ACS Full Text ], [CAS], [Google Scholar]

  
189. Yamaguchi, S.; Akiyama, S.; Tamao, K. Synthesis, Structures, Photophysical Properties, and Dynamic
Stereochemistry of Tri(9-anthryl)silane Derivatives Organometallics 1998, 17, 4347– 4352
 DOI: 10.1021/om9804778 [ACS Full Text ], [CAS], [Google Scholar]

190. Sedo, J.; Ventosa, N.; Molins, M. A.; Pons, M.; Rovira, C.; Veciana, J. Stereoisomerism of Molecular
Multipropellers. 2. Dynamic Stereochemistry of Bis- and Tris-Triaryl Systems J. Org. Chem. 2001, 66, 1579–
1589 DOI: 10.1021/jo000474g [ACS Full Text ], [CAS], [Google Scholar]

191. Grilli, S.; Lunazzi, L.; Mazzanti, A. Conformational Studies by Dynamic NMR. 83.1 Correlated
Enantiomerization Pathways for the Stereolabile Propeller Antipodes of Dimesityl Substituted Ethanol and Ethers
J. Org. Chem. 2001, 66, 5853– 5858 DOI: 10.1021/jo010420m [ACS Full Text ], [CAS], [Google Scholar]

192. Benincori, T.; Celentano, G.; Pilati, T.; Ponti, A.; Rizzo, S.; Sannicolò, F. Con gurationally Stable Molecular
Propellers: First Resolution of Residual Enantiomers Angew. Chem., Int. Ed. 2006, 45, 6193– 6196
 DOI: 10.1002/anie.200601869 [Crossref], [PubMed], [CAS], [Google Scholar]

193. Bulo, R. E.; Allaart, F.; Ehlers, A. W.; de Kanter, F. J. J.; Schakel, M.; Lutz, M.; Spek, A. L.; Lammertsma, K.
Circumambulatory Rearrangement with Characteristics of a 2:1 Covalent Molecular Bevel Gear J. Am. Chem.
Soc. 2006, 128, 12169– 12173 DOI: 10.1021/ja0627895 [ACS Full Text ], [CAS], [Google Scholar]

194. Romeo, R.; Carnabuci, S.; Fenech, L.; Plutino, M. R.; Albinati, A. Overcrowded Organometallic Platinum(II)
Complexes That Behave as Molecular Gears Angew. Chem., Int. Ed. 2006, 45, 4494– 4498
 DOI: 10.1002/anie.200600827 [Crossref], [PubMed], [CAS], [Google Scholar]

195. Peck, T.-G.; Lai, Y.-H. Conformational Isomerism in 1,2-Di-o-tolylnaphthalenes: Selective Rotation of the 2-
Aryl Ring Tetrahedron 2009, 65, 3664– 3667 DOI: 10.1016/j.tet.2009.02.068 [Crossref], [CAS], [Google Scholar]

196. Mati, I. K.; Cockroft, S. L. Molecular Balances for Quantifying Non-covalent Interactions Chem. Soc. Rev.
2010, 39, 4195– 4205 DOI: 10.1039/b822665m [Crossref], [PubMed], [CAS], [Google Scholar]

197. Hiraoka, S.; Harano, K.; Tanaka, T.; Shiro, M.; Shionoya, M. Quantitative Formation of Sandwich-shaped
Trinuclear Silver(I) Complexes and Dynamic Nature of Their P ≤ > M Flip Motion in Solution Angew. Chem., Int.
Ed. 2003, 42, 5182– 5185 DOI: 10.1002/anie.200351068 [Crossref], [PubMed], [CAS], [Google Scholar]
198. Hiraoka, S.; Shiro, M.; Shionoya, M. Heterotopic Assemblage of Two Different Disk-Shaped Ligands through
Trinuclear Silver (I) Complexation: Ligand Exchange-Driven Molecular Motion J. Am. Chem. Soc. 2004, 
 126, 
1214– 1218 DOI: 10.1021/ja036388q [ACS Full Text ], [CAS], [Google Scholar]

199. Hiraoka, S.; Okuno, E.; Tanaka, T.; Shiro, M.; Shionoya, M. Ranging Correlated Motion (1.5 nm) of Two
Coaxially Arranged Rotors Mediated by Helix Inversion of a Supramolecular Transmitter J. Am. Chem. Soc. 2008,
130, 9089– 9098 DOI: 10.1021/ja8014583 [ACS Full Text ], [CAS], [Google Scholar]

200. Hiraoka, S.; Hirata, K.; Shionoya, M. A Molecular Ball Bearing Mediated by Multiligand Exchange in Concert
Angew. Chem., Int. Ed. 2004, 43, 3814– 3818 DOI: 10.1002/anie.200453753 [Crossref], [PubMed], [CAS], [Google
Scholar]

201. Okuno, E.; Hiraoka, S.; Shionoya, M. A Synthetic Approach to a Molecular Crank Mechanism: Toward
Intramolecular Motion Transformation Between Rotation and Translation Dalton Trans. 2010, 39, 4107– 4116
 DOI: 10.1039/b926154k [Crossref], [PubMed], [CAS], [Google Scholar]

202. Liu, S.; Kondratuk, D. V.; Rousseaux, S. A.; Gil-Ramirez, G.; O’Sullivan, M. C.; Cremers, J.; Claridge, T. D.;
Anderson, H. L. Caterpillar Track Complexes in Template-directed Synthesis and Correlated Molecular Motion
Angew. Chem., Int. Ed. 2015, 54, 5355– 5359 DOI: 10.1002/anie.201412293 [Crossref], [PubMed], [CAS], [Google
Scholar]

203. Kelly, T. R.; Bowyer, M. C.; Bhaskar, K. V.; Bebbington, D.; Garcia, A.; Lang, F.; Kim, M. H.; Jette, M. P. A
Molecular Brake J. Am. Chem. Soc. 1994, 116, 3657– 3658 DOI: 10.1021/ja00087a085 [ACS Full Text ],
[CAS], [Google Scholar]

204. Kelly, T. R. Progress Toward a Rationally Designed Molecular Motor Acc. Chem. Res. 2001, 34, 514– 522
 DOI: 10.1021/ar000167x [ACS Full Text ], [CAS], [Google Scholar]

205. Sestelo, J. P.; Kelly, T. R. A Prototype of a Rationally Designed Chemically Powered Brownian Motor Appl.
Phys. A: Mater. Sci. Process. 2002, 75, 337– 343 DOI: 10.1007/s003390201328 [Crossref], [CAS], [Google
Scholar]

206. Jog, P. V.; Brown, R. E.; Bates, D. K. A Redox-Mediated Molecular Brake: Dynamic NMR Study of 2-[2-
(Methylthio)phenyl]isoindolin-1-one and S-Oxidized Counterparts J. Org. Chem. 2003, 68, 8240– 8243
 DOI: 10.1021/jo034613g [ACS Full Text ], [CAS], [Google Scholar]
207. Dial, B. E.; Pellechia, P. J.; Smith, M. D.; Shimizu, K. D. Proton Grease: An Acid Accelerated Molecular Rotor
J. Am. Chem. Soc. 2012, 134, 3675– 3678 DOI: 10.1021/ja2120184 [ACS Full Text ], [CAS], [Google 
Scholar] 

208. Dial, B. E.; Rasberry, R. D.; Bullock, B. N.; Smith, M. D.; Pellechia, P. J.; Profeta, S.; Shimizu, K. D. Guest-
Accelerated Molecular Rotor Org. Lett. 2011, 13, 244– 247 DOI: 10.1021/ol102659n [ACS Full Text ],
[CAS], [Google Scholar]

209. Kanazawa, H.; Higuchi, M.; Yamamoto, K. An Electric Cyclophane: Cavity Control Based on the Rotation of
a Paraphenylene by Redox Switching J. Am. Chem. Soc. 2005, 127, 16404– 16405 DOI: 10.1021/ja055681i [ACS
Full Text ], [CAS], [Google Scholar]

210. Tomohiro, Y.; Satake, A.; Kobuke, Y. Synthesis of Bipyridylene-Bridged Bisporphyrin by Nickel-Mediated
Coupling Reaction: ON–OFF Control of Cofacial Porphyrin Unit by Reversible Complexation J. Org. Chem. 2001,
66, 8442– 8446 DOI: 10.1021/jo015852b [ACS Full Text ], [CAS], [Google Scholar]

211. Zehm, D.; Fudickar, W.; Linker, T. Molecular Switches Flipped by Oxygen Angew. Chem., Int. Ed. 2007, 46,
7689– 7692 DOI: 10.1002/anie.200700443 [Crossref], [PubMed], [CAS], [Google Scholar]

212. Kelly, T. R.; Tellitu, I.; Sestelo, J. P. In Search of Molecular Ratchets Angew. Chem., Int. Ed. Engl. 1997, 36,
1866– 1868 DOI: 10.1002/anie.199718661 [Crossref], [CAS], [Google Scholar]

213. Kelly, T. R.; Sestelo, J. P.; Tellitu, I. New Molecular Devices: In Search of a Molecular Ratchet J. Org. Chem.
1998, 63, 3655– 3665 DOI: 10.1021/jo9723218 [ACS Full Text ], [CAS], [Google Scholar]

214. Sebastian, K. L. Molecular Ratchets: Veri cation of the Principle of Detailed Balance and the Second Law
of Dynamics Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 2000, 61, 937– 939
 DOI: 10.1103/PhysRevE.61.937 [Crossref], [PubMed], [CAS], [Google Scholar]

215. Davis, A. P. Tilting at Windmills? The Second Law Survives Angew. Chem., Int. Ed. 1998, 37, 909– 910
 DOI: 10.1002/(SICI)1521-3773(19980420)37:7<909::AID-ANIE909>3.0.CO;2-X [Crossref], [CAS], [Google Scholar]

216. Kelly, T. R.; De Silva, H.; Silva, R. A. Unidirectional Rotary Motion in a Molecular System Nature 1999, 401,
150– 152 DOI: 10.1038/43639 [Crossref], [PubMed], [CAS], [Google Scholar]

217. Kelly, T. R.; Silva, R. A.; Silva, H. D.; Jasmin, S.; Zhao, Y. A Rationally Designed Prototype of a Molecular
Motor J. Am. Chem. Soc. 2000, 122, 6935– 6949 DOI: 10.1021/ja001048f [ACS Full Text ], [CAS], [Google
Scholar]

  
218. Kelly, T. R.; Cai, X.; Damkaci, F.; Panicker, S. B.; Tu, B.; Bushell, S. M.; Cornella, I.; Piggott, M. J.; Salives, R.;
Cavero, M. Progress Toward a Rationally Designed, Chemically Powered Rotary Molecular Motor J. Am. Chem.
Soc. 2007, 129, 376– 386 DOI: 10.1021/ja066044a [ACS Full Text ], [CAS], [Google Scholar]

219. Mock, W. L.; Ochwat, K. J. Theory and Example of a Small-molecule Motor J. Phys. Org. Chem. 2003, 16,
175– 182 DOI: 10.1002/poc.591 [Crossref], [CAS], [Google Scholar]

220. Dahl, B. J.; Branchaud, B. P. Synthesis and Characterization of a Functionalized Chiral Biaryl Capable of
Exhibiting Unidirectional Bond Rotation Tetrahedron Lett. 2004, 45, 9599– 9602
 DOI: 10.1016/j.tetlet.2004.10.147 [Crossref], [CAS], [Google Scholar]

221. Lin, Y.; Dahl, B. J.; Branchaud, B. P. Net Directed 180° Aryl–aryl Bond Rotation in a Prototypical Achiral Biaryl
Lactone Synthetic Molecular Motor Tetrahedron Lett. 2005, 46, 8359– 8362 DOI: 10.1016/j.tetlet.2005.09.151
[Crossref], [CAS], [Google Scholar]

222. Fletcher, S. P.; Dumur, F.; Pollard, M. M.; Feringa, B. L. A Reversible, Unidirectional Molecular Rotary Motor
Driven by Chemical Energy Science 2005, 310, 80– 82 DOI: 10.1126/science.1117090 [Crossref], [PubMed],
[CAS], [Google Scholar]

223. Siegel, J. Chemistry. Inventing the Nanomolecular Wheel Science 2005, 310, 63– 64
 DOI: 10.1126/science.1118765 [Crossref], [PubMed], [CAS], [Google Scholar]

224. Tashiro, K.; Konishi, K.; Aida, T. Enantiomeric Resolution of Chiral Metallobis(porphyrin)s: Studies on
Rotatbility of Electronically Coupled Porphyrin Ligands Angew. Chem., Int. Ed. Engl. 1997, 36, 856– 858
 DOI: 10.1002/anie.199708561 [Crossref], [CAS], [Google Scholar]

225. Tashiro, K.; Fujiwara, T.; Konishi, K.; Aida, T. Chem. Commun. 1998, 1121– 1122 DOI: 10.1039/a801350k
[Crossref], [CAS], [Google Scholar]

226. Ikeda, M.; Takeuchi, M.; Shinkai, S.; Tani, F.; Naruta, Y. Synthesis of New Diaryl-Substituted Triple-Decker
and Tetraaryl-substituted Double-Decker Lanthanum(III) Porphyrins and Their Porphyrin Ring Rotational Speed
as Compared with that of Double-Decker Cerium(IV) Porphyrins Bull. Chem. Soc. Jpn. 2001, 74, 739– 746
 DOI: 10.1246/bcsj.74.739 [Crossref], [CAS], [Google Scholar]
227. Tashiro, K.; Konishi, K.; Aida, T. Metal Bisporphyrin Double-decker Complexes as Redox-Responsive
Rotating Modules. Studies on Ligand Rotation Activities of the Reduced and Oxidized Forms Using  as 
Chirality a
Probe J. Am. Chem. Soc. 2000, 122, 7921– 7926 DOI: 10.1021/ja000356a [ACS Full Text ], [CAS], [Google
Scholar]

228. Buchler, J. W.; Decian, A.; Fischer, J.; Hammerschmitt, P.; Lo er, J.; Scharbert, B.; Weiss, R. Metal-
Complexes with Tetrapyrrole Ligands 0.54. Synthesis, Spectra, Structure, and Redox Properties of Cerium(Iv)
Bisporphyrinates with Identical and Different Porphyrin Rings in the Sandwich System Chem. Ber. 1989, 122,
2219– 2228 DOI: 10.1002/cber.19891221203 [Crossref], [CAS], [Google Scholar]

229. Takeuchi, M.; Imada, I.; Shinkai, S. A Strong Positive Allosteric Effect in the Molecular Recognition of
Dicarboxylic Acids by a Cerium(IV)Bis[tetrakis(4-pyridyl)-porphyrinte] Double Decker Angew. Chem., Int. Ed. 1998,
37, 2096– 2099 DOI: 10.1002/(SICI)1521-3773(19980817)37:15<2096::AID-ANIE2096>3.0.CO;2-B [Crossref],
[CAS], [Google Scholar]

230. Sugasaki, A.; Ikeda, M.; Takeuchi, M.; Robertson, A.; Shinkai, S. E cient Chirality Transcription Utilizing a
Cerium(IV) Double Decker Porphyrin: A Prototype for Development of a Molecular Memory Systems J. Chem.
Soc., Perkin Trans. 1 1999, 3259– 3264 DOI: 10.1039/a904890a [Crossref], [CAS], [Google Scholar]

231. Yamamoto, M.; Sugasaki, A.; Ikeda, M.; Takeuchi, M.; Frimat, K.; James, T. D.; Shinkai, S. E cient Anion
Binding to Cerium(IV)Bis(porphyrinate) Double Decker Utilizing Positive Homotropic Allosterim Chem. Lett. 2001,
520– 521 DOI: 10.1246/cl.2001.520 [Crossref], [CAS], [Google Scholar]

232. Robertson, A.; Ikeda, M.; Takeuchi, M.; Shinkai, S. Allosteric Binding of K+ to Crown Ether Macrocycles
Appended to a Lanthanum Double Decker System Bull. Chem. Soc. Jpn. 2001, 74, 883– 888
 DOI: 10.1246/bcsj.74.883 [Crossref], [CAS], [Google Scholar]

233. Sugasaki, A.; Ikeda, M.; Takeuchi, M.; Shinkai, S. Novel Oligosaccharide Binding to the Cerium(IV)
Bis(porphyrinoate) Double Decker: Effective Ampli cation of a Binding Signal through Positive Homotropic
Allosterism Angew. Chem., Int. Ed. 2000, 39, 3839– 3842
 DOI: 10.1002/1521-3773(20001103)39:21<3839::AID-ANIE3839>3.0.CO;2-2 [Crossref], [CAS], [Google Scholar]

234. Ikeda, M.; Shinkai, S.; Osuka, A. Meso–meso-linked Porphyrin Dimer as a Novel Scaffold for the Selective
Binding of Oligosaccharides Chem. Commun. 2000, 1047– 1048 DOI: 10.1039/b003365k [Crossref],
[CAS], [Google Scholar]
235. Sugasaki, A.; Ikeda, M.; Takeuchi, M.; Koumoto, K.; Shinkai, S. The First Example of Positive Allosterism in
an Aqueous Saccharide-Binding System Designed on a Ce(IV) Bis(porphyrinate) Double Decker Scaffold
  
Tetrahedron 2000, 56, 4717– 4723 DOI: 10.1016/S0040-4020(00)00395-1 [Crossref], [CAS], [Google Scholar]

236. Ikeda, M.; Tanida, T.; Takeuchi, M.; Shinkai, S. Allosteric Silver (I) Ion Binding with Peripheral π Clefts of a Ce
(IV) Double Decker Porphyrin Org. Lett. 2000, 2, 1803– 1805 DOI: 10.1021/ol005807a [ACS Full Text ],
[CAS], [Google Scholar]

237. Ikeda, M.; Takeuchi, M.; Shinkai, S.; Tani, F.; Naruta, Y.; Sakamoto, S.; Yamaguchi, K. Allosteric Binding of a
Ag+ Ion to Cerium (IV) Bis-porphyrinates Enhances the Rotational Activity of Porphyrin Ligands Chem. - Eur. J.
2002, 8, 5541– 5550 DOI: 10.1002/1521-3765(20021216)8:24<5541::AID-CHEM5541>3.0.CO;2-X [Crossref],
[PubMed], [CAS], [Google Scholar]

238. Kubo, Y.; Ikeda, M.; Sugasaki, A.; Takeuchi, M.; Shinkai, S. A Porphyrin Tetramer for a Positive Homotropic
Allosteric Recognition System: E cient Binding Information Transduction Through Butadiynyl Axis Rotation
Tetrahedron Lett. 2001, 42, 7435– 7438 DOI: 10.1016/S0040-4039(01)01546-5 [Crossref], [CAS], [Google
Scholar]

239. Shinkai, S.; Ikeda, M.; Sugasaki, A.; Takeuchi, M. Positive Allosteric Systems Designed on Dynamic
Supramolecular Scaffolds: Toward Switching and Ampli cation of Guest A nity and Selectivity Acc. Chem. Res.
2001, 34, 494– 503 DOI: 10.1021/ar000177y [ACS Full Text ], [CAS], [Google Scholar]

240. Takeuchi, M.; Ikeda, M.; Sugasaki, A.; Shinkai, S. Molecular Design of Arti cial Molecular and Ion
Recognition Systems with Allosteric Guest Responses Acc. Chem. Res. 2001, 34, 865– 873
 DOI: 10.1021/ar0000410 [ACS Full Text ], [CAS], [Google Scholar]

241. Ayabe, M.; Ikeda, A.; Kubo, Y.; Takeuchi, M.; Shinkai, S. A Dendritic Porphyrin Receptor for C60 Which
Features a Profound Positive Allosteric Effect Angew. Chem., Int. Ed. 2002, 41, 2790– 2792
 DOI: 10.1002/1521-3773(20020802)41:15<2790::AID-ANIE2790>3.0.CO;2-W [Crossref], [PubMed], [CAS], [Google
Scholar]

242. Kubo, Y.; Sugasaki, A.; Ikeda, M.; Sugiyasu, K.; Sonoda, K.; Ikeda, A.; Takeuchi, M.; Shinkai, S. Cooperative C-
60 Binding to a Porphyrin Tetramer Arranged Around a p-Terphenyl Axis in 1:2 Host-guest Stoichiometry Org.
Lett. 2002, 4, 925– 928 DOI: 10.1021/ol017299q [ACS Full Text ], [CAS], [Google Scholar]

243. Ercolani, G. The Origin of Cooperativity in Double-Wheel Receptors. Freezing of Internal Rotation or Ligand-
Induced Torsional Strain? Org. Lett. 2005, 7, 803– 805 DOI: 10.1021/ol047620f [ACS Full Text ], [CAS], [Google
Scholar]

  
244. Shinkai, S.; Takeuchi, M. Molecular Design of Synthetic Receptors with Dynamic, Imprinting, and Allosteric
Functions Bull. Chem. Soc. Jpn. 2005, 78, 40– 51 DOI: 10.1246/bcsj.78.40 [Crossref], [CAS], [Google Scholar]

245. Ikeda, T.; Tsukahara, T.; Iino, R.; Takeuchi, M.; Noji, H. Motion Capture and Manipulation of a Single
Synthetic Molecular Rotor by Optical Microscopy Angew. Chem., Int. Ed. 2014, 53, 10082– 10085
 DOI: 10.1002/anie.201403091 [Crossref], [PubMed], [CAS], [Google Scholar]

246. Tanaka, H.; Ikeda, T.; Takeuchi, M.; Sada, K.; Shinkai, S.; Kawai, T. Molecular Rotation in Self-Assembled
Multidecker Porphyrin Complexes ACS Nano 2011, 5, 9575– 9582 DOI: 10.1021/nn203773p [ACS Full Text ],
[CAS], [Google Scholar]

247. Ogi, S.; Ikeda, T.; Wakabayashi, R.; Shinkai, S.; Takeuchi, M. A Bevel-gear-shaped Rotor Bearing a Double-
decker Porphyrin Complex Chem. - Eur. J. 2010, 16, 8285– 8290 DOI: 10.1002/chem.201000276 [Crossref],
[PubMed], [CAS], [Google Scholar]

248. Ogi, S.; Ikeda, T.; Wakabayashi, R.; Shinkai, S.; Takeuchi, M. Mechanically Interlocked Porphyrin Gears
Propagating Two Different Rotational Frequencies Eur. J. Org. Chem. 2011, 2011, 1831– 1836
 DOI: 10.1002/ejoc.201001656 [Crossref], [Google Scholar]

249. Ogi, S.; Ikeda, T.; Takeuchi, M. Synthetic Molecular Gear Based on Double-Decker Porphyrin Complexes J.
Inorg. Organomet. Polym. Mater. 2013, 23, 193– 199 DOI: 10.1007/s10904-012-9749-x [Crossref], [CAS], [Google
Scholar]

250. Shibata, M.; Tanaka, S.; Ikeda, T.; Shinkai, S.; Kaneko, K.; Ogi, S.; Takeuchi, M. Stimuli-Responsive Folding
and Unfolding of a Polymer Bearing Multiple Cerium(IV) Bis(porphyrinate) Joints: Mechano-imitation of the
Action of a Folding Ruler Angew. Chem., Int. Ed. 2013, 52, 397– 400 DOI: 10.1002/anie.201205584 [Crossref],
[PubMed], [CAS], [Google Scholar]

251. Hiraoka, S.; Hisanaga, Y.; Shiro, M.; Shionoya, M. A Molecular Double Ball Bearing: An AgI–PtII
Dodecanuclear Quadruple-Decker Complex with Three Rotors Angew. Chem., Int. Ed. 2010, 49, 1669– 1673
 DOI: 10.1002/anie.200905947 [Crossref], [PubMed], [CAS], [Google Scholar]

252. Bohn, R. K.; Haaland, A. On the Molecular Structure of Ferrocene, Fe(C5H5)2 J. Organomet. Chem. 1966, 5,
470– 476 DOI: 10.1016/S0022-328X(00)82382-7 [Crossref], [CAS], [Google Scholar]
253. Wang, X. B.; Dai, B.; Woo, H. K.; Wang, L. S. Intramolecular Rotation Through Proton Transfer: [Fe(η5-
C5H4CO2-)2] versus [(η5-C5H4CO2-)Fe(η5-C5H4CO2H)] Angew. Chem., Int. Ed. 2005, 44, 6022– 6024  
 DOI: 10.1002/anie.200501564 [Crossref], [PubMed], [CAS], [Google Scholar]

254. Crowley, J. D.; Steele, I. M.; Bosnich, B. Protonmotive Force: Development of Electrostatic Drivers for
Synthetic Molecular Motors Chem. - Eur. J. 2006, 12, 8935– 8951 DOI: 10.1002/chem.200500519 [Crossref],
[PubMed], [CAS], [Google Scholar]

255. Iordache, A.; Oltean, M.; Milet, A.; Thomas, F.; Baptiste, B.; Saint-Aman, E.; Bucher, C. Redox Control of
Rotary Motions in Ferrocene-based Elemental Ball Bearings J. Am. Chem. Soc. 2012, 134, 2653– 2671
 DOI: 10.1021/ja209766e [ACS Full Text ], [CAS], [Google Scholar]

256. Fukino, T.; Joo, H.; Hisada, Y.; Obana, M.; Yamagishi, H.; Hikima, T.; Takata, M.; Fujita, N.; Aida, T.
Manipulation of Discrete Nanostructures by Selective Modulation of Noncovalent Forces Science 2014, 344,
499– 504 DOI: 10.1126/science.1252120 [Crossref], [PubMed], [CAS], [Google Scholar]

257. Warren, L. F., Jr.; Hawthorne, M. F. The Chemistry of the Bis[π-(3)-1,2-dicarbollyl] Metalates of Nickel and
Palladium J. Am. Chem. Soc. 1970, 92, 1157– 1173 DOI: 10.1021/ja00708a009 [ACS Full Text ], [CAS], [Google
Scholar]

258. St. Clair, D.; Zalkin, A.; Templeton, D. H. The Crystal Structure of 3,3′-commo-Bis[undecahydro-1,2-dicarba-
3-nickela-closo-dodecaborane], a Nickel(IV) Complex of the Dicarbollide Ion J. Am. Chem. Soc. 1970, 92, 1173–
1179 DOI: 10.1021/ja00708a010 [ACS Full Text ], [CAS], [Google Scholar]

259. Gold, K.; Churchill, M. R. The Crystal Structure and Molecular Con guration of an Asymmetric 1,2-Dimethyl-
1,2-dicarbollide Complex of Nickel, Racemic (3,4′)-[(CH8)2B9C2H9]2Ni J. Am. Chem. Soc. 1970, 92, 1180– 1187
 DOI: 10.1021/ja00708a011 [ACS Full Text ], [CAS], [Google Scholar]

260. Hawthorne, M. F.; Dunks, G. B. Metallocarboranes That Exhibit Novel Chemical Features: A Virtually
Unlimited Variety of Structural and Dynamic Features are Observed in Metallocarborane Chemistry Science
1972, 178, 462– 471 DOI: 10.1126/science.178.4060.462 [Crossref], [PubMed], [CAS], [Google Scholar]

261. Hawthorne, M. F.; Zink, J. I.; Skelton, J. M.; Bayer, M. J.; Liu, C.; Livshits, E.; Baer, R.; Neuhauser, D. Electrical
or Photocontrol of the Rotary Motion of a Metallacarborane Science 2004, 303, 1849– 1851
 DOI: 10.1126/science.1093846 [Crossref], [PubMed], [CAS], [Google Scholar]
262. Rebek, J. Binding Forces, Equilibria and Rates: New Models for Enzymic Catalysis Acc. Chem. Res. 1984,
17, 258– 264 DOI: 10.1021/ar00103a006 [ACS Full Text ], [CAS], [Google Scholar]   

263. Rebek, J.; Trend, J. E.; Wattley, R. V.; Chakravorti, S. Allosteric Effects in Organic Chemistry. Site-speci c
Binding J. Am. Chem. Soc. 1979, 101, 4333– 4337 DOI: 10.1021/ja00509a047 [ACS Full Text ], [CAS], [Google
Scholar]

264. Rebek, J.; Wattley, R. V. Allosteric Effects. Remote Control of Ion Transport Selectivity J. Am. Chem. Soc.
1980, 102, 4853– 4854 DOI: 10.1021/ja00534a058 [ACS Full Text ], [CAS], [Google Scholar]

265. Rebek, J.; Costello, T.; Marshall, L.; Wattley, R.; Gadwood, R. C.; Onan, K. Allosteric Effects in Organic
Chemistry: Binding Cooperativity in a Model for Subunit Interactions J. Am. Chem. Soc. 1985, 107, 7481– 7487
 DOI: 10.1021/ja00311a043 [ACS Full Text ], [CAS], [Google Scholar]

266. van Veggel, F. C. J. M.; Verboom, W.; Reinhoudt, D. N. Metallomacrocycles: Supramolecular Chemistry with
Hard and Soft Metal Cations in Action Chem. Rev. 1994, 94, 279– 299 DOI: 10.1021/cr00026a001 [ACS Full Text
], [CAS], [Google Scholar]

267. Linton, B.; Hamilton, A. D. Formation of Arti cial Receptors by Metal-Templated Self-Assembly Chem. Rev.
1997, 97, 1669– 1680 DOI: 10.1021/cr960375w [ACS Full Text ], [CAS], [Google Scholar]

268. Robertson, A.; Shinkai, S. Cooperative Binding in Selective Sensors, Catalysts and Actuators Coord. Chem.
Rev. 2000, 205, 157– 199 DOI: 10.1016/S0010-8545(00)00243-5 [Crossref], [CAS], [Google Scholar]

269. Shinkai, S.; Ikeda, M.; Sugasaki, A.; Takeuchi, M. Positive Allosteric Systems Designed on Dynamic
Supramolecular Scaffolds: Toward Switching and Ampli cation of Guest A nity and Selectivity Acc. Chem. Res.
2001, 34, 494– 503 DOI: 10.1021/ar000177y [ACS Full Text ], [CAS], [Google Scholar]

270. Kovbasyuk, L.; Krämer, R. Allosteric Supramolecular Receptors and Catalysts Chem. Rev. 2004, 104, 3161–
3188 DOI: 10.1021/cr030673a [ACS Full Text ], [CAS], [Google Scholar]

271. Chong, Y. S.; Smith, M. D.; Shimizu, K. D. A Conformationally Programmable Ligand J. Am. Chem. Soc.
2001, 123, 7463– 7464 DOI: 10.1021/ja0158713 [ACS Full Text ], [CAS], [Google Scholar]

272. Degenhardt, C. F.; Lavin, J. M.; Smith, M. D.; Shimizu, K. D. Conformationally Imprinted Receptors:
Atropisomers with “Write”, “Save”, and “Erase” Recognition Properties Org. Lett. 2005, 7, 4079– 4081
 DOI: 10.1021/ol051325t [ACS Full Text ], [CAS], [Google Scholar]

  
273. Leighton, P.; Sanders, J. K. M. A Molecular Switch for Control of Conformation: Strained Intramolecular Co-
ordination in 4,4′-Bipyridyl-capped Zinc Porphyrins J. Chem. Soc., Chem. Commun. 1984, 854– 856
 DOI: 10.1039/c39840000854 [Crossref], [CAS], [Google Scholar]

274. Abraham, R. J.; Leighton, P.; Sanders, J. K. M. = Coordination Chemistry and Geometries of Some 4,4′-
Bipyridyl-capped Porphyrins. Proton- and Ligand-induced Switching of Conformations J. Am. Chem. Soc. 1985,
107, 3472– 3478 DOI: 10.1021/ja00298a012 [ACS Full Text ], [CAS], [Google Scholar]

275. Mendez-Arroyo, J.; Barroso-Flores, J.; Lifschitz, A. M.; Sarjeant, A. A.; Stern, C. L.; Mirkin, C. A. A Multi-state,
Allosterically-regulated Molecular Receptor with Switchable Selectivity J. Am. Chem. Soc. 2014, 136, 10340–
10348 DOI: 10.1021/ja503506a [ACS Full Text ], [CAS], [Google Scholar]

276. Kennedy, R. D.; Machan, C. W.; McGuirk, C. M.; Rosen, M. S.; Stern, C. L.; Sarjeant, A. A.; Mirkin, C. A. General
Strategy for the Synthesis of Rigid Weak-link Approach Platinum(II) Complexes: Tweezers, Triple-layer
Complexes, and Macrocycles Inorg. Chem. 2013, 52, 5876– 5888 DOI: 10.1021/ic302855f [ACS Full Text ],
[CAS], [Google Scholar]

277. Kuwabara, J.; Yoon, H. J.; Mirkin, C. A.; DiPasquale, A. G.; Rheingold, A. L. Pseudo-allosteric Regulation of
the Anion Binding A nity of a Macrocyclic Coordination Complex Chem. Commun. 2009, 4557– 4559
 DOI: 10.1039/b905150c [Crossref], [PubMed], [CAS], [Google Scholar]

278. Shinkai, S.; Nakaji, T.; Ogawa, T.; Shigematsu, K.; Manabe, O. Photoresponsive Crown Ethers. 2.
Photocontrol of Ion Extraction and Ion Transport by a Bis(crown ether) with a Butter y-like Motion J. Am. Chem.
Soc. 1981, 103, 111– 115 DOI: 10.1021/ja00391a021 [ACS Full Text ], [CAS], [Google Scholar]

279. Hunter, C. A.; Anderson, H. L. What is Cooperativity? Angew. Chem., Int. Ed. 2009, 48, 7488– 7499
 DOI: 10.1002/anie.200902490 [Crossref], [PubMed], [CAS], [Google Scholar]

280. Klarner, F. G.; Kahlert, B. Molecular Tweezers and Clips as Synthetic Receptors. Molecular Recognition and
Dynamics in Receptor-substrate Complexes Acc. Chem. Res. 2003, 36, 919– 932 DOI: 10.1021/ar0200448 [ACS
Full Text ], [CAS], [Google Scholar]

281. Harmata, M. Chiral Molecular Tweezers Acc. Chem. Res. 2004, 37, 862– 873 DOI: 10.1021/ar030164v
[ACS Full Text ], [CAS], [Google Scholar]
282. Poulsen, T.; Nielsen, K. A.; Bond, A. D.; Jeppesen, J. O. Bis(tetrathiafulvalene)-Calix[2]pyrrole[2]- thiophene
Text ],
and Its Complexation with TCNQ Org. Lett. 2007, 9, 5485– 5488 DOI: 10.1021/ol7024235 [ACS Full 
[CAS], [Google Scholar]

283. Sijbesma, R. P.; Wijmenga, S. S.; Nolte, R. J. M. A Molecular Clip That Binds Aromatic Guests by an
Induced- t Mechanism J. Am. Chem. Soc. 1992, 114, 9807– 9813 DOI: 10.1021/ja00051a013 [ACS Full Text ],
[CAS], [Google Scholar]

284. Rowan, A. E.; Elemans, J. A. A. W.; Nolte, R. J. M. Molecular and Supramolecular Objects from Glycoluril
Acc. Chem. Res. 1999, 32, 995– 1006 DOI: 10.1021/ar9702684 [ACS Full Text ], [CAS], [Google Scholar]

285. Reek, J. N. H.; Engelkamp, H.; Rowan, A. E.; Elemans, J. A. A. W.; Nolte, R. J. M. Conformational Behavior
and Binding Properties of Naphthalene-Walled Clips Chem. - Eur. J. 1998, 4, 716– 722
 DOI: 10.1002/(SICI)1521-3765(19980416)4:4<716::AID-CHEM716>3.0.CO;2-C [Crossref], [CAS], [Google Scholar]

286. Sijbesma, R. P.; Nolte, R. J. M. A Molecular Clip with Allosteric Binding Properties J. Am. Chem. Soc. 1991,
113, 6695– 6696 DOI: 10.1021/ja00017a063 [ACS Full Text ], [CAS], [Google Scholar]

287. Klarner, F. G.; Benkhoff, J.; Boese, R.; Burkert, U.; Kamieth, M.; Naatz, U. Molecular Tweezers as Synthetic
Receptors in Host-guest Chemistry: Inclusion of Cyclohexane and Self-assembly of Aliphatic Side Chains Angew.
Chem., Int. Ed. Engl. 1996, 35, 1130– 1133 DOI: 10.1002/anie.199611301 [Crossref], [Google Scholar]

288. Petitjean, A.; Khoury, R. G.; Kyritsakas, N.; Lehn, J.-M. Dynamic Devices. Shape Switching and Substrate
Binding in Ion-Controlled Nanomechanical Molecular Tweezers J. Am. Chem. Soc. 2004, 126, 6637– 6647
 DOI: 10.1021/ja031915r [ACS Full Text ], [CAS], [Google Scholar]

289. Hoffmann, R. W. Flexible Molecules with De ned Shape—Conformational Design Angew. Chem., Int. Ed.
Engl. 1992, 31, 1124– 1134 DOI: 10.1002/anie.199211241 [Crossref], [Google Scholar]

290. Yuasa, H.; Hashimoto, H. Bending Trisaccharides by a Chelation-Induced Ring Flip of a Hinge-Like
Monosaccharide Unit J. Am. Chem. Soc. 1999, 121, 5089– 5090 DOI: 10.1021/ja984062p [ACS Full Text ],
[CAS], [Google Scholar]

291. Izumi, T.; Hashimoto, H.; Yuasa, H. Switching Extended 1,3-Diequatorial and Bent 1,3-Diaxial States of a
Disubstituted Hinge Sugar by Ligand Exchange Reactions on Pt(II) Chem. Commun. 2004, 94– 95
 DOI: 10.1039/b311811h [Crossref], [PubMed], [CAS], [Google Scholar]
292. Menger, F. M.; Chicklo, P. A.; Sherrod, M. J. Ion-induced Conformational Changes in Kemp’s Triacid
 
Tetrahedron Lett. 1989, 30, 6943– 6946 DOI: 10.1016/S0040-4039(01)93393-3 [Crossref], [CAS], [Google 
Scholar]

293. Shirodkar, S. M.; Weisman, G. R. Cyclohexane-based 1,3-Dipodands: Complexation and Conformational


Biasing J. Chem. Soc., Chem. Commun. 1989, 236– 238 DOI: 10.1039/c39890000236 [Crossref], [CAS], [Google
Scholar]

294. Raban, M.; Burch, D. L.; Hortelano, E. R.; Durocher, D.; Kost, D. Complete Conformational Switching in a
Calcium Ionophore J. Org. Chem. 1994, 59, 1283– 1287 DOI: 10.1021/jo00085a014 [ACS Full Text ],
[CAS], [Google Scholar]

295. Kemp, D. S.; Petrakis, K. S. Synthesis and Conformational Analysis of cis,cis-1,3,5-Trimethylcyclohexane-


1,3,5-tricarboxylic acid J. Org. Chem. 1981, 46, 5140– 5143 DOI: 10.1021/jo00338a014 [ACS Full Text ],
[CAS], [Google Scholar]

296. Samoshin, V. V.; Chertkov, V. A.; Vatlina, L. P.; Dobretsova, E. K.; Simonov, N. A.; Kastorsky, L. P.;
Gremyachinsky, D. E.; Schneider, H.-J. trans-1,2-Cyclohexanedicarboxylic Acid Derivatives as pH-Trigger for
Conformationally Controlled Crowns Tetrahedron Lett. 1996, 37, 3981– 3984
 DOI: 10.1016/0040-4039(96)00761-7 [Crossref], [CAS], [Google Scholar]

297. Gil de Oliveira Santos, A.; Klute, W.; Torode, J.; P. W. Bohm, V.; Cabrita, E.; Runsink, J.; W. Hoffmann, R.
Flexible Molecules with De ned Shape. X. Synthesis and Conformational Study of 1,5-Diaza-cis-decalin New J.
Chem. 1998, 22, 993– 997 DOI: 10.1039/a801160e [Crossref], [CAS], [Google Scholar]

298. Collin, J.-P.; Durola, F.; Heitz, V.; Reviriego, F.; Sauvage, J.-P.; Trolez, Y. A Cyclic [4]rotaxane that Behaves as a
Switchable Molecular Receptor: Formation of a Rigid Scaffold from a Collapsed Structure by Complexation with
Copper(I) Ions Angew. Chem., Int. Ed. 2010, 49, 10172– 10175 DOI: 10.1002/anie.201004008 [Crossref],
[PubMed], [CAS], [Google Scholar]

299. Haberhauer, G. Control of Planar Chirality: The Construction of a Copper-Ion-Controlled Chiral Molecular
Hinge Angew. Chem., Int. Ed. 2008, 47, 3635– 3638 DOI: 10.1002/anie.200800062 [Crossref], [PubMed],
[CAS], [Google Scholar]

300. Haberhauer, G. A Metal-Ion-Driven Supramolecular Chirality Pendulum Angew. Chem., Int. Ed. 2010, 49,
9286– 9289 DOI: 10.1002/anie.201004460 [Crossref], [PubMed], [CAS], [Google Scholar]
301. Haberhauer, G.; Kallweit, C. A Bridged Azobenzene Derivative as a Reversible, Light-Induced Chirality
 
Switch Angew. Chem., Int. Ed. 2010, 49, 2418– 2421 DOI: 10.1002/anie.200906731 [Crossref], [PubMed], 
[CAS], [Google Scholar]

302. Timmerman, P.; Verboom, W.; Reinhoudt, D. N. Resorcinarenes Tetrahedron 1996, 52, 2663– 2704
 DOI: 10.1016/0040-4020(95)00984-1 [Crossref], [CAS], [Google Scholar]

303. Moran, J. R.; Ericson, J. L.; Dalcanale, E.; Bryant, J. A.; Knobler, C. B.; Cram, D. J. Vases and Kites as
Cavitands J. Am. Chem. Soc. 1991, 113, 5707– 5714 DOI: 10.1021/ja00015a026 [ACS Full Text ],
[CAS], [Google Scholar]

304. Moran, J. R.; Karbach, S.; Cram, D. J. Cavitands: Synthetic Molecular Vessels J. Am. Chem. Soc. 1982, 104,
5826– 5828 DOI: 10.1021/ja00385a064 [ACS Full Text ], [CAS], [Google Scholar]

305. Skinner, P. J.; Cheetham, A. G.; Beeby, A.; Gramlich, V.; Diederich, F. Conformational Switching of
Resorcin[4]arene Cavitands by Protonation, Preliminary Communication Helv. Chim. Acta 2001, 84, 2146– 2153
 DOI: 10.1002/1522-2675(20010711)84:7<2146::AID-HLCA2146>3.0.CO;2-K [Crossref], [CAS], [Google Scholar]

306. Azov, V. A.; Diederich, F.; Lill, Y.; Hecht, B. Synthesis and Conformational Switching of Partially and
Differentially Bridged Resorcin[4]arenes Bearing Fluorescent Dye Labels. Preliminary Communication Helv. Chim.
Acta 2003, 86, 2149– 2155 DOI: 10.1002/hlca.200390172 [Crossref], [CAS], [Google Scholar]

307. Azov, V. A.; Jaun, B.; Diederich, F. NMR Investigations into the Vase-Kite Conformational Switching of
Resorcin[4]arene Cavitands Helv. Chim. Acta 2004, 87, 449– 462 DOI: 10.1002/hlca.200490043 [Crossref],
[CAS], [Google Scholar]

308. Frei, M.; Marotti, F.; Diederich, F. ZnII-induced Conformational Control of Amphiphilic Cavitands in Langmuir
Monolayers Chem. Commun. 2004, 1362– 1363 DOI: 10.1039/b405331a [Crossref], [PubMed], [CAS], [Google
Scholar]

309. Pochorovski, I.; Ebert, M.-O.; Gisselbrecht, J.-P.; Boudon, C.; Schweizer, W. B.; Diederich, F. Redox-Switchable
Resorcin[4]arene Cavitands: Molecular Grippers J. Am. Chem. Soc. 2012, 134, 14702– 14705
 DOI: 10.1021/ja306473x [ACS Full Text ], [CAS], [Google Scholar]

310. Pochorovski, I.; Milić, J.; Kolarski, D.; Gropp, C.; Schweizer, W. B.; Diederich, F. Evaluation of Hydrogen-Bond
Acceptors for Redox-Switchable Resorcin[4]arene Cavitands J. Am. Chem. Soc. 2014, 136, 3852– 3858
 DOI: 10.1021/ja411429b [ACS Full Text ], [CAS], [Google Scholar]
 Rates
311. Rudkevich, D. M.; Hilmersson, G.; Rebek, J. Intramolecular Hydrogen Bonding Controls the Exchange 
of Guests in a Cavitand J. Am. Chem. Soc. 1997, 119, 9911– 9912 DOI: 10.1021/ja971592x [ACS Full Text ],
[CAS], [Google Scholar]

312. Rudkevich, D. M.; Hilmersson, G.; Rebek, J. Self-Folding Cavitands J. Am. Chem. Soc. 1998, 120, 12216–
12225 DOI: 10.1021/ja982970g [ACS Full Text ], [CAS], [Google Scholar]

313. Ma, S.; Rudkevich, D. M.; Rebek, J. J. Supramolecular Isomerism in Caviplexes Angew. Chem., Int. Ed. 1999,
38, 2600– 2602 DOI: 10.1002/(SICI)1521-3773(19990903)38:17<2600::AID-ANIE2600>3.0.CO;2-T [Crossref],
[PubMed], [CAS], [Google Scholar]

314. Tucci, F. C.; Rudkevich, D. M.; Rebek, J. J. Velcrands with Snaps and Their Conformational Control Chem. -
Eur. J. 2000, 6, 1007– 1016 DOI: 10.1002/(SICI)1521-3765(20000317)6:6<1007::AID-CHEM1007>3.0.CO;2-A
[Crossref], [PubMed], [CAS], [Google Scholar]

315. Haino, T.; Rudkevich, D. M.; Shivanyuk, A.; Rissanen, K.; Rebek, J. J. Induced-Fit Molecular Recognition with
Water-Soluble Cavitands Chem. - Eur. J. 2000, 6, 3797– 3805
 DOI: 10.1002/1521-3765(20001016)6:20<3797::AID-CHEM3797>3.0.CO;2-1 [Crossref], [PubMed], [CAS], [Google
Scholar]

316. Far, A. R.; Shivanyuk, A.; Rebek, J. Water-Stabilized Cavitands J. Am. Chem. Soc. 2002, 124, 2854– 2855
 DOI: 10.1021/ja012453p [ACS Full Text ], [CAS], [Google Scholar]

317. Amrhein, P.; Wash, P. L.; Shivanyuk, A.; Rebek, J. Metal Ligation Regulates Conformational Equilibria and
Binding Properties of Cavitands Org. Lett. 2002, 4, 319– 321 DOI: 10.1021/ol0167661 [ACS Full Text ],
[CAS], [Google Scholar]

318. Tucker, J. A.; Knobler, C. B.; Trueblood, K. N.; Cram, D. J. Host-guest Complexation. 49. Cavitands
Containing Two Binding Cavities J. Am. Chem. Soc. 1989, 111, 3688– 3699 DOI: 10.1021/ja00192a028 [ACS Full
Text ], [CAS], [Google Scholar]

319. Amrhein, P.; Shivanyuk, A.; Johnson, D. W.; Rebek, J. Metal-Switching and Self-Inclusion of Functional
Cavitands J. Am. Chem. Soc. 2002, 124, 10349– 10358 DOI: 10.1021/ja0204269 [ACS Full Text ],
[CAS], [Google Scholar]
320. Ikeda, A.; Shinkai, S. Novel Cavity Design Using Calix[n]arene Skeletons: Toward Molecular Recognition and
Metal Binding Chem. Rev. 1997, 97, 1713– 1734 DOI: 10.1021/cr960385x [ACS Full Text  
], [CAS], [Google 
Scholar]

321. Ashton, P. R.; Ballardini, R.; Balzani, V.; Boyd, S. E.; Credi, A.; Gandol , M. T.; Gómez-López, M.; Iqbal, S.; Philp,
D.; Preece, J. A. Simple Mechanical Molecular and Supramolecular Machines: Photochemical and
Electrochemical Control of Switching Processes Chem. - Eur. J. 1997, 3, 152– 170
 DOI: 10.1002/chem.19970030123 [Crossref], [CAS], [Google Scholar]

322. Ashton, P. R.; Gómez-López, M.; Iqbal, S.; Preece, J. A.; Stoddart, J. F. A Self-Complexing Macrocycle Acting
as a Chromophoric Receptor Tetrahedron Lett. 1997, 38, 3635– 3638 DOI: 10.1016/S0040-4039(97)00688-6
[Crossref], [CAS], [Google Scholar]

323. Brondsted Nielsen, M.; Becher, J. ’Self-complexing’ Tetrathiafulvalene Macrocycles; A Tetrathiafulvalene


Switch Chem. Commun. 1998, 475– 476 DOI: 10.1039/a707026h [Crossref], [Google Scholar]

324. Brøndsted Nielsen, M.; Hansen, J. G.; Becher, J. Self-Complexing Tetrathiafulvalene-Based Donor–
Acceptor Macrocycles Eur. J. Org. Chem. 1999, 1999, 2807– 2815
 DOI: 10.1002/(SICI)1099-0690(199911)1999:11<2807::AID-EJOC2807>3.3.CO;2-E [Crossref], [Google Scholar]

325. Balzani, V.; Ceroni, P.; Credi, A.; Gomez-Lopez, M.; Hamers, C.; Stoddart, J. F.; Wolf, R. Controlled
Dethreading/rethreading of a Scorpion-like Pseudorotaxane and a Related Macrobicyclic Self-complexing
System New J. Chem. 2001, 25, 25– 31 DOI: 10.1039/b004934o [Crossref], [CAS], [Google Scholar]

326. Liu, Y.; Flood, A. H.; Stoddart, J. F. Thermally and Electrochemically Controllable Self-Complexing Molecular
Switches J. Am. Chem. Soc. 2004, 126, 9150– 9151 DOI: 10.1021/ja048164t [ACS Full Text ], [CAS], [Google
Scholar]

327. Qu, D.-H.; Feringa, B. L. Controlling Molecular Rotary Motion with a Self-Complexing Lock Angew. Chem.,
Int. Ed. 2010, 49, 1107– 1110 DOI: 10.1002/anie.200906064 [Crossref], [PubMed], [CAS], [Google Scholar]

328. Jones, I. M.; Lingard, H.; Hamilton, A. D. pH-Dependent Conformational Switching in 2,6-
Benzamidodiphenylacetylenes Angew. Chem., Int. Ed. 2011, 50, 12569– 12571 DOI: 10.1002/anie.201106241
[Crossref], [PubMed], [CAS], [Google Scholar]

329. Pramanik, S.; De, S.; Schmittel, M. Bidirectional Chemical Communication between Nanomechanical
Switches Angew. Chem., Int. Ed. 2014, 53, 4709– 4713 DOI: 10.1002/anie.201400804 [Crossref], [PubMed],
[CAS], [Google Scholar]

  
330. Dolain, C.; Maurizot, V.; Huc, I. Protonation-Induced Transition between Two Distinct Helical Conformations
of a Synthetic Oligomer via a Linear Intermediate Angew. Chem., Int. Ed. 2003, 42, 2738– 2740
 DOI: 10.1002/anie.200351080 [Crossref], [PubMed], [CAS], [Google Scholar]

331. Piguet, C.; Bernardinelli, G.; Hopfgartner, G. Helicates as Versatile Supramolecular Complexes Chem. Rev.
1997, 97, 2005– 2062 DOI: 10.1021/cr960053s [ACS Full Text ], [CAS], [Google Scholar]

332. Hill, D. J.; Mio, M. J.; Prince, R. B.; Hughes, T. S.; Moore, J. S. A Field Guide to Foldamers Chem. Rev. 2001,
101, 3893– 4012 DOI: 10.1021/cr990120t [ACS Full Text ], [CAS], [Google Scholar]

333. Cheng, R. P.; Gellman, S. H.; DeGrado, W. F. β-Peptides: From Structure to Function Chem. Rev. 2001, 101,
3219– 3232 DOI: 10.1021/cr000045i [ACS Full Text ], [CAS], [Google Scholar]

334. Albrecht, M. Let’s Twist Again”Double-Stranded, Triple-Stranded, and Circular Helicates Chem. Rev. 2001,
101, 3457– 3498 DOI: 10.1021/cr0103672 [ACS Full Text ], [CAS], [Google Scholar]

335. Williams, A. Helical Complexes and Beyond Chem. - Eur. J. 1997, 3, 15– 19
 DOI: 10.1002/chem.19970030104 [Crossref], [CAS], [Google Scholar]

336. Gellman, S. H. Foldamers: A Manifesto Acc. Chem. Res. 1998, 31, 173– 180 DOI: 10.1021/ar960298r [ACS
Full Text ], [CAS], [Google Scholar]

337. Rowan, A. E.; Nolte, R. J. M. Helical Molecular Programming Angew. Chem., Int. Ed. 1998, 37, 63– 68
 DOI: 10.1002/(SICI)1521-3773(19980202)37:1/2<63::AID-ANIE63>3.0.CO;2-4 [Crossref], [CAS], [Google Scholar]

338. Cubberley, M. S.; Iverson, B. L. Models of Higher-order Structure: Foldamers and Beyond Curr. Opin. Chem.
Biol. 2001, 5, 650– 653 DOI: 10.1016/S1367-5931(01)00261-7 [Crossref], [PubMed], [CAS], [Google Scholar]

339. Schmuck, C. Molecules with Helical Structure: How To Build a Molecular Spiral Staircase Angew. Chem.,
Int. Ed. 2003, 42, 2448– 2452 DOI: 10.1002/anie.200201625 [Crossref], [PubMed], [CAS], [Google Scholar]

340. Albrecht, M. Arti cial Molecular Double-Stranded Helices Angew. Chem., Int. Ed. 2005, 44, 6448– 6451
 DOI: 10.1002/anie.200501472 [Crossref], [PubMed], [CAS], [Google Scholar]
341. Gong, B. Crescent Oligoamides: From Acyclic “Macrocycles” to Folding Nanotubes Chem. - Eur.
 J. 2001,
  7,
4336– 4342 DOI: 10.1002/1521-3765(20011015)7:20<4336::AID-CHEM4336>3.0.CO;2-1 [Crossref], [PubMed],
[CAS], [Google Scholar]

342. Sanford, A. R.; Yamato, K.; Yang, X.; Yuan, L.; Han, Y.; Gong, B. Well-de ned Secondary Structures Eur. J.
Biochem. 2004, 271, 1416– 1425 DOI: 10.1111/j.1432-1033.2004.04062.x [Crossref], [PubMed], [CAS], [Google
Scholar]

343. Huc, I. Aromatic Oligoamide Foldamers Eur. J. Org. Chem. 2004, 2004, 17– 29
 DOI: 10.1002/ejoc.200300495 [Crossref], [Google Scholar]

344. Constable, E. C. Oligopyridines as Helicating Ligands Tetrahedron 1992, 48, 10013– 10059
 DOI: 10.1016/S0040-4020(01)89035-9 [Crossref], [CAS], [Google Scholar]

345. Seebach, D.; Matthews, J. β-Peptides: A Surprise at Every Turn Chem. Commun. 1997, 2015– 2022
 DOI: 10.1039/a704933a [Crossref], [CAS], [Google Scholar]

346. Nielsen, P. E.; Haaima, G. Peptide Nucleic Acid (PNA). A DNA Mimic with a Pseudopeptide Backbone
Chem. Soc. Rev. 1997, 26, 73– 78 DOI: 10.1039/cs9972600073 [Crossref], [CAS], [Google Scholar]

347. Nowick, J. S. Chemical Models of Protein β-Sheets Acc. Chem. Res. 1999, 32, 287– 296
 DOI: 10.1021/ar970204t [ACS Full Text ], [CAS], [Google Scholar]

348. Stigers, K. D.; Soth, M. J.; Nowick, J. S. Designed Molecules that Fold to Mimic Protein Secondary
Structures Curr. Opin. Chem. Biol. 1999, 3, 714– 723 DOI: 10.1016/S1367-5931(99)00030-7 [Crossref], [PubMed],
[CAS], [Google Scholar]

349. Nielsen, P. E. Peptide Nucleic Acid. A Molecule with Two Identities Acc. Chem. Res. 1999, 32, 624– 630
 DOI: 10.1021/ar980010t [ACS Full Text ], [CAS], [Google Scholar]

350. Kirshenbaum, K.; Zuckermann, R. N.; Dill, K. A. Designing Polymers That Mimic Biomolecules Curr. Opin.
Struct. Biol. 1999, 9, 530– 535 DOI: 10.1016/S0959-440X(99)80075-X [Crossref], [PubMed], [CAS], [Google
Scholar]
351. Herdewijn, P. Conformationally Restricted Carbohydrate-modi ed Nucleic Acids and Antisense Technology
  
Biochim. Biophys. Acta, Gene Struct. Expression 1999, 1489, 167– 179 DOI: 10.1016/S0167-4781(99)00152-9
[Crossref], [PubMed], [CAS], [Google Scholar]

352. Patch, J. A.; Barron, A. E. Mimicry of Bioactive Peptides via Non-natural, Sequence-speci c Peptidomimetic
Oligomers Curr. Opin. Chem. Biol. 2002, 6, 872– 877 DOI: 10.1016/S1367-5931(02)00385-X [Crossref], [PubMed],
[CAS], [Google Scholar]

353. Martinek, T. A.; Fülöp, F. Side-chain Control of β-Peptide Secondary Structures Eur. J. Biochem. 2003, 270,
3657– 3666 DOI: 10.1046/j.1432-1033.2003.03756.x [Crossref], [PubMed], [CAS], [Google Scholar]

354. Cheng, R. P. Beyond de novo Protein Design — de novo Design of Non-natural Folded Oligomers Curr. Opin.
Struct. Biol. 2004, 14, 512– 520 DOI: 10.1016/j.sbi.2004.07.001 [Crossref], [PubMed], [CAS], [Google Scholar]

355. Licini, G.; Prins, L. J.; Scrimin, P. Oligopeptide Foldamers: From Structure to Function Eur. J. Org. Chem.
2005, 2005, 969– 977 DOI: 10.1002/ejoc.200400521 [Crossref], [Google Scholar]

356. Nelson, J. C.; Saven, J. G.; Moore, J. S.; Wolynes, P. G. Solvophobically Driven Folding of Nonbiological
Oligomers Science 1997, 277, 1793– 1796 DOI: 10.1126/science.277.5333.1793 [Crossref], [PubMed],
[CAS], [Google Scholar]

357. Prince, R. B.; Saven, J. G.; Wolynes, P. G.; Moore, J. S. Cooperative Conformational Transitions in Phenylene
Ethynylene Oligomers: Chain-Length Dependence J. Am. Chem. Soc. 1999, 121, 3114– 3121
 DOI: 10.1021/ja983995i [ACS Full Text ], [CAS], [Google Scholar]

358. Lahiri, S.; Thompson, J. L.; Moore, J. S. Solvophobically Driven π-Stacking of Phenylene Ethynylene
Macrocycles and Oligomers J. Am. Chem. Soc. 2000, 122, 11315– 11319 DOI: 10.1021/ja002129e [ACS Full
Text ], [CAS], [Google Scholar]

359. Hill, D. J.; Moore, J. S. Helicogenicity of Solvents in the Conformational Equilibrium of Oligo(m-phenylene
ethynylene)s: Implications for Foldamer Research Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 5053– 5057
 DOI: 10.1073/pnas.072642799 [Crossref], [PubMed], [CAS], [Google Scholar]

360. Hofacker, A. L.; Parquette, J. R. Dendrimer Folding in Aqueous Media: An Example of Solvent-Mediated
Chirality Switching Angew. Chem., Int. Ed. 2005, 44, 1053– 1057 DOI: 10.1002/anie.200460943 [Crossref],
[PubMed], [CAS], [Google Scholar]
361. Berl, V.; Huc, I.; Khoury, R. G.; Krische, M. J.; Lehn, J.-M. Interconversion of Single and Double Helices
 
Formed from Synthetic Molecular Strands Nature 2000, 407, 720– 723 DOI: 10.1038/35037545 [Crossref], 
[PubMed], [CAS], [Google Scholar]

362. Berl, V.; Huc, I.; Khoury, R. G.; Lehn, J.-M. Helical Molecular Programming: Folding of Oligopyridine-
dicarboxamides into Molecular Single Helices Chem. - Eur. J. 2001, 7, 2798– 2809
 DOI: 10.1002/1521-3765(20010702)7:13<2798::AID-CHEM2798>3.0.CO;2-L [Crossref], [PubMed], [CAS], [Google
Scholar]

363. Berl, V.; Huc, I.; Khoury, R. G.; Lehn, J.-M. Helical Molecular Programming: Supramolecular Double Helices
by Dimerization of Helical Oligopyridine-dicarboxamide Strands Chem. - Eur. J. 2001, 7, 2810– 2820
 DOI: 10.1002/1521-3765(20010702)7:13<2810::AID-CHEM2810>3.0.CO;2-5 [Crossref], [PubMed], [CAS], [Google
Scholar]

364. Hanan, G. S.; Lehn, J.-M.; Kyritsakas, N.; Fischer, J. Molecular Helicity: A General Approach for Helicity
Induction in a Polyheterocyclic Molecular Strand J. Chem. Soc., Chem. Commun. 1995, 765– 766
 DOI: 10.1039/c39950000765 [Crossref], [CAS], [Google Scholar]

365. Bassani, D. M.; Lehn, J.-M.; Baum, G.; Fenske, D. Designed Self-Generation of an Extended Helical Structure
From an Achiral Polyheterocylic Strand Angew. Chem., Int. Ed. Engl. 1997, 36, 1845– 1847
 DOI: 10.1002/anie.199718451 [Crossref], [CAS], [Google Scholar]

366. Ohkita, M.; Lehn, J.-M.; Baum, G.; Fenske, D. Helicity Coding: Programmed Molecular Self-Organization of
Achiral Nonbiological Strands into Multiturn Helical Superstructures: Synthesis and Characterization of
Alternating Pyridine–Pyrimidine Oligomers Chem. - Eur. J. 1999, 5, 3471– 3481
 DOI: 10.1002/(SICI)1521-3765(19991203)5:12<3471::AID-CHEM3471>3.0.CO;2-5 [Crossref], [CAS], [Google
Scholar]

367. Stadler, A.-M.; Kyritsakas, N.; Lehn, J.-M. Reversible Folding/unfolding of Linear Molecular Strands into
Helical Channel-like Complexes upon Proton-modulated Binding and Release of Metal Ions Chem. Commun.
2004, 2024– 2025 DOI: 10.1039/b407168a [Crossref], [PubMed], [CAS], [Google Scholar]

368. Kolomiets, E.; Berl, V.; Odriozola, I.; Stadler, A.-M.; Kyritsakas, N.; Lehn, J.-M. Contraction/extension
Molecular Motion by Protonation/deprotonation Induced Structural Switching of Pyridine Derived Oligoamides
Chem. Commun. 2003, 2868– 2869 DOI: 10.1039/b311578j [Crossref], [PubMed], [CAS], [Google Scholar]
369. Barboiu, M.; Lehn, J.-M. Dynamic Chemical Devices: Modulation of Contraction/extension Molecular
Motion by Coupled-ion Binding/pH Change-induced Structural Switching Proc. Natl. Acad. Sci. U. S.  99,
A. 2002,
5201– 5206 DOI: 10.1073/pnas.082099199 [Crossref], [PubMed], [CAS], [Google Scholar]

370. Barboiu, M.; Vaughan, G.; Kyritsakas, N.; Lehn, J.-M. Dynamic Chemical Devices: Generation of Reversible
Extension/Contraction Molecular Motion by Ion-Triggered Single/Double Helix Interconversion Chem. - Eur. J.
2003, 9, 763– 769 DOI: 10.1002/chem.200390085 [Crossref], [PubMed], [CAS], [Google Scholar]

371. Stadler, A.-M.; Lehn, J.-M. P. Coupled Nanomechanical Motions: Metal-Ion-Effected, pH-Modulated,
Simultaneous Extension/Contraction Motions of Double-Domain Helical/Linear Molecular Strands J. Am. Chem.
Soc. 2014, 136, 3400– 3409 DOI: 10.1021/ja408752m [ACS Full Text ], [CAS], [Google Scholar]

372. Dugave, C.; Demange, L. cis-trans Isomerization of Organic Molecules and Biomolecules: Implications and
Applications Chem. Rev. 2003, 103, 2475– 2532 DOI: 10.1021/cr0104375 [ACS Full Text ], [CAS], [Google
Scholar]

373. Waldeck, D. H. Photoisomerization Dynamics of Stilbenes Chem. Rev. 1991, 91, 415– 436
 DOI: 10.1021/cr00003a007 [ACS Full Text ], [CAS], [Google Scholar]

374. Bandara, H. M.; Burdette, S. C. Photoisomerization in Different Classes of Azobenzene Chem. Soc. Rev.
2012, 41, 1809– 1825 DOI: 10.1039/C1CS15179G [Crossref], [PubMed], [CAS], [Google Scholar]

375. Cheetham, A. G.; Hutchings, M. G.; Claridge, T. D. W.; Anderson, H. L. Enzymatic Synthesis and
Photoswitchable Enzymatic Cleavage of a Peptide-linked Rotaxane Angew. Chem., Int. Ed. 2006, 45, 1596– 1599
 DOI: 10.1002/anie.200504064 [Crossref], [PubMed], [CAS], [Google Scholar]

376. Haberhauer, G.; Kallweit, C.; Wölper, C.; Bläser, D. An Azobenzene Unit Embedded in a Cyclopeptide as a
Type-Speci c and Spatially Directed Switch Angew. Chem., Int. Ed. 2013, 52, 7879– 7882
 DOI: 10.1002/anie.201301516 [Crossref], [PubMed], [CAS], [Google Scholar]

377. Clever, G. H.; Tashiro, S.; Shionoya, M. Light-Triggered Crystallization of a Molecular Host-Guest Complex J.
Am. Chem. Soc. 2010, 132, 9973– 9975 DOI: 10.1021/ja103620z [ACS Full Text ], [CAS], [Google Scholar]

378. Irie, M. Diarylethenes for Memories and Switches Chem. Rev. 2000, 100, 1685– 1716
 DOI: 10.1021/cr980069d [ACS Full Text ], [CAS], [Google Scholar]
379. Irie, M. Photochromism: Memories and Switches - Introduction Chem. Rev. 2000, 100, 1683– 1683
 DOI: 10.1021/cr980068l [ACS Full Text ], [CAS], [Google Scholar]   

380. van Herpt, J. T.; Areephong, J.; Stuart, M. C.; Browne, W. R.; Feringa, B. L. Light-controlled Formation of
Vesicles and Supramolecular Organogels by a Cholesterol-bearing Amphiphilic Molecular Switch Chem. - Eur. J.
2014, 20, 1737– 1742 DOI: 10.1002/chem.201302902 [Crossref], [PubMed], [CAS], [Google Scholar]

381. Hou, L.; Zhang, X.; Pijper, T. C.; Browne, W. R.; Feringa, B. L. Reversible Photochemical Control of Singlet
Oxygen Generation Using Diarylethene Photochromic Switches J. Am. Chem. Soc. 2014, 136, 910– 913
 DOI: 10.1021/ja4122473 [ACS Full Text ], [CAS], [Google Scholar]

382. Kudernac, T.; Kobayashi, T.; Uyama, A.; Uchida, K.; Nakamura, S.; Feringa, B. L. Tuning the Temperature
Dependence for Switching in Dithienylethene Photochromic Switches J. Phys. Chem. A 2013, 117, 8222– 8229
 DOI: 10.1021/jp404924q [ACS Full Text ], [CAS], [Google Scholar]

383. Yokoyama, Y. Fulgides for Memories and Switches Chem. Rev. 2000, 100, 1717– 1739
 DOI: 10.1021/cr980070c [ACS Full Text ], [CAS], [Google Scholar]

384. Santiago, A.; Becker, R. S. Photochromic Fulgides - Spectroscopy and Mechanism of Photoreactions J.
Am. Chem. Soc. 1968, 90, 3654– 3658 DOI: 10.1021/ja01016a009 [ACS Full Text ], [CAS], [Google Scholar]

385. Berkovic, G.; Weiss, V. Spiropyrans and Spirooxazines for Memories and Switches Chem. Rev. 2000, 100,
1741– 1753 DOI: 10.1021/cr9800715 [ACS Full Text ], [CAS], [Google Scholar]

386. Feringa, B. L.; Huck, N. P. M.; Schoevaars, A. M. Chiroptical Molecular Switches Adv. Mater. 1996, 8, 681–
684 DOI: 10.1002/adma.19960080819 [Crossref], [CAS], [Google Scholar]

387. Feringa, B. L.; van Delden, R. A.; Koumura, N.; Geertsema, E. M. Chiroptical Molecular Switches Chem. Rev.
2000, 100, 1789– 1816 DOI: 10.1021/cr9900228 [ACS Full Text ], [CAS], [Google Scholar]

388. Browne, B. L. F. a. W. Molecular Switches, 2nd ed.; Wiley-VCH: Weinheim, 2011. [Crossref], [Google Scholar]

389. Shinkai, S.; Honda, Y.; Kusano, Y.; Manabe, O. A Photoresponsive Cylindrical Ionophore J. Chem. Soc.,
Chem. Commun. 1982, 848– 850 DOI: 10.1039/c39820000848 [Crossref], [CAS], [Google Scholar]
390. Feringa, B. L.; Jager, W. F.; de Lange, B. Organic Materials for Reversible Optical Data Storage Tetrahedron
1993, 49, 8267– 8310 DOI: 10.1016/S0040-4020(01)81913-X [Crossref], [CAS], [Google Scholar]   

391. Momotake, A.; Arai, T. Synthesis, Excited State Properties, and Dynamic Structural Change of
Photoresponsive Dendrimers Polymer 2004, 45, 5369– 5390 DOI: 10.1016/j.polymer.2004.05.075 [Crossref],
[CAS], [Google Scholar]

392. Momotake, A.; Arai, T. Photochemistry and Photophysics of Stilbene Dendrimers and Related Compounds
J. Photochem. Photobiol., C 2004, 5, 1– 25 DOI: 10.1016/j.jphotochemrev.2004.01.001 [Crossref], [CAS], [Google
Scholar]

393. Tie, C.; Gallucci, J. C.; Parquette, J. R. Helical Conformational Dynamics and Photoisomerism of
Alternating Pyridinedicarboxamide/m-(Phenylazo)azobenzene Oligomers J. Am. Chem. Soc. 2006, 128, 1162–
1171 DOI: 10.1021/ja0547228 [ACS Full Text ], [CAS], [Google Scholar]

394. Yager, K. G.; Barrett, C. J. Novel Photo-switching Using Azobenzene Functional Materials J. Photochem.
Photobiol., A 2006, 182, 250– 261 DOI: 10.1016/j.jphotochem.2006.04.021 [Crossref], [CAS], [Google Scholar]

395. Russew, M. M.; Hecht, S. Photoswitches: From Molecules to Materials Adv. Mater. 2010, 22, 3348– 3360
 DOI: 10.1002/adma.200904102 [Crossref], [PubMed], [CAS], [Google Scholar]

396. Fischer, G. Peptidyl-Prolyl cis/trans Isomerases and Their Effectors Angew. Chem., Int. Ed. Engl. 1994, 33,
1415– 1436 DOI: 10.1002/anie.199414151 [Crossref], [Google Scholar]

397. Fischer, G. Chemical Aspects of Peptide Bond Isomerization Chem. Soc. Rev. 2000, 29, 119– 127
 DOI: 10.1039/a803742f [Crossref], [CAS], [Google Scholar]

398. Beharry, A. A.; Woolley, G. A. Azobenzene Photoswitches for Biomolecules Chem. Soc. Rev. 2011, 40,
4422– 4437 DOI: 10.1039/c1cs15023e [Crossref], [PubMed], [CAS], [Google Scholar]

399. Beharry, A. A.; Wong, L.; Tropepe, V.; Woolley, G. A. Fluorescence Imaging of Azobenzene Photoswitching In
Vivo Angew. Chem., Int. Ed. 2011, 50, 1325– 1327 DOI: 10.1002/anie.201006506 [Crossref], [PubMed],
[CAS], [Google Scholar]

400. Szymanski, W.; Beierle, J. M.; Kistemaker, H. A.; Velema, W. A.; Feringa, B. L. Reversible Photocontrol of
Biological Systems by the Incorporation of Molecular Photoswitches Chem. Rev. 2013, 113, 6114– 6178
 DOI: 10.1021/cr300179f [ACS Full Text ], [CAS], [Google Scholar]

  
401. Kienzler, M. A.; Reiner, A.; Trautman, E.; Yoo, S.; Trauner, D.; Isacoff, E. Y. A Red-Shifted, Fast-Relaxing
Azobenzene Photoswitch for Visible Light Control of an Ionotropic Glutamate Receptor J. Am. Chem. Soc. 2013,
135, 17683– 17686 DOI: 10.1021/ja408104w [ACS Full Text ], [CAS], [Google Scholar]

402. Kamiya, Y.; Asanuma, H. Light-Driven DNA Nanomachine with a Photoresponsive Molecular Engine Acc.
Chem. Res. 2014, 47, 1663– 1672 DOI: 10.1021/ar400308f [ACS Full Text ], [CAS], [Google Scholar]

403. Velema, W. A.; Szymanski, W.; Feringa, B. L. Photopharmacology: Beyond Proof of Principle J. Am. Chem.
Soc. 2014, 136, 2178– 2191 DOI: 10.1021/ja413063e [ACS Full Text ], [CAS], [Google Scholar]

404. Velema, W. A.; van der Berg, J. P.; Hansen, M. J.; Szymanski, W.; Driessen, A. J.; Feringa, B. L. Optical
Control of Antibacterial Activity Nat. Chem. 2013, 5, 924– 928 DOI: 10.1038/nchem.1750 [Crossref], [PubMed],
[CAS], [Google Scholar]

405. Szymanski, W.; Yilmaz, D.; Kocer, A.; Feringa, B. L. Bright Ion Channels and Lipid Bilayers Acc. Chem. Res.
2013, 46, 2910– 2923 DOI: 10.1021/ar4000357 [ACS Full Text ], [CAS], [Google Scholar]

406. Kocer, A.; Walko, M.; Meijberg, W.; Feringa, B. L. A Light-actuated Nanovalve Derived from a Channel Protein
Science 2005, 309, 755– 758 DOI: 10.1126/science.1114760 [Crossref], [PubMed], [CAS], [Google Scholar]

407. Muraoka, T.; Kinbara, K.; Aida, T. Mechanical Twisting of a Guest by a Photoresponsive Host Nature 2006,
440, 512– 515 DOI: 10.1038/nature04635 [Crossref], [PubMed], [CAS], [Google Scholar]

408. Muraoka, T.; Kinbara, K.; Aida, T. Reversible Operation of Chiral Molecular Scissors by Redox and UV Light
Chem. Commun. 2007, 1441– 1443 DOI: 10.1039/b618248h [Crossref], [PubMed], [CAS], [Google Scholar]

409. Merino, E.; Ribagorda, M. Control Over Molecular Motion Using the cis-trans Photoisomerization of the Azo
Group Beilstein J. Org. Chem. 2012, 8, 1071– 1090 DOI: 10.3762/bjoc.8.119 [Crossref], [PubMed], [CAS], [Google
Scholar]

410. Marchi, E.; Baroncini, M.; Bergamini, G.; Van Heyst, J.; Vogtle, F.; Ceroni, P. Photoswitchable Metal
Coordinating Tweezers Operated by Light-harvesting Dendrimers J. Am. Chem. Soc. 2012, 134, 15277– 15280
 DOI: 10.1021/ja307522f [ACS Full Text ], [CAS], [Google Scholar]
411. Raymo, F. M. Intermolecular Coupling of Motion Under Photochemical Control Angew. Chem., Int. Ed.
2006, 45, 5249– 5251 DOI: 10.1002/anie.200602516 [Crossref], [PubMed], [CAS], [Google Scholar]  

412. Kai, H.; Nara, S.; Kinbara, K.; Aida, T. Toward Long-distance Mechanical Communication: Studies on a
Ternary Complex Interconnected by a Bridging Rotary Module J. Am. Chem. Soc. 2008, 130, 6725– 6727
 DOI: 10.1021/ja801646b [ACS Full Text ], [CAS], [Google Scholar]

413. Muraoka, T.; Kinbara, K.; Kobayashi, Y.; Aida, T. Light-Driven Open-Close Motion of Chiral Molecular Scissors
J. Am. Chem. Soc. 2003, 125, 5612– 5613 DOI: 10.1021/ja034994f [ACS Full Text ], [CAS], [Google Scholar]

414. Norikane, Y.; Tamaoki, N. Light-Driven Molecular Hinge: A New Molecular Machine Showing a Light-
Intensity-Dependent Photoresponse that Utilizes the Trans-Cis Isomerization of Azobenzene Org. Lett. 2004, 6,
2595– 2598 DOI: 10.1021/ol049082c [ACS Full Text ], [CAS], [Google Scholar]

415. Muraoka, T.; Kinbara, K. Development of Photoresponsive Supramolecular Machines Inspired by Biological
Molecular Systems J. Photochem. Photobiol., C 2012, 13, 136– 147 DOI: 10.1016/j.jphotochemrev.2012.04.001
[Crossref], [CAS], [Google Scholar]

416. Li, Z.; Liang, J.; Xue, W.; Liu, G.; Liu, S. H.; Yin, J. Switchable Azo-macrocycles: From Molecules to
Functionalisation Supramol. Chem. 2014, 26, 54– 65 DOI: 10.1080/10610278.2013.822970 [Crossref],
[CAS], [Google Scholar]

417. Basheer, M. C.; Oka, Y.; Mathews, M.; Tamaoki, N. A Light-controlled Molecular Brake with Complete ON-
OFF Rotation Chem. - Eur. J. 2010, 16, 3489– 3496 DOI: 10.1002/chem.200902123 [Crossref], [PubMed],
[CAS], [Google Scholar]

418. Hashim, P. K.; Thomas, R.; Tamaoki, N. Induction of Molecular Chirality by Circularly Polarized Light in
Cyclic Azobenzene with a Photoswitchable Benzene Rotor Chem. - Eur. J. 2011, 17, 7304– 7312
 DOI: 10.1002/chem.201003526 [Crossref], [PubMed], [CAS], [Google Scholar]

419. Baroncini, M.; Silvi, S.; Venturi, M.; Credi, A. Reversible Photoswitching of Rotaxane Character and Interplay
of Thermodynamic Stability and Kinetic Lability in a Self-assembling Ring-axle Molecular System Chem. - Eur. J.
2010, 16, 11580– 11587 DOI: 10.1002/chem.201001409 [Crossref], [PubMed], [CAS], [Google Scholar]

420. Liu, F.; Morokuma, K. Computational Study on the Working Mechanism of a Stilbene Light-Driven Molecular
Rotary Motor: Sloped Minimal Energy Path and Unidirectional Nonadiabatic Photoisomerization J. Am. Chem.
Soc. 2012, 134, 4864– 4876 DOI: 10.1021/ja211441n [ACS Full Text ], [CAS], [Google Scholar]

 J.;Teplý,
421. Pospíšil, L.; Bednárová, L.; Štěpánek, P.; Slavíček, P.; Vávra, J.; Hromadová, M.; Dlouhá, H.; Tarábek,
F. Intense Chiroptical Switching in a Dicationic Helicene-Like Derivative: Exploration of a Viologen-Type Redox
Manifold of a Non-Racemic Helquat J. Am. Chem. Soc. 2014, 136, 10826– 10829 DOI: 10.1021/ja500220j [ACS
Full Text ], [CAS], [Google Scholar]

422. Ruangsupapichat, N.; Pollard, M. M.; Harutyunyan, S. R.; Feringa, B. L. Reversing the Direction in a Light-
driven Rotary Molecular Motor Nat. Chem. 2011, 3, 53– 60 DOI: 10.1038/nchem.872 [Crossref], [PubMed],
[CAS], [Google Scholar]

423. Feringa, B. L. In Control of Motion: From Molecular Switches to Molecular Motors Acc. Chem. Res. 2001,
34, 504– 513 DOI: 10.1021/ar0001721 [ACS Full Text ], [CAS], [Google Scholar]

424. Feringa, B. L.; van Delden, R. A.; ter Wiel, M. K. J. In Control of Switching, Motion, and Organization Pure
Appl. Chem. 2003, 75, 563– 575 DOI: 10.1351/pac200375050563 [Crossref], [CAS], [Google Scholar]

425. Feringa, B. L. The Art of Building Small: From Molecular Switches to Molecular Motors J. Org. Chem. 2007,
72, 6635– 6652 DOI: 10.1021/jo070394d [ACS Full Text ], [CAS], [Google Scholar]

426. Feringa, B. L.; Koumura, N.; van Delden, R. A.; ter Wiel, M. K. J. Light-driven Molecular Switches and Motors
Appl. Phys. A: Mater. Sci. Process. 2002, 75, 301– 308 DOI: 10.1007/s003390201338 [Crossref], [CAS], [Google
Scholar]

427. Feringa, B. L. Molecular Switches and Motors AIP Conf. Proc. 2013, 1519, 73– 75 DOI: 10.1063/1.4794713
[Crossref], [CAS], [Google Scholar]

428. Cnossen, A.; Browne, W. R.; Feringa, B. L. Unidirectional Light-Driven Molecular Motors Based on
Overcrowded Alkenes Top. Curr. Chem. 2014, 354, 139– 162 DOI: 10.1007/128_2013_512 [Crossref], [PubMed],
[CAS], [Google Scholar]

429. Koumura, N.; Zijlstra, R. W. J.; van Delden, R. A.; Harada, N.; Feringa, B. L. Light-driven Monodirectional
Molecualr Rotor Nature 1999, 401, 152– 155 DOI: 10.1038/43646 [Crossref], [PubMed], [CAS], [Google Scholar]

430. Harada, N.; Koumura, N.; Feringa, B. L. Chemistry of Unique Chiral Ole ns. 3. synthesis and Absolute
Stereochemistry of trans-and cis-1,1′,2,2′3,3′,4,4′-Octahydro-3,3′-dimethyl-4,4′-biphenanthrylidenes J. Am. Chem.
Soc. 1997, 119, 7256– 7264 DOI: 10.1021/ja970669e [ACS Full Text ], [CAS], [Google Scholar]
431. Harada, N.; Saito, A.; Koumura, N.; Uda, H.; deLange, B.; Jager, W. F.; Wynberg, H.; Feringa, B. L.  
Chemistry of
Unique Chiral Ole ns 0.1. Synthesis, Enantioresolution, Circular Dichroism, and Theoretical Determination of the
Absolute Stereochemistry of trans- and cis-1,1′,2,2′,3,3′,4,4′-octahydro-4,4′-biphenanthrylidenes J. Am. Chem.
Soc. 1997, 119, 7241– 7248 DOI: 10.1021/ja970667u [ACS Full Text ], [CAS], [Google Scholar]

432. Harada, N.; Saito, A.; Koumura, N.; Roe, D. C.; Jager, W. F.; Zijlstra, R. W. J.; deLange, B.; Feringa, B. L.
Chemistry of Unique Chiral Ole ns 0.2. Unexpected Thermal Racemization of cis-1,1′,2,2′,3,3′,4,4′-octahydro-4,4′-
biphenanthrylidene J. Am. Chem. Soc. 1997, 119, 7249– 7255 DOI: 10.1021/ja970668m [ACS Full Text ],
[CAS], [Google Scholar]

433. Zijlstra, R. W. J.; van Duijnen, P. T.; Feringa, B. L.; Steffen, T.; Duppen, K.; Wiersma, D. A. Excited-State
Dynamics of Tetraphenylethylene: Ultrafast Stokes Shift, Isomerization, and Charge Separation J. Phys. Chem. A
1997, 101, 9828– 9836 DOI: 10.1021/jp971763k [ACS Full Text ], [CAS], [Google Scholar]

434. Koumura, N.; Harada, N. Photochemistry and Absolute Stereochemistry of Unique Chiral Ole ns, trans- and
cis-1,1′,2,2′,3,3′,4,4′-Octahydro-3,3′-dimethyl-4,4′-biphenanthrylidenes Chem. Lett. 1998, 1151– 1152
 DOI: 10.1246/cl.1998.1151 [Crossref], [CAS], [Google Scholar]

435. Zijlstra, R. W. J.; Jager, W. F.; de Lange, B.; van Duijnen, P. Th.; Feringa, B. L.; Goto, H.; Saito, A.; Koumura, N.;
Harada, N. Chemistry of Unique Chiral Ole ns. 4. Theoretical Studies of the Racemization Mechanism of trans-
and cis-1,1′,2,2′,3,3′,4,4′-Octahydro-4,4′-biphenanthrylidenes J. Org. Chem. 1999, 64, 1667– 1674
 DOI: 10.1021/jo982381t [ACS Full Text ], [CAS], [Google Scholar]

436. Zijlstra, R. W. J.; Jager, W. F.; de Lange, B.; van Duijnen, P. T.; Feringa, B. L.; Goto, H.; Saito, A.; Koumura, N.;
Harada, N. Chemistry of Unique Chiral Ole ns. 4. Theoretical Studies of the Racemization Mechanism of trans-
and cis-1,1 ′,2,2 ′,3,3 ′,4,4 ′-octahydro-4,4 ′-biphenanthrylidenes J. Org. Chem. 1999, 64, 1667– 1674
 DOI: 10.1021/jo982381t [ACS Full Text ], [CAS], [Google Scholar]

437. ter Wiel, M. K. J.; Koumura, N.; Van Delden, R. A.; Meetsma, A.; Harada, N.; Feringa, B. L. Chiral Overcrowded
Alkenes; Asymmetric Synthesis of (3S,3′S)-(M,M)-(E)-(+)-1,1′,2,2′,3,3′,4,4′-octahydro-3,3′,7,7′-Tetramethyl-4,4′-
biphenanthrylidenes Chirality 2000, 12, 734– 741
 DOI: 10.1002/1520-636X(2000)12:10<734::AID-CHIR6>3.0.CO;2-Q [Crossref], [PubMed], [CAS], [Google Scholar]

438. ter Wiel, M. K. J.; van Delden, R. A.; Meetsma, A.; Feringa, B. L. Increased Speed of Rotation for The
Smallest Light-Driven Molecular Motor J. Am. Chem. Soc. 2003, 125, 15076– 15086 DOI: 10.1021/ja036782o
[ACS Full Text ], [CAS], [Google Scholar]
439. ter Wiel, M. K. J.; van Delden, R. A.; Meetsma, A.; Feringa, B. L. Light-Driven Molecular Motors: Stepwise
Thermal Helix Inversion during Unidirectional Rotation of sterically Overcrowded Biphenanthrylidenes 
 J. Am. 
Chem. Soc. 2005, 127, 14208– 14222 DOI: 10.1021/ja052201e [ACS Full Text ], [CAS], [Google Scholar]

440. Kuwahara, S.; Fujita, T.; Harada, N. A New Model of Light-Powered Chiral Molecular Motor with Higher
Speed of Rotation, Part 2 - Dynamics of Motor Rotation Eur. J. Org. Chem. 2005, 2005, 4544– 4556
 DOI: 10.1002/ejoc.200500323 [Crossref], [Google Scholar]

441. Klok, M.; Walko, M.; Geertsema, E. M.; Ruangsupapichat, N.; Kistemaker, J. C. M.; Meetsma, A.; Feringa, B.
L. New Mechanistic Insight in the Thermal Helix Inversion of Second - Second Generation Moelcular Motors
Chem. - Eur. J. 2008, 14, 11183– 11193 DOI: 10.1002/chem.200800969 [Crossref], [PubMed], [CAS], [Google
Scholar]

442. ter Wiel, M. K.; Kwit, M. G.; Meetsma, A.; Feringa, B. L. Synthesis, Stereochemistry, and Photochemical and
Thermal Behavior of Bis-tert-butyl Substituted Overcrowded Alkenes Org. Biomol. Chem. 2007, 5, 87– 96
 DOI: 10.1039/B611070C [Crossref], [PubMed], [CAS], [Google Scholar]

443. Pollard, M. M.; Meetsma, A.; Feringa, B. L. A Redesign of Light-driven Rotary Molecular Motors Org. Biomol.
Chem. 2008, 6, 507– 512 DOI: 10.1039/B715652A [Crossref], [PubMed], [CAS], [Google Scholar]

444. Caroli, G.; Kwit, M. G.; Feringa, B. L. Photochemical and Thermal Behavior of Light-driven Unidirectional
Molecular Motor with Long Alkyl Chains Tetrahedron 2008, 64, 5956– 5962 DOI: 10.1016/j.tet.2008.03.098
[Crossref], [CAS], [Google Scholar]

445. Koumura, N.; Geertsma, E. M.; Meetsma, A.; Feringa, B. L. Light-Driven Molecular Rotor: Unidirectional
Rotation Controlled by a Single Stereogenic Center J. Am. Chem. Soc. 2000, 122, 12005– 12006
 DOI: 10.1021/ja002755b [ACS Full Text ], [CAS], [Google Scholar]

446. Pollard, M. M.; ter Wiel, M. K. J.; van Delden, R. A.; Vicario, J.; Koumura, N.; van den Brom, J. R.; Meetsma,
A.; Feringa, B. L. Light-Driven Rotary Molecular Motors on Gold Nanoparticles Chem. - Eur. J. 2008, 14, 11610–
11622 DOI: 10.1002/chem.200800814 [Crossref], [PubMed], [CAS], [Google Scholar]

447. van Delden, R. A.; Koumura, N.; Schoevaars, A.; Meetsma, A.; Feringa, B. L. A Donor–acceptor Substituted
Molecular Motor: Unidirectional Rotation Driven by Visible Light Org. Biomol. Chem. 2003, 1, 33– 35
 DOI: 10.1039/b209378b [Crossref], [PubMed], [CAS], [Google Scholar]
448. Geertsema, E. M.; Koumura, N.; ter Wiel, M. K. J.; Meetsma, A.; Feringa, B. L. In Control of the Speed of
Rotation in Molecular Motors. Unexpected Retardation of Rotary Motion Chem. Commun. 2002, 2962– 
 2963 
 DOI: 10.1039/b208323j [Crossref], [PubMed], [CAS], [Google Scholar]

449. Koumura, N.; Geertsema, E. M.; van Gelder, M. B.; Meetsma, A.; Feringa, B. L. Second Generation Light-
Driven Molecular Motors. Unidirectional Rotation Controlled by a Single Stereogenic Center with Near-Perfect
Photoequilibria and Acceleration of the Speed of Rotation by Structural Modi cation J. Am. Chem. Soc. 2002,
124, 5037– 5051 DOI: 10.1021/ja012499i [ACS Full Text ], [CAS], [Google Scholar]

450. ter Wiel, M. K.; van Delden, R. A.; Meetsma, A.; Feringa, B. L. Increased Speed of Rotation for the Smallest
Light-Driven Molecular Motor J. Am. Chem. Soc. 2003, 125, 15076– 15086 DOI: 10.1021/ja036782o [ACS Full
Text ], [CAS], [Google Scholar]

451. Pijper, D.; van Delden, R. A.; Meetsma, A.; Feringa, B. L. Acceleration of a Nanomotor: Electronic Control of
the Rotary Speed of a Light-Driven Molecular Rotor J. Am. Chem. Soc. 2005, 127, 17612– 17613
 DOI: 10.1021/ja054499e [ACS Full Text ], [CAS], [Google Scholar]

452. Vicario, J.; Walko, M.; Meetsma, A.; Feringa, B. L. Fine Tuning of the Rotary Motion by Structural
Modi cation in Light-driven Unidirectional Molecular Motors J. Am. Chem. Soc. 2006, 128, 5127– 5135
 DOI: 10.1021/ja058303m [ACS Full Text ], [CAS], [Google Scholar]

453. Conyard, J.; Addison, K.; Heisler, I. A.; Cnossen, A.; Browne, W. R.; Feringa, B. L.; Meech, S. R. Ultrafast
Dynamics in the Power Stroke of a Molecular Rotary Motor Nat. Chem. 2012, 4, 547– 551
 DOI: 10.1038/nchem.1343 [Crossref], [PubMed], [CAS], [Google Scholar]

454. Schoevaars, A. M.; Kruizinga, W.; Zijlstra, R. W. J.; Veldman, N.; Spek, A. L.; Feringa, B. L. Toward a
Switchable Molecular Rotor. Unexpected Dynamic Behavior of Functionalized Overcrowded Alkenes J. Org.
Chem. 1997, 62, 4943– 4948 DOI: 10.1021/jo962210t [ACS Full Text ], [CAS], [Google Scholar]

455. ter Wiel, M. K.; van Delden, R. A.; Meetsma, A.; Feringa, B. L. Control of Rotor Motion in a Light-driven
Molecular Motor: Towards a Molecular Gearbox Org. Biomol. Chem. 2005, 3, 4071– 4076
 DOI: 10.1039/b510641a [Crossref], [PubMed], [CAS], [Google Scholar]

456. Greb, L.; Lehn, J.-M. Light-Driven Molecular Motors: Imines as Four-Step or Two-Step Unidirectional Rotors
J. Am. Chem. Soc. 2014, 136, 13114– 13117 DOI: 10.1021/ja506034n [ACS Full Text ], [CAS], [Google Scholar]
457. Padwa, A. Photochemistry of the Crbon-Nitrogen Double Bond Chem. Rev. 1977, 77, 37– 68
 DOI: 10.1021/cr60305a004 [ACS Full Text ], [CAS], [Google Scholar]   

458. Pratt, A. C. The Photochemistry of Imines Chem. Soc. Rev. 1977, 6, 63– 81 DOI: 10.1039/cs9770600063
[Crossref], [CAS], [Google Scholar]

459. Pierre Courot, R. P.; le Saint, Jaques Photochromisme par Isomerization Syn-anti de Phenylhydrazones-2
de Tricetones-1,2,3 et de Dicetones-1,2 Substituees Tetrahedron Lett. 1976, 17, 1181– 1184
 DOI: 10.1016/S0040-4039(00)78012-9 [Crossref], [Google Scholar]

460. Pichon, R.; le Saint, J.; Courtot, P. Photoisomerization d’Arylhydrazones-2 de Dicetones-1,2 Substituees en
2 Mecanisme d’Isomerization Thermique de la Double Liaison C═N Tetrahedron 1981, 37, 1517– 1524
 DOI: 10.1016/S0040-4020(01)92091-5 [Crossref], [CAS], [Google Scholar]

461. Chaur, M. N.; Collado, D.; Lehn, J.-M. Con gurational and Constitutional Information Storage: Multiple
Dynamics in Systems Based on Pyridyl and Acyl Hydrazones Chem. - Eur. J. 2011, 17, 248– 258
 DOI: 10.1002/chem.201002308 [Crossref], [PubMed], [CAS], [Google Scholar]

462. Burdette, S. C. Molecular Switches: Hydrazones Double Down on Zinc Nat. Chem. 2012, 4, 695– 696
 DOI: 10.1038/nchem.1438 [Crossref], [PubMed], [CAS], [Google Scholar]

463. Tatum, L. A.; Su, X.; Aprahamian, I. Simple Hydrazone Building Blocks for Complicated Functional Materials
Acc. Chem. Res. 2014, 47, 2141– 2149 DOI: 10.1021/ar500111f [ACS Full Text ], [CAS], [Google Scholar]

464. Su, X.; Aprahamian, I. Hydrazone-based Switches, Metallo-assemblies and Sensors Chem. Soc. Rev. 2014,
43, 1963– 1981 DOI: 10.1039/c3cs60385g [Crossref], [PubMed], [CAS], [Google Scholar]

465. Su, X.; Aprahamian, I. Zinc(II)-Regulation of Hydrazone Switch Isomerization Kinetics Org. Lett. 2013, 15,
5952– 5955 DOI: 10.1021/ol402789y [ACS Full Text ], [CAS], [Google Scholar]

466. Croteau, M. L.; Su, X.; Wilcox, D. E.; Aprahamian, I. Metal Coordination and Isomerization of a Hydrazone
Switch ChemPlusChem 2014, 79, 1214– 1224 DOI: 10.1002/cplu.201402134 [Crossref], [CAS], [Google Scholar]

467. Foy, J. T.; Ray, D.; Aprahamian, I. Regulating Signal Enhancement with Coordination-coupled Deprotonation
of a Hydrazone Switch Chem. Sci. 2015, 6, 209– 213 DOI: 10.1039/C4SC02882A [Crossref], [CAS], [Google
Scholar]
468. Langde, S. M.; Aprahamian, I. A ph Activated Con gurational Rotary Switch: Controlling the E
/Z  
Isomerization in Hydrazones J. Am. Chem. Soc. 2009, 131, 18269– 18271 DOI: 10.1021/ja909149z [ACS Full
Text ], [Google Scholar]

469. Su, X.; Robbins, T. F.; Aprahamian, I. Switching Through Coordination-coupled Proton Transfer Angew.
Chem., Int. Ed. 2011, 50, 1841– 1844 DOI: 10.1002/anie.201006982 [Crossref], [PubMed], [CAS], [Google Scholar]

470. Landge, S. M.; Tkatchouk, E.; Benitez, D.; Lanfranchi, D. A.; Elhabiri, M.; Goddard, W. A., III; Aprahamian, I.
Isomerization Mechanism in Hydrazone-based Rotary Switches: Lateral Shift, Rotation, or Tautomerization? J.
Am. Chem. Soc. 2011, 133, 9812– 9823 DOI: 10.1021/ja200699v [ACS Full Text ], [CAS], [Google Scholar]

471. Ray, D.; Foy, J. T.; Hughes, R. P.; Aprahamian, I. A Switching Cascade of Hydrazone-based Rotary Switches
Through Coordination-coupled Proton Relays Nat. Chem. 2012, 4, 757– 762 DOI: 10.1038/nchem.1408
[Crossref], [PubMed], [CAS], [Google Scholar]

472. Su, X.; Voskian, S.; Hughes, R. P.; Aprahamian, I. Manipulating Liquid-Crystal Properties Using a pH
Activated Hydrazone Switch Angew. Chem., Int. Ed. 2013, 52, 10734– 10739 DOI: 10.1002/anie.201305514
[Crossref], [PubMed], [CAS], [Google Scholar]

473. Siegel, J. S. Supramolecular Chemistry. Concepts and Perspectives - Lehn, J. -M Science 1996, 271, 949–
949 DOI: 10.1126/science.271.5251.949 [Crossref], [CAS], [Google Scholar]

474. Amendola, V.; Fabbrizzi, L.; Mangano, C.; Pallavicini, P. Molecular Movements and Translocations
Controlled by Transition Metals and Signaled by Light Emission Struct. Bonding (Berlin) 2001, 99, 79–
115 [Crossref], [CAS], [Google Scholar]

475. Miyazawa, A.; Fujiyoshi, Y.; Unwin, N. Structure and Gating Mechanism of the Acetylcholine Receptor Pore
Nature 2003, 423, 949– 955 DOI: 10.1038/nature01748 [Crossref], [PubMed], [CAS], [Google Scholar]

476. Kaempfer, R. Ribosomal Subunit Exchange during Protein Synthesis Proc. Natl. Acad. Sci. U. S. A. 1968,
61, 106– 113 DOI: 10.1073/pnas.61.1.106 [Crossref], [PubMed], [CAS], [Google Scholar]

477. Nakahata, M.; Takashima, Y.; Hashidzume, A.; Harada, A. Redox-Generated Mechanical Motion of a
Supramolecular Polymeric Actuator Based on Host-Guest Interactions Angew. Chem., Int. Ed. 2013, 52, 5731–
5735 DOI: 10.1002/anie.201300862 [Crossref], [PubMed], [CAS], [Google Scholar]
478. De Santis, G.; Fabbrizzi, L.; Iacopino, D.; Pallavicini, P.; Perotti, A.; Poggi, A. Electrochemically Switched Anion
Translocation in a Multicomponent Coordination Compound Inorg. Chem. 1997, 36, 827– 832   
 DOI: 10.1021/ic960892x [ACS Full Text ], [Google Scholar]

479. Akutagawa, T.; Koshinaka, H.; Sato, D.; Takeda, S.; Noro, S. I.; Takahashi, H.; Kumai, R.; Tokura, Y.; Nakamura,
T. Ferroelectricity and Polarity Control in Solid-state Flip- op Supramolecular Rotators Nat. Mater. 2009, 8, 342–
347 DOI: 10.1038/nmat2377 [Crossref], [PubMed], [CAS], [Google Scholar]

480. Ikeda, A.; Tsudera, T.; Shinkai, S. Molecular Design of a ’’Molecular Syringe’’ Mimic for Metal Cations Using a
1,3-Alternate Calix[4]arene Cavity J. Org. Chem. 1997, 62, 3568– 3574 DOI: 10.1021/jo962040k [ACS Full Text
], [CAS], [Google Scholar]

481. Knipe, P. C.; Thompson, S.; Hamilton, A. D. Ion-mediated Conformational Switches Chem. Sci. 2015, 6,
1630– 1639 DOI: 10.1039/C4SC03525A [Crossref], [PubMed], [CAS], [Google Scholar]

482. Su, X.; Robbins, T. F.; Aprahamian, I. Switching Through Coordination-Coupled Proton Transfer Angew.
Chem., Int. Ed. 2011, 50, 1841– 1844 DOI: 10.1002/anie.201006982 [Crossref], [PubMed], [CAS], [Google Scholar]

483. Su, X.; Voskian, S.; Hughes, R. P.; Aprahamian, I. Manipulating Liquid-Crystal Properties Using a pH
Activated Hydrazone Switch Angew. Chem., Int. Ed. 2013, 52, 10734– 10739 DOI: 10.1002/anie.201305514
[Crossref], [PubMed], [CAS], [Google Scholar]

484. Su, X.; Lessing, T.; Aprahamian, I. The Importance of the Rotor in Hydrazone-based Molecular Switches
Beilstein J. Org. Chem. 2012, 8, 872– 876 DOI: 10.3762/bjoc.8.98 [Crossref], [PubMed], [CAS], [Google Scholar]

485. Su, X.; Aprahamian, I. Switching Around Two Axles: Controlling the Con guration and Conformation of a
Hydrazone-Based Switch Org. Lett. 2011, 13, 30– 33 DOI: 10.1021/ol102422h [ACS Full Text ], [CAS], [Google
Scholar]

486. del Barrio, J.; Horton, P. N.; Lairez, D.; Lloyd, G. O.; Toprakcioglu, C.; Scherman, O. A. Photocontrol over
Cucurbit[8]uril Complexes: Stoichiometry and Supramolecular Polymers J. Am. Chem. Soc. 2013, 135, 11760–
11763 DOI: 10.1021/ja406556h [ACS Full Text ], [CAS], [Google Scholar]

487. Lan, Y.; Wu, Y. C.; Karas, A.; Scherman, O. A. Photoresponsive Hybrid Raspberry-Like Colloids Based on
Cucurbit[8]uril Host-Guest Interactions Angew. Chem., Int. Ed. 2014, 53, 2166– 2169
 DOI: 10.1002/anie.201309204 [Crossref], [PubMed], [CAS], [Google Scholar]
488. Ashton, P. R.; Campbell, P. J.; Chrystal, E. J. T.; Glink, P. T.; Menzer, S.; Philp, D.; Spencer, N.; Stoddart, J. F.;
Tasker, P. A.; Williams, D. J. Dialkylammonium Ion Crown-Ether Complexes - the Forerunners of a New  of
 Family
Interlocked Molecules Angew. Chem., Int. Ed. Engl. 1995, 34, 1865– 1869 DOI: 10.1002/anie.199518651
[Crossref], [CAS], [Google Scholar]

489. Ashton, P. R.; Chrystal, E. J. T.; Glink, P. T.; Menzer, S.; Schiavo, C.; Stoddart, J. F.; Tasker, P. A.; Williams, D. J.
Doubly Encircled and Double-Stranded Pseudorotaxanes Angew. Chem., Int. Ed. Engl. 1995, 34, 1869– 1871
 DOI: 10.1002/anie.199518691 [Crossref], [CAS], [Google Scholar]

490. Ashton, P. R.; Chrystal, E. J. T.; Glink, P. T.; Menzer, S.; Schiavo, C.; Spencer, N.; Stoddart, J. F.; Tasker, P. A.;
White, A. J. P.; Williams, D. J. Pseudorotaxanes Formed Between Secondary Dialkylammonium Salts and Crown
Ethers Chem. - Eur. J. 1996, 2, 709– 728 DOI: 10.1002/chem.19960020616 [Crossref], [CAS], [Google Scholar]

491. Ashton, P. R.; Fyfe, M. C. T.; Glink, P. T.; Menzer, S.; Stoddart, J. F.; White, A. J. P.; Williams, D. J. Molecular
Meccano, Part 24. Multiply Stranded and Multiply Encircled Pseudorotaxanes J. Am. Chem. Soc. 1997, 119,
12514– 12524 DOI: 10.1021/ja9714806 [ACS Full Text ], [CAS], [Google Scholar]

492. Ashton, P. R.; Fyfe, M. C. T.; Martinez-Diaz, M. V.; Menzer, S.; Schiavo, C.; Stoddart, J. F.; White, A. J. P.;
Williams, D. J. Molecular Meccano. Part 39 - Doubly Docked Pseudorotaxanes Chem. - Eur. J. 1998, 4, 1523–
1534 DOI: 10.1002/(SICI)1521-3765(19980807)4:8<1523::AID-CHEM1523>3.0.CO;2-V [Crossref], [CAS], [Google
Scholar]

493. Loeb, S. J.; Tiburcio, J.; Vella, S. J. [2]Pseudorotaxane Formation with N-benzylanilinium Axles and 24-
Crown-8 Ether Wheels Org. Lett. 2005, 7, 4923– 4926 DOI: 10.1021/ol051786e [ACS Full Text ], [CAS], [Google
Scholar]

494. Baroncini, M.; Gao, C.; Carboni, V.; Credi, A.; Previtera, E.; Semeraro, M.; Venturi, M.; Silvi, S. Light Control of
Stoichiometry and Motion in Pseudorotaxanes Comprising a Cucurbit[7]uril Wheel and an Azobenzene-
Bipyridinium Axle Chem. - Eur. J. 2014, 20, 10737– 10744 DOI: 10.1002/chem.201402821 [Crossref], [PubMed],
[CAS], [Google Scholar]

495. Credi, A.; Dumas, S.; Silvi, S.; Venturi, M.; Arduini, A.; Pochini, A.; Secchi, A. Viologen-calix[6]arene
Pseudorotaxanes. Ion-pair Recognition and Threading/dethreading Molecular Motions J. Org. Chem. 2004, 69,
5881– 5887 DOI: 10.1021/jo0494127 [ACS Full Text ], [CAS], [Google Scholar]

496. Balzani, V.; Credi, A.; Marchioni, F.; Stoddart, J. F. Arti cial Molecular-level Machines. Dethreading-
rethreading of a Pseudorotaxane Powered Exclusively by Light Energy Chem. Commun. 2001, 1860– 1861
 DOI: 10.1039/b105160c [Crossref], [PubMed], [CAS], [Google Scholar]

  
497. Baroncini, M. A Simple Molecular Machine Operated by Photoinduced Proton Transfer. Springer Theses;
Springer: New York, 2011; Chapter 7, pp 71– 76;  DOI: 10.1007/978-3-642-19285-2_7 . [Crossref], [Google Scholar]

498. Jeppesen, J. O.; Becher, J.; Stoddart, J. F. Poised on the Brink Between a Bistable Complex and a
Compound Org. Lett. 2002, 4, 557– 560 DOI: 10.1021/ol0171362 [ACS Full Text ], [CAS], [Google Scholar]

499. Rekharsky, M. V.; Yamamura, H.; Kawai, M.; Osaka, I.; Arakawa, R.; Sato, A.; Ko, Y. H.; Selvapalam, N.; Kim, K.;
Inoue, Y. Sequential Formation of a Ternary Complex Among Dihexylammonium, Cucurbit[6]uril, and Cyclodextrin
with Positive Cooperativity Org. Lett. 2006, 8, 815– 818 DOI: 10.1021/ol0528742 [ACS Full Text ],
[CAS], [Google Scholar]

500. Balzani, V.; Credi, A.; Venturi, M. Controlled Disassembling of Self-assembling Systems: Toward Arti cial
Molecular-level Devices and Machines Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 4814– 4817
 DOI: 10.1073/pnas.022631599 [Crossref], [PubMed], [CAS], [Google Scholar]

501. Huang, F.; Fronczek, F. R.; Gibson, H. W. A Cryptand/bisparaquat [3]Pseudorotaxane by Cooperative


Complexation J. Am. Chem. Soc. 2003, 125, 9272– 9273 DOI: 10.1021/ja0355877 [ACS Full Text ],
[CAS], [Google Scholar]

502. Huang, F.; Gibson, H. W.; Bryant, W. S.; Nagvekar, D. S.; Fronczek, F. R. First Pseudorotaxane-like
[3]Complexes based on Cryptands and Paraquat: Self-assembly and Crystal Structures J. Am. Chem. Soc. 2003,
125, 9367– 9371 DOI: 10.1021/ja034968h [ACS Full Text ], [CAS], [Google Scholar]

503. Inoue, Y.; Kanbara, T.; Yamamoto, T. Construction of New [2]Pseudorotaxanes by Hydrogen Bonding
Assembly of Macrocyclic Tetrathiolactam with Amides and an Ester Tetrahedron Lett. 2004, 45, 4603– 4606
 DOI: 10.1016/j.tetlet.2004.04.114 [Crossref], [CAS], [Google Scholar]

504. Sambrook, M. R.; Beer, P. D.; Wisner, J. A.; Paul, R. L.; Cowley, A. R.; Szemes, F.; Drew, M. G. Anion-templated
Assembly of Pseudorotaxanes: Importance of Anion Template, Strength of Ion-pair Thread Association, and
Macrocycle Ring Size J. Am. Chem. Soc. 2005, 127, 2292– 2302 DOI: 10.1021/ja046278z [ACS Full Text ],
[CAS], [Google Scholar]

505. Yonemura, H.; Kusano, S.; Matsuo, T.; Yamada, S. Effect of pi-System on Long-range Photoinduced
Electron Transfer in Through-ring alpha-Cyclodextrin Complexes of Carbazole-viologen Linked Compounds
Tetrahedron Lett. 1998, 39, 6915– 6918 DOI: 10.1016/S0040-4039(98)01451-8 [Crossref], [CAS], [Google
Scholar]   

506. Montalti, M.; Ballardini, R.; Prodi, L.; Balzani, V. Electronic Energy Transfer in Adducts of Aromatic Crown
Ethers with Protonated 9-Methylaminomethylanthracene Chem. Commun. 1996, 2011– 2012
 DOI: 10.1039/cc9960002011 [Crossref], [CAS], [Google Scholar]

507. Asakawa, M.; Ashton, P. R.; Balzani, V.; Credi, A.; Mattersteig, G.; Matthews, O. A.; Montalti, M.; Spencer, N.;
Stoddart, J. F.; Venturi, M. Electrochemically Induced Molecular Motions in Pseudorotaxanes: A Case of Dual-
mode (Oxidative and Reductive) Dethreading Chem. - Eur. J. 1997, 3, 1992– 1996
 DOI: 10.1002/chem.19970031214 [Crossref], [CAS], [Google Scholar]

508. Credi, A.; Balzani, V.; Langford, S. J.; Stoddart, J. F. Logic Operations at the Molecular Level. An XOR Gate
Based on a Molecular Machine J. Am. Chem. Soc. 1997, 119, 2679– 2681 DOI: 10.1021/ja963572l [ACS Full
Text ], [CAS], [Google Scholar]

509. Ashton, P. R.; Ballardini, R.; Balzani, V.; GomezLopez, M.; Lawrence, S. E.; MartinezDiaz, M. V.; Montalti, M.;
Piersanti, A.; Prodi, L.; Stoddart, J. F., et al. Molecular Meccano. 26. Hydrogen-bonded Complexes of Aromatic
Crown Ethers with (9-Anthracenyl)methylammonium Derivatives. Supramolecular Photochemistry and
Photophysics. pH-Controllable Supramolecular Switching J. Am. Chem. Soc. 1997, 119, 10641– 10651
 DOI: 10.1021/ja9715760 [ACS Full Text ], [CAS], [Google Scholar]

510. Devonport, W.; Blower, M. A.; Bryce, M. R.; Goldenberg, L. M. A Redox-active Tetrathiafulvalene
[2]Pseudorotaxane: Spectroelectrochemical and Cyclic Voltammetric Studies of the Highly-reversible
Complexation/decomplexation Process J. Org. Chem. 1997, 62, 885– 887 DOI: 10.1021/jo960951o [ACS Full
Text ], [CAS], [Google Scholar]

511. Montalti, M.; Prodi, L. A Supramolecular Assembly Controlled by Anions: Threading and Unthreading of a
Pseudorotaxane Chem. Commun. 1998, 1461– 1462 DOI: 10.1039/a802157k [Crossref], [CAS], [Google Scholar]

512. Matthews, O. A.; Raymo, F. M.; Stoddart, J. F.; White, A. J. P.; Williams, D. J. Acid/Base-controlled
Supramolecular Switch New J. Chem. 1998, 22, 1131– 1134 DOI: 10.1039/a804742a [Crossref], [CAS], [Google
Scholar]

513. Asakawa, M.; Ashton, P. R.; Balzani, V.; Boyd, S. E.; Credi, A.; Mattersteig, G.; Menzer, S.; Montalti, M.; Raymo,
F. M.; Ru lli, C., et al. Molecular Meccano, 49 - Pseudorotaxanes and Catenanes Containing a Redox-active Unit
Derived from Tetrathiafulvalene Eur. J. Org. Chem. 1999, 1999, 985– 994
 DOI: 10.1002/(SICI)1099-0690(199905)1999:5<985::AID-EJOC985>3.0.CO;2-O [Crossref], [Google  
Scholar] 

514. Balzani, V. V.; Credi, A.; Mattersteig, G.; Matthews, O. A.; Raymo, F. M.; Stoddart, J. F.; Venturi, M.; White, A. J.;
Williams, D. J. Switching of Pseudorotaxanes and Catenanes Incorporating a Tetrathiafulvalene Unit by Redox
and Chemical Inputs J. Org. Chem. 2000, 65, 1924– 1936 DOI: 10.1021/jo991781t [ACS Full Text ],
[CAS], [Google Scholar]

515. Balzani, V. V.; Becher, J.; Credi, A.; Nielsen, M. B.; Raymo, F. M.; Stoddart, J. F.; Talarico, A. M.; Venturi, M. The
Electrochemically-driven Decomplexation/recomplexation of Inclusion Adducts of Ferrocene Derivatives with an
Electron-accepting Receptor J. Org. Chem. 2000, 65, 1947– 1956 DOI: 10.1021/jo991467z [ACS Full Text ],
[CAS], [Google Scholar]

516. Fujimoto, T.; Nakamura, A.; Inoue, Y.; Sakata, Y.; Kaneda, T. Photoswitching of the Association of a
Permethylated alpha-Cyclodextrin-azobenzene Dyad Forming a Janus [2]Pseudorotaxane Tetrahedron Lett.
2001, 42, 7987– 7989 DOI: 10.1016/S0040-4039(01)01563-5 [Crossref], [CAS], [Google Scholar]

517. Kuwabara, J.; Stern, C. L.; Mirkin, C. A. A Coordination Chemistry Approach to a Multieffector Enzyme
Mimic J. Am. Chem. Soc. 2007, 129, 10074– 10075 DOI: 10.1021/ja073447h [ACS Full Text ], [CAS], [Google
Scholar]

518. Zheng, Y.; Yu, Z. Y.; Parker, R. M.; Wu, Y. C.; Abell, C.; Scherman, O. A. Interfacial Assembly of Dendritic
Microcapsules with Host-guest Chemistry Nat. Commun. 2014, 5, 5772 DOI: 10.1038/ncomms6772 [Crossref],
[PubMed], [CAS], [Google Scholar]

519. Horie, M.; Suzaki, Y.; Osakada, K. Formation of Pseudorotaxane Induced by Electrochemical Oxidation of
Ferrocene-containing Axis Molecule in the Presence of Crown Ether J. Am. Chem. Soc. 2004, 126, 3684– 3685
 DOI: 10.1021/ja039899l [ACS Full Text ], [CAS], [Google Scholar]

520. Ballardini, R.; Balzani, V.; Credi, A.; Gandol , M. T.; Langford, S. J.; Menzer, S.; Prodi, L.; Stoddart, J. F.;
Venturi, M.; Williams, D. J. Simple Molecular Machines: Chemically Driven Unthreading and Rethreading of a
[2]Pseudorotaxane Angew. Chem., Int. Ed. Engl. 1996, 35, 978– 981 DOI: 10.1002/anie.199609781 [Crossref],
[CAS], [Google Scholar]

521. Ballardini, R.; Balzani, V.; Gandol , M. T.; Prodi, L.; Venturi, M.; Philp, D.; Ricketts, H. G.; Stoddart, J. F. A
Photochemically Driven Molecular Machine Angew. Chem., Int. Ed. Engl. 1993, 32, 1301– 1303
 DOI: 10.1002/anie.199313011 [Crossref], [Google Scholar]
522. Mock, W. L.; Pierpont, J. A Cucurbituril-Based Molecular Switch J. Chem. Soc., Chem. Commun. 1990,
1509– 1511 DOI: 10.1039/c39900001509 [Crossref], [CAS], [Google Scholar]   

523. Asakawa, M.; Ashton, P. R.; Balzani, V.; Brown, C. L.; Credi, A.; Matthews, O. A.; Newton, S. P.; Raymo, F. M.;
Shipway, A. N.; Spencer, N., et al. Photoactive Azobenzene-containing Supramolecular Complexes and Related
Interlocked Molecular Compounds Chem. - Eur. J. 1999, 5, 860– 875
 DOI: 10.1002/(SICI)1521-3765(19990301)5:3<860::AID-CHEM860>3.0.CO;2-K [Crossref], [CAS], [Google Scholar]

524. Jeppesen, J. O.; Vignon, S. A.; Stoddart, J. F. In the Twilight Zone Between [2]Pseudorotaxanes and
[2]Rotaxanes Chem. - Eur. J. 2003, 9, 4611– 4625 DOI: 10.1002/chem.200304798 [Crossref], [PubMed],
[CAS], [Google Scholar]

525. Jun, S. I.; Lee, J. W.; Sakamoto, S.; Yamaguchi, K.; Kim, K. Rotaxane-based Molecular Switch with
Fluorescence Signaling Tetrahedron Lett. 2000, 41, 471– 475 DOI: 10.1016/S0040-4039(99)02094-8 [Crossref],
[CAS], [Google Scholar]

526. Asakawa, M.; Iqbal, S.; Stoddart, J. F.; Tinker, N. D. Prototype of an Optically Responsive Molecular Switch
Based on Pseudorotaxane Angew. Chem., Int. Ed. Engl. 1996, 35, 976– 978 DOI: 10.1002/anie.199609761
[Crossref], [CAS], [Google Scholar]

527. Ashton, P. R.; Balzani, V.; Kocian, O.; Prodi, L.; Spencer, N.; Stoddart, J. F. A Light-fueled ″Piston Cylinder″
Molecular-level Machine J. Am. Chem. Soc. 1998, 120, 11190– 11191 DOI: 10.1021/ja981889a [ACS Full Text
], [CAS], [Google Scholar]

528. Ishow, E.; Credi, A.; Balzani, V.; Spadola, F.; Mandolini, L. A Molecular-level Plug/socket System: Electronic
Energy Transfer From a Binaphthyl Unit Incorporated into a Crown Ether to an Anthracenyl Unit Linked to an
Ammonium Ion Chem. - Eur. J. 1999, 5, 984– 989
 DOI: 10.1002/(SICI)1521-3765(19990301)5:3<984::AID-CHEM984>3.0.CO;2-T [Crossref], [CAS], [Google Scholar]

529. Ballardini, R.; Balzani, V.; Clemente-Leon, M.; Credi, A.; Gandol , M. T.; Ishow, E.; Perkins, J.; Stoddart, J. F.;
Tseng, H. R.; Wenger, S. Photoinduced Electron Transfer in a Triad That can be Assembled/disassembled by Two
Different External Inputs. Toward Molecular-level Electrical Extension Cables J. Am. Chem. Soc. 2002, 124,
12786– 12795 DOI: 10.1021/ja025813x [ACS Full Text ], [CAS], [Google Scholar]

530. Jeon, W. S.; Ziganshina, A. Y.; Lee, J. W.; Ko, Y. H.; Kang, J. K.; Lee, C.; Kim, K. A [2]Pseudorotaxane-based
Molecular Machine: Reversible Formation of a Molecular Loop Driven by Electrochemical and Photochemical
Stimuli Angew. Chem., Int. Ed. 2003, 42, 4097– 4100 DOI: 10.1002/anie.200351925 [Crossref], [PubMed],
[CAS], [Google Scholar]   

531. Jeon, W. S.; Kim, E.; Ko, Y. H.; Hwang, I.; Lee, J. W.; Kim, S. Y.; Kim, H. J.; Kim, K. Molecular Loop Lock: A
Redox-driven Molecular Machine Based on a Host-stabilized Charge-transfer Complex Angew. Chem., Int. Ed.
2005, 44, 87– 91 DOI: 10.1002/anie.200461806 [Crossref], [CAS], [Google Scholar]

532. Credi, A.; Montalti, M.; Balzani, V.; Langford, S. J.; Raymo, F. M.; Stoddart, J. F. Simple Molecular-level
Machines. Interchange Between Different Threads in Pseudorotaxanes New J. Chem. 1998, 22, 1061– 1065
 DOI: 10.1039/a804787a [Crossref], [CAS], [Google Scholar]

533. Ashton, P. R.; Balzani, V.; Becher, J.; Credi, A.; Fyfe, M. C. T.; Mattersteig, G.; Menzer, S.; Nielsen, M. B.;
Raymo, F. M.; Stoddart, J. F., et al. A Three-pole Supramolecular Switch J. Am. Chem. Soc. 1999, 121, 3951–
3957 DOI: 10.1021/ja984341c [ACS Full Text ], [CAS], [Google Scholar]

534. Chang, K. J.; An, Y. J.; Uh, H.; Jeong, K. S. Reversible Control of Assembly and Disassembly of Interlocked
Supermolecules J. Org. Chem. 2004, 69, 6556– 6563 DOI: 10.1021/jo0490548 [ACS Full Text ], [CAS], [Google
Scholar]

535. Kaiser, G.; Jarrosson, T.; Otto, S.; Ng, Y. F.; Bond, A. D.; Sanders, J. K. M. Lithium-templated Synthesis of a
Donor-acceptor Pseudorotaxane and Catenane Angew. Chem., Int. Ed. 2004, 43, 1959– 1962
 DOI: 10.1002/anie.200353075 [Crossref], [PubMed], [CAS], [Google Scholar]

536. Venturi, M.; Dumas, S.; Balzani, V.; Cao, J. G.; Stoddart, J. F. Threading/dethreading Processes in
Pseudorotaxanes. A Thermodynamic and Kinetic Study New J. Chem. 2004, 28, 1032– 1037
 DOI: 10.1039/b315933g [Crossref], [CAS], [Google Scholar]

537. Nygaard, S.; Laursen, B. W.; Flood, A. H.; Hansen, C. N.; Jeppesen, J. O.; Stoddart, J. F. Quantifying the
Working Stroke of Tetrathiafulvalene-based Electrochemically-driven Linear Motor-molecules Chem. Commun.
2006, 144– 146 DOI: 10.1039/B511575B [Crossref], [PubMed], [CAS], [Google Scholar]

538. Mirzoian, A.; Kaifer, A. E. Reactive Pseudorotaxanes: Inclusion Complexation of Reduced Viologens by the
Hosts beta-Cyclodextrin and Heptakis(2,6-di-O-methyl)-beta-cyclodextrin Chem. - Eur. J. 1997, 3, 1052– 1058
 DOI: 10.1002/chem.19970030711 [Crossref], [CAS], [Google Scholar]

539. Eliadou, K.; Yannakopoulou, K.; Rontoyianni, A.; Mavridis, I. M. NMR Detection of Simultaneous Formation
of [2]- and [3]Pseudorotaxanes in Aqueous Solution Between alpha-Cyclodextrin and Linear Aliphatic
alpha,omega-Amino Acids, an alpha,omega-Diamine and an alpha,omega-Diacid of Similar Length, and

 
Comparison with the Solid-state Structures J. Org. Chem. 1999, 64, 6217– 6226 DOI: 10.1021/jo990021f [ACS
Full Text ], [CAS], [Google Scholar]

540. Huh, K. M.; Ooya, T.; Sasaki, S.; Yui, N. Polymer Inclusion Complex Consisting of Poly(epsilon-lysine) and
alpha-Cyclodextrin Macromolecules 2001, 34, 2402– 2404 DOI: 10.1021/ma0018648 [ACS Full Text ],
[CAS], [Google Scholar]

541. Wisner, J. A.; Beer, P. D.; Drew, M. G. B. A Demonstration of Anion Templation and Selectivity in
Pseudorotaxane Formation Angew. Chem., Int. Ed. 2001, 40, 3606– 3609
 DOI: 10.1002/1521-3773(20011001)40:19<3606::AID-ANIE3606>3.0.CO;2-O [Crossref], [PubMed], [CAS], [Google
Scholar]

542. Wisner, J. A.; Beer, P. D.; Berry, N. G.; Tomapatanaget, B. Anion Recognition as a Method for Templating
Pseudorotaxane Formation Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 4983– 4986 DOI: 10.1073/pnas.062637999
[Crossref], [PubMed], [CAS], [Google Scholar]

543. Anelli, P. L.; Ashton, P. R.; Spencer, N.; Slawin, A. M. Z.; Stoddart, J. F.; Williams, D. J. Self-Assembling
[2]Pseudorotaxanes Angew. Chem., Int. Ed. Engl. 1991, 30, 1036– 1039 DOI: 10.1002/anie.199110361
[Crossref], [Google Scholar]

544. Kraus, T.; Budesinsky, M.; Cvacka, J. C.; Sauvage, J.-P. Copper(I)-directed Formation of a Cyclic
Pseudorotaxane Tetramer and its Trimeric Homologue Angew. Chem., Int. Ed. 2006, 45, 258– 261
 DOI: 10.1002/anie.200503163 [Crossref], [CAS], [Google Scholar]

545. Ashton, P. R.; Baxter, I.; Fyfe, M. C. T.; Raymo, F. M.; Spencer, N.; Stoddart, J. F.; White, A. J. P.; Williams, D. J.
Rotaxane or Pseudorotaxane? That is the Question! J. Am. Chem. Soc. 1998, 120, 2297– 2307
 DOI: 10.1021/ja9731276 [ACS Full Text ], [CAS], [Google Scholar]

546. Collin, J.-P.; Gavina, P.; Sauvage, J.-P. Electrochemically Induced Molecular Motions in a Copper(I) Complex
Pseudorotaxane Chem. Commun. 1996, 2005– 2006 DOI: 10.1039/cc9960002005 [Crossref], [CAS], [Google
Scholar]

547. Wezenberg, S. J.; Vlatkovic, M.; Kistemaker, J. C.; Feringa, B. L. Multi-state Regulation of the Dihydrogen
Phosphate Binding A nity to a Light- and Heat-responsive Bis-urea Receptor J. Am. Chem. Soc. 2014, 136,
16784– 16787 DOI: 10.1021/ja510700j [ACS Full Text ], [CAS], [Google Scholar]
548. Terashima, T.; Mes, T.; De Greef, T. F.; Gillissen, M. A.; Besenius, P.; Palmans, A. R.; Meijer, E. W. Single-chain
Folding of Polymers for Catalytic Systems in Water J. Am. Chem. Soc. 2011, 133, 4742– 4745   
 DOI: 10.1021/ja2004494 [ACS Full Text ], [CAS], [Google Scholar]

549. Giuseppone, N.; Lutz, J. F. Materials Chemistry: Catalytic Accordions Nature 2011, 473, 40– 41
 DOI: 10.1038/473040a [Crossref], [PubMed], [CAS], [Google Scholar]

550. Zayed, J. M.; Nouvel, N.; Rauwald, U.; Scherman, O. A. Chemical Complexity--supramolecular Self-assembly
of Synthetic and Biological Building Blocks in Water Chem. Soc. Rev. 2010, 39, 2806– 2816
 DOI: 10.1039/b922348g [Crossref], [PubMed], [CAS], [Google Scholar]

551. Lee, T. C.; Kalenius, E.; Lazar, A. I.; Assaf, K. I.; Kuhnert, N.; Grun, C. H.; Janis, J.; Scherman, O. A.; Nau, W. M.
Chemistry Inside Molecular Containers in the Gas Phase Nat. Chem. 2013, 5, 376– 382
 DOI: 10.1038/nchem.1618 [Crossref], [PubMed], [CAS], [Google Scholar]

552. Zhang, J.; Coulston, R. J.; Jones, S. T.; Geng, J.; Scherman, O. A.; Abell, C. One-step Fabrication of
Supramolecular Microcapsules from Micro uidic Droplets Science 2012, 335, 690– 694
 DOI: 10.1126/science.1215416 [Crossref], [PubMed], [CAS], [Google Scholar]

553. Goldup, S. Arti cial Molecular Machines: Two Steps Uphill Nat. Nanotechnol. 2015, 10, 488– 489
 DOI: 10.1038/nnano.2015.116 [Crossref], [PubMed], [CAS], [Google Scholar]

554. Nakamura, T.; Takashima, Y.; Hashidzume, A.; Yamaguchi, H.; Harada, A. A Metal-ion-responsive Adhesive
Material via Switching of Molecular Recognition Properties Nat. Commun. 2014, 5, 4622
 DOI: 10.1038/ncomms5622 [Crossref], [PubMed], [CAS], [Google Scholar]

555. Adams, H.; Chekmeneva, E.; Hunter, C. A.; Misuraca, M. C.; Navarro, C.; Turega, S. M. Quanti cation of the
Effect of Conformational Restriction on Supramolecular Effective Molarities J. Am. Chem. Soc. 2013, 135, 1853–
1863 DOI: 10.1021/ja310221t [ACS Full Text ], [CAS], [Google Scholar]

556. Ariga, K.; Ito, H.; Hill, J. P.; Tsukube, H. Molecular Recognition: From Solution Science to Nano/materials
Technology Chem. Soc. Rev. 2012, 41, 5800– 5835 DOI: 10.1039/c2cs35162e [Crossref], [PubMed],
[CAS], [Google Scholar]

557. Rieth, S.; Hermann, K.; Wang, B. Y.; Badjic, J. D. Controlling the Dynamics of Molecular Encapsulation and
Gating Chem. Soc. Rev. 2011, 40, 1609– 1622 DOI: 10.1039/C005254J [Crossref], [PubMed], [CAS], [Google
Scholar]
 E.
558. Witlicki, E. H.; Johnsen, C.; Hansen, S. W.; Silverstein, D. W.; Bottomley, V. J.; Jeppesen, J. O.; Wong, W.; 
Jensen, L.; Flood, A. H. Molecular Logic Gates Using Surface-Enhanced Raman-Scattered Light J. Am. Chem.
Soc. 2011, 133, 7288– 7291 DOI: 10.1021/ja200992x [ACS Full Text ], [CAS], [Google Scholar]

559. Berryman, O. B.; Sather, A. C.; Rebek, J. A Light Controlled Cavitand Wall Regulates Guest Binding Chem.
Commun. 2011, 47, 656– 658 DOI: 10.1039/C0CC03865B [Crossref], [PubMed], [CAS], [Google Scholar]

560. Dube, H.; Ajami, D.; Rebek, J. Photochemical Control of Reversible Encapsulation Angew. Chem., Int. Ed.
2010, 49, 3192– 3195 DOI: 10.1002/anie.201000876 [Crossref], [PubMed], [CAS], [Google Scholar]

561. Rogez, G.; Ribera, B. F.; Credi, A.; Ballardini, R.; Gandol , M. T.; Balzani, V.; Liu, Y.; Northrop, B. H.; Stoddart, J.
F. A Molecular Plug-socket Connector J. Am. Chem. Soc. 2007, 129, 4633– 4642 DOI: 10.1021/ja067739e [ACS
Full Text ], [CAS], [Google Scholar]

562. Arduini, A.; Bussolati, R.; Credi, A.; Secchi, A.; Silvi, S.; Semeraro, M.; Venturi, M. Toward Directionally
Controlled Molecular Motions and Kinetic Intra- and Intermolecular Self-Sorting: Threading Processes of
Nonsymmetric Wheel and Axle Components J. Am. Chem. Soc. 2013, 135, 9924– 9930 DOI: 10.1021/ja404270c
[ACS Full Text ], [CAS], [Google Scholar]

563. Baroncini, M.; Silvi, S.; Venturi, M.; Credi, A. Photoactivated Directionally Controlled Transit of a Non-
Symmetric Molecular Axle Through a Macrocycle Angew. Chem., Int. Ed. 2012, 51, 4223– 4226
 DOI: 10.1002/anie.201200555 [Crossref], [PubMed], [CAS], [Google Scholar]

564. Cao, D.; Amelia, M.; Klivansky, L. M.; Koshkakaryan, G.; Khan, S. I.; Semeraro, M.; Silvi, S.; Venturi, M.; Credi,
A.; Liu, Y. Probing Donor-Acceptor Interactions and Co-Conformational Changes in Redox Active Desymmetrized
[2]Catenanes J. Am. Chem. Soc. 2010, 132, 1110– 1122 DOI: 10.1021/ja909041g [ACS Full Text ],
[CAS], [Google Scholar]

565. Leblond, J.; Petitjean, A. Molecular Tweezers: Concepts and Applications ChemPhysChem 2011, 12,
1043– 1051 DOI: 10.1002/cphc.201001050 [Crossref], [PubMed], [CAS], [Google Scholar]

566. Legouin, B.; Gayral, M.; Uriac, P.; Cupif, J. F.; Levoin, N.; Toupet, L.; van de Weghe, P. Molecular Tweezers:
Synthesis and Formation of Host-Guest Complexes Eur. J. Org. Chem. 2010, 2010, 5503– 5508
 DOI: 10.1002/ejoc.201000729 [Crossref], [Google Scholar]
567. Bier, D.; Rose, R.; Bravo-Rodriguez, K.; Bartel, M.; Ramirez-Anguita, J. M.; Dutt, S.; Wilch, C.; Klarner, F. G.;
Sanchez-Garcia, E.; Schrader, T.; Ottmann, C. Molecular Tweezers Modulate 14–3-3 Protein-protein  
Interactions
Nat. Chem. 2013, 5, 234– 239 DOI: 10.1038/nchem.1570 [Crossref], [PubMed], [CAS], [Google Scholar]

568. Colquhoun, H. M.; Zhu, Z. X. Recognition of Polyimide Sequence Information by a Molecular Tweezer
Angew. Chem., Int. Ed. 2004, 43, 5040– 5045 DOI: 10.1002/anie.200460382 [Crossref], [PubMed], [CAS], [Google
Scholar]

569. Gianneschi, N. C.; Cho, S. H.; Nguyen, S. T.; Mirkin, C. A. Reversibly Addressing an Allosteric Catalyst in situ:
Catalytic Molecular Tweezers Angew. Chem., Int. Ed. 2004, 43, 5503– 5507 DOI: 10.1002/anie.200460932
[Crossref], [PubMed], [CAS], [Google Scholar]

570. Guther, R.; Nieger, M.; Vogtle, F. Molecular Tweezers with a Hydrocarbon Skeleton and Convergent Carboxyl
Groups Angew. Chem., Int. Ed. Engl. 1993, 32, 601– 603 DOI: 10.1002/anie.199306011 [Crossref], [Google
Scholar]

571. Hardouin-Lerouge, M.; Hudhomme, P.; Salle, M. Molecular Clips and Tweezers Hosting Neutral Guests
Chem. Soc. Rev. 2011, 40, 30– 43 DOI: 10.1039/B915145C [Crossref], [PubMed], [CAS], [Google Scholar]

572. Klarner, F. G.; Schrader, T. Aromatic Interactions by Molecular Tweezers and Clips in Chemical and
Biological Systems Acc. Chem. Res. 2013, 46, 967– 978 DOI: 10.1021/ar300061c [ACS Full Text ],
[CAS], [Google Scholar]

573. Leblond, J.; Gao, H.; Petitjean, A.; Leroux, J. C. pH-Responsive Molecular Tweezers J. Am. Chem. Soc.
2010, 132, 8544– 8545 DOI: 10.1021/ja103153t [ACS Full Text ], [CAS], [Google Scholar]

574. Zimmerman, S. C. Rigid Molecular Tweezers as Hosts for the Complexation of Neutral Guests Top. Curr.
Chem. 1993, 165, 71– 102 [Crossref], [CAS], [Google Scholar]

575. Shinkai, S.; Nakaji, T.; Ogawa, T.; Shigematsu, K.; Manabe, O. Photoresponsive Crown Ethers 0.2.
Photocontrol of Ion Extraction and Ion-Transport by a Bis(Crown Ether) with a Butter y-Like Motion J. Am. Chem.
Soc. 1981, 103, 111– 115 DOI: 10.1021/ja00391a021 [ACS Full Text ], [CAS], [Google Scholar]

576. Shinkai, S.; Ishihara, M.; Ueda, K.; Manabe, O. Photoresponsive Crown Ethers 0.14. Photoregulated Crown
Metal Complexation by Competitive Intramolecular Tail(Ammonium)-Biting J. Chem. Soc., Perkin Trans. 2 1985,
511– 518 DOI: 10.1039/p29850000511 [Crossref], [CAS], [Google Scholar]
577. Perez, E. M.; Sanchez, L.; Fernandez, G.; Martin, N. exTTF as a Building Block for Fullerene Receptors.
Unexpected Solvent-dependent Positive Homotropic Cooperativity J. Am. Chem. Soc. 2006, 128, 7172– 
 7173 
 DOI: 10.1021/ja0621389 [ACS Full Text ], [CAS], [Google Scholar]

578. Zimmerman, S. C.; Vanzyl, C. M.; Hamilton, G. S. Rigid Molecular Tweezers - Preorganized Hosts for
Electron-Donor Acceptor Complexation in Organic-Solvents J. Am. Chem. Soc. 1989, 111, 1373– 1381
 DOI: 10.1021/ja00186a035 [ACS Full Text ], [CAS], [Google Scholar]

579. Molt, O.; Rubeling, D.; Schrader, T. A Selective Biomimetic Tweezer for Noradrenaline J. Am. Chem. Soc.
2003, 125, 12086– 12087 DOI: 10.1021/ja035212l [ACS Full Text ], [CAS], [Google Scholar]

580. Etxebarria, J.; Vidal-Ferran, A.; Ballester, P. The Effect of Complex Stoichiometry in Supramolecular Chirality
Transfer to Zinc Bisporphyrin Systems Chem. Commun. 2008, 5939– 5941 DOI: 10.1039/b812819g [Crossref],
[PubMed], [CAS], [Google Scholar]

581. Ulrich, S.; Petitjean, A.; Lehn, J.-M. Metallo-Controlled Dynamic Molecular Tweezers: Design, Synthesis, and
Self-Assembly by Metal-Ion Coordination Eur. J. Inorg. Chem. 2010, 2010, 1913– 1928
 DOI: 10.1002/ejic.200901262 [Crossref], [Google Scholar]

582. Petitjean, A.; Khoury, R. G.; Kyritsakas, N.; Lehn, J.-M. Dynamic Devices. Shape Switching and Substrate
Binding in Ion-controlled Nanomechanical Molecular Tweezers J. Am. Chem. Soc. 2004, 126, 6637– 6647
 DOI: 10.1021/ja031915r [ACS Full Text ], [CAS], [Google Scholar]

583. Petitjean, A.; Kyritsakas, N.; Lehn, J.-M. Ion-triggered Multistate Molecular Switching Device Based on
Regioselective Coordination-controlled Ion Binding Chem. - Eur. J. 2005, 11, 6818– 6828
 DOI: 10.1002/chem.200500625 [Crossref], [PubMed], [CAS], [Google Scholar]

584. Linke-Schaetzel, M.; Anson, C. E.; Powell, A. K.; Buth, G.; Palomares, E.; Durrant, J. D.; Balaban, T. S.; Lehn, J.-
M. Dynamic Chemical Devices: Photoinduced Electron Transfer and its Ion-triggered Switching in
Nanomechanical Butter y-type Bis(porphyrin)terpyridines Chem. - Eur. J. 2006, 12, 1931– 1940
 DOI: 10.1002/chem.200500602 [Crossref], [PubMed], [CAS], [Google Scholar]

585. Oliveri, C. G.; Ulmann, P. A.; Wiester, M. J.; Mirkin, C. A. Heteroligated Supramolecular Coordination
Complexes Formed via the Halide-Induced Ligand Rearrangement Reaction Acc. Chem. Res. 2008, 41, 1618–
1629 DOI: 10.1021/ar800025w [ACS Full Text ], [CAS], [Google Scholar]
586. Oliveri, C. G.; Nguyen, S. T.; Mirkin, C. A. A Highly Modular and Convergent Approach for the Synthesis of
Stimulant-responsive Heteroligated Cofacial Porphyrin Tweezer Complexes Inorg. Chem. 2008, 47, 2763
2755–
 DOI: 10.1021/ic702150y [ACS Full Text ], [CAS], [Google Scholar]

587. Skibinski, M.; Gomez, R.; Lork, E.; Azov, V. A. Redox Responsive Molecular Tweezers with Tetrathiafulvalene
Units: Synthesis, Electrochemistry, and Binding Properties Tetrahedron 2009, 65, 10348– 10354
 DOI: 10.1016/j.tet.2009.10.052 [Crossref], [CAS], [Google Scholar]

588. Tian, F.; Jiao, D. Z.; Biedermann, F.; Scherman, O. A. Orthogonal Switching of a Single Supramolecular
Complex Nat. Commun. 2012, 3, 1207 DOI: 10.1038/ncomms2198 [Crossref], [PubMed], [CAS], [Google Scholar]

589. Miwa, K.; Furusho, Y.; Yashima, E. Ion-triggered Spring-like Motion of a Double helicate Accompanied by
Anisotropic Twisting Nat. Chem. 2010, 2, 444– 449 DOI: 10.1038/nchem.649 [Crossref], [PubMed],
[CAS], [Google Scholar]

590. Yamamoto, S.; Iida, H.; Yashima, E. Guest-Induced Unidirectional Dual Rotary and Twisting Motions of a
Spiroborate-Based Double-Stranded Helicate Containing a Bisporphyrin Unit Angew. Chem., Int. Ed. 2013, 52,
6849– 6853 DOI: 10.1002/anie.201302560 [Crossref], [PubMed], [CAS], [Google Scholar]

591. Ferrand, Y.; Gan, Q.; Kauffmann, B.; Jiang, H.; Huc, I. Template-Induced Screw Motions within an Aromatic
Amide Foldamer Double Helix Angew. Chem., Int. Ed. 2011, 50, 7572– 7575 DOI: 10.1002/anie.201101697
[Crossref], [PubMed], [CAS], [Google Scholar]

592. Hua, Y.; Liu, Y.; Chen, C. H.; Flood, A. H. Hydrophobic Collapse of Foldamer Capsules Drives Picomolar-level
Chloride Binding in Aqueous Acetonitrile Solutions J. Am. Chem. Soc. 2013, 135, 14401– 14412
 DOI: 10.1021/ja4074744 [ACS Full Text ], [CAS], [Google Scholar]

593. Juwarker, H.; Jeong, K. S. Anion-controlled Foldamers Chem. Soc. Rev. 2010, 39, 3664– 3674
 DOI: 10.1039/b926162c [Crossref], [PubMed], [CAS], [Google Scholar]

594. Gan, Q.; Ronson, T. K.; Vosburg, D. A.; Thoburn, J. D.; Nitschke, J. R. Cooperative Loading and Release
Behavior of a Metal-Organic Receptor J. Am. Chem. Soc. 2015, 137, 1770– 1773 DOI: 10.1021/ja5120437 [ACS
Full Text ], [CAS], [Google Scholar]

595. Hua, Y. R.; Flood, A. H. Flipping the Switch on Chloride Concentrations with a Light-Active Foldamer J. Am.
Chem. Soc. 2010, 132, 12838– 12840 DOI: 10.1021/ja105793c [ACS Full Text ], [CAS], [Google Scholar]
596. Li, Y.; Flood, A. H.; Pure, C.-H. Hydrogen Bonding to Chloride Ions: A Preorganized and Rigid Macrocyclic
 
Receptor Angew. Chem., Int. Ed. 2008, 47, 2649– 2652 DOI: 10.1002/anie.200704717 [Crossref], [PubMed], 
[CAS], [Google Scholar]

597. Li, Y.; Flood, A. H. Strong, Size-selective, and Electronically Tunable C-H···Halide Binding with Steric Control
over Aggregation From Synthetically Modular, Shape-persistent [34]Triazolophanes J. Am. Chem. Soc. 2008, 130,
12111– 12122 DOI: 10.1021/ja803341y [ACS Full Text ], [CAS], [Google Scholar]

598. Hua, Y.; Flood, A. H. Click Chemistry Generates Privileged CH Hydrogen-bonding Triazoles: The Latest
Addition to Anion Supramolecular Chemistry Chem. Soc. Rev. 2010, 39, 1262– 1271 DOI: 10.1039/b818033b
[Crossref], [PubMed], [CAS], [Google Scholar]

599. Juwarker, H.; Lenhardt, J. M.; Pham, D. M.; Craig, S. L. 1,2,3-Triazole CH···Cl– Contacts Guide Anion Binding
and Concomitant Folding in 1,4-Diaryl Triazole Oligomers Angew. Chem., Int. Ed. 2008, 47, 3740– 3743
 DOI: 10.1002/anie.200800548 [Crossref], [PubMed], [CAS], [Google Scholar]

600. Meudtner, R. M.; Hecht, S. Helicity Inversion in Responsive Foldamers Induced by Achiral Halide Ion Guests
Angew. Chem., Int. Ed. 2008, 47, 4926– 4930 DOI: 10.1002/anie.200800796 [Crossref], [PubMed], [CAS], [Google
Scholar]

601. Lee, S.; Hua, Y. R.; Flood, A. H. beta-Sheet-like Hydrogen Bonds Interlock the Helical Turns of a
Photoswitchable Foldamer To Enhance the Binding and Release of Chloride J. Org. Chem. 2014, 79, 8383– 8396
 DOI: 10.1021/jo501595k [ACS Full Text ], [CAS], [Google Scholar]

602. Lee, S.; Flood, A. H. Photoresponsive Receptors for Binding and Releasing Anions J. Phys. Org. Chem.
2013, 26, 79– 86 DOI: 10.1002/poc.2973 [Crossref], [CAS], [Google Scholar]

603. Suk, J.-m.; Naidu, V. R.; Liu, X.; Lah, M. S.; Jeong, K.-S. A Foldamer-Based Chiroptical Molecular Switch That
Displays Complete Inversion of the Helical Sense upon Anion Binding J. Am. Chem. Soc. 2011, 133, 13938–
13941 DOI: 10.1021/ja206546b [ACS Full Text ], [CAS], [Google Scholar]

604. Campbell, V. E.; de Hatten, X.; Delsuc, N.; Kauffmann, B.; Huc, I.; Nitschke, J. R. Cascading Transformations
within a Dynamic Self-assembled System Nat. Chem. 2010, 2, 684– 687 DOI: 10.1038/nchem.693 [Crossref],
[PubMed], [CAS], [Google Scholar]

605. Lanyi, J. K. Bacteriorhodopsin as a Model for Proton Pumps Nature 1995, 375, 461– 463
 DOI: 10.1038/375461a0 [Crossref], [PubMed], [CAS], [Google Scholar]

606. Amendola, V.; Fabbrizzi, L.; Mangano, C.; Pallavicini, P. Molecular Machines Based on Metal Ion  
Translocation Acc. Chem. Res. 2001, 34, 488– 493 DOI: 10.1021/ar010011c [ACS Full Text ], [CAS], [Google
Scholar]

607. Zhao, J. Z.; Ji, S. M.; Chen, Y. H.; Guo, H. M.; Yang, P. Excited State Intramolecular Proton Transfer (ESIPT):
From Principal Photophysics to the Development of New Chromophores and Applications in Fluorescent
Molecular Probes and Luminescent Materials Phys. Chem. Chem. Phys. 2012, 14, 8803– 8817
 DOI: 10.1039/C2CP23144A [Crossref], [PubMed], [CAS], [Google Scholar]

608. Iijima, T.; Momotake, A.; Shinohara, Y.; Sato, T.; Nishimura, Y.; Arai, T. Excited-State Intramolecular Proton
Transfer of Naphthalene-Fused 2-(2′-Hydroxyaryl)benzazole Family J. Phys. Chem. A 2010, 114, 1603– 1609
 DOI: 10.1021/jp904370t [ACS Full Text ], [CAS], [Google Scholar]

609. Fabbrizzi, L.; Gatti, F.; Pallavicini, P.; Zambarbieri, E. Redox-driven Intramolecular Anion Translocation
Between Transition Metal Centers Chem. - Eur. J. 1999, 5, 682– 690
 DOI: 10.1002/(SICI)1521-3765(19990201)5:2<682::AID-CHEM682>3.0.CO;2-E [Crossref], [CAS], [Google Scholar]

610. Yoo, H. J.; Mirkin, C. A.; DiPasquale, A. G.; Rheingold, A. L.; Stern, C. L. Reversible CO-Induced Chloride
Shuttling in Rh-I Tweezer Complexes Containing Urea-Functionalized Hemilabile Ligands Inorg. Chem. 2008, 47,
9727– 9729 DOI: 10.1021/ic8008909 [ACS Full Text ], [CAS], [Google Scholar]

611. Zelikovich, L.; Libman, J.; Shanzer, A. Molecular Redox Switches Based on Chemical Triggering of Iron
Translocation in Triple-Stranded Helical Complexes Nature 1995, 374, 790– 792 DOI: 10.1038/374790a0
[Crossref], [CAS], [Google Scholar]

612. Amendola, V.; Fabbrizzi, L.; Mangano, C.; Miller, H.; Pallavicini, P.; Perotti, A.; Taglietti, A. Signal Ampli cation
by a Fluorescent Indicator of a pH-Driven Intramolecular Translocation of a Copper(II) Ion Angew. Chem., Int. Ed.
2002, 41, 2553– 2356 DOI: 10.1002/1521-3773(20020715)41:14<2553::AID-ANIE2553>3.0.CO;2-U [Crossref],
[PubMed], [CAS], [Google Scholar]

613. Amendola, V.; Fabbrizzi, L.; Mangano, C.; Pallavicini, P.; Perotti, A.; Taglietti, A. pH-Controlled Translocation
of NiII Within a Ditopic Receptor Bearing an Appended Anthracene Fragment: A Mechanical Switch of
Fluorescence J. Chem. Soc. Dalton 2000, 185– 189 DOI: 10.1039/a907966a [Crossref], [Google Scholar]

614. Colasson, B.; Le Poul, N.; Le Mest, Y.; Reinaud, O. Electrochemically Triggered Double Translocation of Two
Different Metal Ions with a Ditopic Calix[6]arene Ligand J. Am. Chem. Soc. 2010, 132, 4393– 4398
 DOI: 10.1021/ja910676z [ACS Full Text ], [CAS], [Google Scholar]

  
615. Dotz, K. H.; Jahr, H. C. Tunable Haptotropic Metal Migration in Fused Arenes: Towards Organometallic
Switches Chem. Rec. 2004, 4, 61– 71 DOI: 10.1002/tcr.20007 [Crossref], [PubMed], [CAS], [Google Scholar]

616. Murahashi, T.; Shirato, K.; Fukushima, A.; Takase, K.; Suenobu, T.; Fukuzumi, S.; Ogoshi, S.; Kurosawa, H.
Redox-induced Reversible Metal Assembly Through Translocation and Reversible Ligand Coupling in
Tetranuclear Metal Sandwich Frameworks Nat. Chem. 2012, 4, 52– 58 DOI: 10.1038/nchem.1202 [Crossref],
[CAS], [Google Scholar]

617. Wozny, M.; Pawlowska, J.; Osior, A.; Swider, P.; Bilewicz, R.; Korybut-Daszkiewicz, B. An Electrochemically
Switchable Foldamer - A Surprising Feature of a Rotaxane with Equivalent Stations Chem. Sci. 2014, 5, 2836–
2842 DOI: 10.1039/c4sc00449c [Crossref], [CAS], [Google Scholar]

618. Chao, S.; Romuald, C.; Fournel-Marotte, K.; Clavel, C.; Coutrot, F. A Strategy Utilizing a Recyclable
Macrocycle Transporter for the E cient Synthesis of a Triazolium-Based [2]Rotaxane Angew. Chem., Int. Ed.
2014, 53, 6914– 6919 DOI: 10.1002/anie.201403765 [Crossref], [PubMed], [CAS], [Google Scholar]

619. Langton, M. J.; Blackburn, O. A.; Lang, T.; Faulkner, S.; Beer, P. D. Nitrite-Templated Synthesis of Lanthanide-
Containing [2]Rotaxanes for Anion Sensing Angew. Chem., Int. Ed. 2014, 53, 11463– 11466
 DOI: 10.1002/anie.201405131 [Crossref], [PubMed], [CAS], [Google Scholar]

620. Nygaard, S.; Laursen, B. W.; Hansen, T. S.; Bond, A. D.; Flood, A. H.; Jeppesen, J. O. Preparation of
Cyclobis(paraquat-p-phenylene)-Based [2]Rotaxanes Without Flexible Glycol Chains Angew. Chem., Int. Ed. 2007,
46, 6093– 6097 DOI: 10.1002/anie.200701722 [Crossref], [PubMed], [CAS], [Google Scholar]

621. Klivansky, L. M.; Koshkakaryan, G.; Cao, D.; Liu, Y. Linear π-Acceptor-Templated Dynamic Clipping to
Macrobicycles and [2]Rotaxanes Angew. Chem., Int. Ed. 2009, 48, 4185– 4189 DOI: 10.1002/anie.200900716
[Crossref], [PubMed], [CAS], [Google Scholar]

622. Prikhod'ko, A. I.; Sauvage, J.-P. Passing Two Strings through the Same Ring Using an Octahedral Metal
Center as Template: A New Synthesis of [3]Rotaxanes J. Am. Chem. Soc. 2009, 131, 6794– 6807
 DOI: 10.1021/ja809267z [ACS Full Text ], [CAS], [Google Scholar]

623. Goldup, S. M.; Leigh, D. A.; Lusby, P. J.; McBurney, R. T.; Slawin, A. M. Z. Active Template Synthesis of
Rotaxanes and Molecular Shuttles with Switchable Dynamics by Four-Component PdII-Promoted Michael
Additions Angew. Chem., Int. Ed. 2008, 47, 3381– 3384 DOI: 10.1002/anie.200705859 [Crossref], [PubMed],
[CAS], [Google Scholar]   

624. Ikeda, T.; Higuchi, M.; Kurth, D. G. From Thiophene [2]Rotaxane to Polythiophene Polyrotaxane J. Am.
Chem. Soc. 2009, 131, 9158– 9159 DOI: 10.1021/ja902992c [ACS Full Text ], [CAS], [Google Scholar]

625. Collin, J.-P.; Durola, F.; Frey, J.; Heitz, V.; Reviriego, F.; Sauvage, J.-P.; Trolez, Y.; Rissanen, K. Templated
Synthesis of Cyclic [4]Rotaxanes Consisting of Two Stiff Rods Threaded through Two Bis-macrocycles with a
Large and Rigid Central Plate as Spacer J. Am. Chem. Soc. 2010, 132, 6840– 6850 DOI: 10.1021/ja101759w
[ACS Full Text ], [CAS], [Google Scholar]

626. Xue, Z.; Mayer, M. F. Actuator Prototype: Capture and Release of a Self-Entangled [1]Rotaxane J. Am.
Chem. Soc. 2010, 132, 3274– 3276 DOI: 10.1021/ja9077655 [ACS Full Text ], [CAS], [Google Scholar]

627. Altieri, A.; Aucagne, V.; Carrillo, R.; Clarkson, G. J.; D’Souza, D. M.; Dunnett, J. A.; Leigh, D. A.; Mullen, K. M.
Sulfur-containing Amide-based [2]Rotaxanes and Molecular Shuttles Chem. Sci. 2011, 2, 1922– 1928
 DOI: 10.1039/c1sc00335f [Crossref], [CAS], [Google Scholar]

628. De Bo, G.; De Winter, J.; Gerbaux, P.; Fustin, C.-A. Rotaxane-Based Mechanically Linked Block Copolymers
Angew. Chem., Int. Ed. 2011, 50, 9093– 9096 DOI: 10.1002/anie.201103716 [Crossref], [PubMed], [CAS], [Google
Scholar]

629. Zhang, Z.-J.; Zhang, H.-Y.; Wang, H.; Liu, Y. A Twin-Axial Hetero[7]rotaxane Angew. Chem., Int. Ed. 2011, 50,
10834– 10838 DOI: 10.1002/anie.201105375 [Crossref], [PubMed], [CAS], [Google Scholar]

630. Li, H.; Fahrenbach, A. C.; Coskun, A.; Zhu, Z.; Barin, G.; Zhao, Y.-L.; Botros, Y. Y.; Sauvage, J.-P.; Stoddart, J. F.
A Light-Stimulated Molecular Switch Driven by Radical–Radical Interactions in Water Angew. Chem., Int. Ed.
2011, 50, 6782– 6788 DOI: 10.1002/anie.201102510 [Crossref], [PubMed], [CAS], [Google Scholar]

631. Nepogodiev, S. A.; Stoddart, J. F. Cyclodextrin-Based Catenanes and Rotaxanes Chem. Rev. 1998, 98,
1959– 1976 DOI: 10.1021/cr970049w [ACS Full Text ], [CAS], [Google Scholar]

632. Dietrich-Buchecker, C. O.; Sauvage, J.-P. Interlocking of Molecular Threads: from the Statistical Approach to
the Templated Synthesis of Catenands Chem. Rev. 1987, 87, 795– 810 DOI: 10.1021/cr00080a007 [ACS Full
Text ], [CAS], [Google Scholar]
633. Anderson, S.; Anderson, H. L.; Sanders, J. K. M. Expanding Roles for Templates in Synthesis Acc. Chem.
Res. 1993, 26, 469– 475 DOI: 10.1021/ar00033a003 [ACS Full Text ], [CAS], [Google Scholar]   

634. Fuller, A.-M.; Leigh, D. A.; Lusby, P. J.; Oswald, I. D. H.; Parsons, S.; Walker, D. B. A 3D Interlocked Structure
from a 2D Template: Structural Requirements for the Assembly of a Square-Planar Metal-Coordinated
[2]Rotaxane Angew. Chem., Int. Ed. 2004, 43, 3914– 3918 DOI: 10.1002/anie.200353622 [Crossref], [PubMed],
[CAS], [Google Scholar]

635. Cantrill, S. J.; Chichak, K. S.; Peters, A. J.; Stoddart, J. F. Nanoscale Borromean Rings Acc. Chem. Res.
2005, 38, 1– 9 DOI: 10.1021/ar040226x [ACS Full Text ], [CAS], [Google Scholar]

636. Chambron, J.-C.; Collin, J.-P.; Heitz, V.; Jouvenot, D.; Kern, J.-M.; Mobian, P.; Pomeranc, D.; Sauvage, J.-P.
Rotaxanes and Catenanes Built Around Octahedral Transition Metals Eur. J. Org. Chem. 2004, 2004, 1627– 1638
 DOI: 10.1002/ejoc.200300341 [Crossref], [Google Scholar]

637. Dietrich-Buchecker, C.; Rapenne, G.; Sauvage, J.-P. Synthesis of Catenanes and Molecular Knots by
Copper(I)-directed Formation of the Precursors Followed by Ruthenium(II)-Catalysed Ring-closing Metathesis
Coord. Chem. Rev. 1999, 185–186, 167– 176 DOI: 10.1016/S0010-8545(98)00266-5 [Crossref], [CAS], [Google
Scholar]

638. Fujita, M. Self-Assembly of [2]Catenanes Containing Metals in Their Backbones Acc. Chem. Res. 1999, 32,
53– 61 DOI: 10.1021/ar9701068 [ACS Full Text ], [CAS], [Google Scholar]

639. Fujita, M.; Ogura, K. Transition-metal-directed Assembly of Well-de ned Organic Architectures Possessing
Large Voids: From Macrocycles to [2] Catenanes Coord. Chem. Rev. 1996, 148, 249– 264
 DOI: 10.1016/0010-8545(95)01212-5 [Crossref], [CAS], [Google Scholar]

640. Busch, D. H.; Stephenson, N. A. Molecular Organization, Portal to Supramolecular Chemistry: Structural
Analysis of the Factors Associated with Molecular Organization in Coordination and Inclusion Chemistry,
Including the Coordination Template Effect Coord. Chem. Rev. 1990, 100, 119– 154
 DOI: 10.1016/0010-8545(90)85007-F [Crossref], [CAS], [Google Scholar]

641. Sauvage, J.-P. Interlacing Molecular Threads on Transition Metals: Catenands, Catenates, and Knots Acc.
Chem. Res. 1990, 23, 319– 327 DOI: 10.1021/ar00178a001 [ACS Full Text ], [CAS], [Google Scholar]

642. Hubin, T. J.; Busch, D. H. Template Routes to Interlocked Molecular Structures and Orderly Molecular
Entanglements Coord. Chem. Rev. 2000, 200–202, 5– 52 DOI: 10.1016/S0010-8545(99)00242-8 [Crossref],
[CAS], [Google Scholar]

  
643. Hoss, R.; Vögtle, F. Template Syntheses Angew. Chem., Int. Ed. Engl. 1994, 33, 375– 384
 DOI: 10.1002/anie.199403751 [Crossref], [Google Scholar]

644. Claessens, C. G.; Stoddart, J. F. π–π Interactions in Self-Assembly J. Phys. Org. Chem. 1997, 10, 254– 272
 DOI: 10.1002/(SICI)1099-1395(199705)10:5<254::AID-POC875>3.0.CO;2-3 [Crossref], [CAS], [Google Scholar]

645. Gillard, R. E.; Raymo, F. M.; Stoddart, J. F. Controlling Self-Assembly Chem. - Eur. J. 1997, 3, 1933– 1940
 DOI: 10.1002/chem.19970031208 [Crossref], [CAS], [Google Scholar]

646. Philp, D.; Stoddart, J. F. Self-Assembly in Organic Synthesis Synlett 1991, 1991, 445– 458
 DOI: 10.1055/s-1991-20759 [Crossref], [Google Scholar]

647. Hogg, L.; Leigh, D. A.; Lusby, P. J.; Morelli, A.; Parsons, S.; Wong, J. K. Y. A Simple General Ligand System for
Assembling Octahedral Metal–Rotaxane Complexes Angew. Chem., Int. Ed. 2004, 43, 1218– 1221
 DOI: 10.1002/anie.200353186 [Crossref], [PubMed], [CAS], [Google Scholar]

648. Hannam, J. S.; Kidd, T. J.; Leigh, D. A.; Wilson, A. J. Magic Rod” Rotaxanes: The Hydrogen Bond-Directed
Synthesis of Molecular Shuttles under Thermodynamic Control Org. Lett. 2003, 5, 1907– 1910
 DOI: 10.1021/ol0344927 [ACS Full Text ], [CAS], [Google Scholar]

649. Leigh, D. A.; Lusby, P. J.; Teat, S. J.; Wilson, A. J.; Wong, J. K. Y. Benzylic Imine Catenates: Readily
Accessible Octahedral Analogues of the Sauvage Catenates Angew. Chem., Int. Ed. 2001, 40, 1538– 1543
 DOI: 10.1002/1521-3773(20010417)40:8<1538::AID-ANIE1538>3.0.CO;2-F [Crossref], [PubMed], [CAS], [Google
Scholar]

650. Kidd, T. J.; Leigh, D. A.; Wilson, A. J. Organic “Magic Rings”: The Hydrogen Bond-Directed Assembly of
Catenanes under Thermodynamic Control J. Am. Chem. Soc. 1999, 121, 1599– 1600 DOI: 10.1021/ja983106r
[ACS Full Text ], [CAS], [Google Scholar]

651. Seel, C.; Vögtle, F. Templates, “Wheeled Reagents”, and a New Route to Rotaxanes by Anion Complexation:
The Trapping Method Chem. - Eur. J. 2000, 6, 21– 24
 DOI: 10.1002/(SICI)1521-3765(20000103)6:1<21::AID-CHEM21>3.0.CO;2-L [Crossref], [PubMed], [CAS], [Google
Scholar]
652. Aucagne, V.; Leigh, D. A.; Lock, J. S.; Thomson, A. R. Rotaxanes of Cyclic Peptides J. Am. Chem. Soc. 2006,
128, 1784– 1785 DOI: 10.1021/ja057206q [ACS Full Text ], [CAS], [Google Scholar]   

653. Biscarini, F.; Cavallini, M.; Leigh, D. A.; León, S.; Teat, S. J.; Wong, J. K. Y.; Zerbetto, F. The Effect of
Mechanical Interlocking on Crystal Packing: Predictions and Testing J. Am. Chem. Soc. 2002, 124, 225– 233
 DOI: 10.1021/ja0159362 [ACS Full Text ], [CAS], [Google Scholar]

654. Johnston, A. G.; Leigh, D. A.; Nezhat, L.; Smart, J. P.; Deegan, M. D. Structurally Diverse and Dynamically
Versatile Benzylic Amide [2]Catenanes Assembled Directly from Commercially Available Precursors Angew.
Chem., Int. Ed. Engl. 1995, 34, 1212– 1216 DOI: 10.1002/anie.199512121 [Crossref], [CAS], [Google Scholar]

655. Johnston, A. G.; Leigh, D. A.; Pritchard, R. J.; Deegan, M. D. Facile Synthesis and Solid-State Structure of a
Benzylic Amide [2]Catenane Angew. Chem., Int. Ed. Engl. 1995, 34, 1209– 1212 DOI: 10.1002/anie.199512091
[Crossref], [CAS], [Google Scholar]

656. Jäger, R.; Vögtle, F. A New Synthetic Strategy towards Molecules with Mechanical Bonds: Nonionic
Template Synthesis of Amide-Linked Catenanes and Rotaxanes Angew. Chem., Int. Ed. Engl. 1997, 36, 930– 944
 DOI: 10.1002/anie.199709301 [Crossref], [CAS], [Google Scholar]

657. Aucagne, V.; Hänni, K. D.; Leigh, D. A.; Lusby, P. J.; Walker, D. B. Catalytic “Click” Rotaxanes: A
Substoichiometric Metal-Template Pathway to Mechanically Interlocked Architectures J. Am. Chem. Soc. 2006,
128, 2186– 2187 DOI: 10.1021/ja056903f [ACS Full Text ], [CAS], [Google Scholar]

658. Crowley, J. D.; Hänni, K. D.; Lee, A.-L.; Leigh, D. A. [2]Rotaxanes through Palladium Active-Template
Oxidative Heck Cross-Couplings J. Am. Chem. Soc. 2007, 129, 12092– 12093 DOI: 10.1021/ja075219t [ACS Full
Text ], [CAS], [Google Scholar]

659. Crowley, J. D.; Hänni, K. D.; Leigh, D. A.; Slawin, A. M. Z. Diels–Alder Active-Template Synthesis of
Rotaxanes and Metal-Ion-Switchable Molecular Shuttles J. Am. Chem. Soc. 2010, 132, 5309– 5314
 DOI: 10.1021/ja101029u [ACS Full Text ], [CAS], [Google Scholar]

660. Crowley, J. D.; Goldup, S. M.; Gowans, N. D.; Leigh, D. A.; Ronaldson, V. E.; Slawin, A. M. Z. An Unusual
Nickel–Copper-Mediated Alkyne Homocoupling Reaction for the Active-Template Synthesis of [2]Rotaxanes J.
Am. Chem. Soc. 2010, 132, 6243– 6248 DOI: 10.1021/ja101121j [ACS Full Text ], [CAS], [Google Scholar]

661. De Bo, G.; Kuschel, S.; Leigh, D. A.; Lewandowski, B.; Papmeyer, M.; Ward, J. W. E cient Assembly of
Threaded Molecular Machines for Sequence-Speci c Synthesis J. Am. Chem. Soc. 2014, 136, 5811– 5814
 DOI: 10.1021/ja5022415 [ACS Full Text ], [CAS], [Google Scholar]

  
662. Goldup, S. M.; Leigh, D. A.; Lusby, P. J.; McBurney, R. T.; Slawin, A. M. Z. Gold(I)-Template Catenane and
Rotaxane Synthesis Angew. Chem., Int. Ed. 2008, 47, 6999– 7003 DOI: 10.1002/anie.200801904 [Crossref],
[PubMed], [CAS], [Google Scholar]

663. Goldup, S. M.; Leigh, D. A.; McBurney, R. T.; McGonigal, P. R.; Plant, A. Ligand-assisted Nickel-catalysed sp3-
sp3 Homocoupling of Unactivated Alkyl Bromides and its Application to the Active Template Synthesis of
Rotaxanes Chem. Sci. 2010, 1, 383– 386 DOI: 10.1039/c0sc00279h [Crossref], [CAS], [Google Scholar]

664. Ahmed, R.; Altieri, A.; D’Souza, D. M.; Leigh, D. A.; Mullen, K. M.; Papmeyer, M.; Slawin, A. M. Z.; Wong, J. K.
Y.; Woollins, J. D. Phosphorus-Based Functional Groups as Hydrogen Bonding Templates for Rotaxane
Formation J. Am. Chem. Soc. 2011, 133, 12304– 12310 DOI: 10.1021/ja2049786 [ACS Full Text ],
[CAS], [Google Scholar]

665. Cheng, H. M.; Leigh, D. A.; Maffei, F.; McGonigal, P. R.; Slawin, A. M. Z.; Wu, J. En Route to a Molecular Sheaf:
Active Metal Template Synthesis of a [3]Rotaxane with Two Axles Threaded through One Ring J. Am. Chem. Soc.
2011, 133, 12298– 12303 DOI: 10.1021/ja205167e [ACS Full Text ], [CAS], [Google Scholar]

666. Lewandowski, B.; De Bo, G.; Ward, J. W.; Papmeyer, M.; Kuschel, S.; Aldegunde, M. J.; Gramlich, P. M. E.;
Heckmann, D.; Goldup, S. M.; D’Souza, D. M.; Fernandes, A. E.; Leigh, D. A. Sequence-Speci c Peptide Synthesis by
an Arti cial Small-Molecule Machine Science 2013, 339, 189– 193 DOI: 10.1126/science.1229753 [Crossref],
[PubMed], [CAS], [Google Scholar]

667. Leigh, D. A.; Lusby, P. J.; McBurney, R. T.; Morelli, A.; Slawin, A. M. Z.; Thomson, A. R.; Walker, D. B. Getting
Harder: Cobalt(III)-Template Synthesis of Catenanes and Rotaxanes J. Am. Chem. Soc. 2009, 131, 3762– 3771
 DOI: 10.1021/ja809627j [ACS Full Text ], [CAS], [Google Scholar]

668. Berná, J.; Goldup, S. M.; Lee, A.-L.; Leigh, D. A.; Symes, M. D.; Teobaldi, G.; Zerbetto, F. Cadiot–Chodkiewicz
Active Template Synthesis of Rotaxanes and Switchable Molecular Shuttles with Weak Intercomponent
Interactions Angew. Chem., Int. Ed. 2008, 47, 4392– 4396 DOI: 10.1002/anie.200800891 [Crossref], [PubMed],
[CAS], [Google Scholar]

669. Liu, L.; Liu, Y.; Liu, P.; Wu, J.; Guan, Y.; Hu, X.; Lin, C.; Yang, Y.; Sun, X.; Ma, J.; Wang, L. Phosphine Oxide
Functional Group Based Three-station Molecular Shuttle Chem. Sci. 2013, 4, 1701– 1706
 DOI: 10.1039/c3sc22048f [Crossref], [CAS], [Google Scholar]
670. Xue, M.; Yang, Y.; Chi, X.; Yan, X.; Huang, F. Development of Pseudorotaxanes and Rotaxanes: From
  
Synthesis to Stimuli-Responsive Motions to Applications Chem. Rev. 2015, 115, 7398 DOI: 10.1021/cr5005869
[ACS Full Text ], [CAS], [Google Scholar]

671. Franz, M.; Januszewski, J. A.; Wendinger, D.; Neiss, C.; Movsisyan, L. D.; Hampel, F.; Anderson, H. L.; Gorling,
A.; Tykwinski, R. R. Cumulene Rotaxanes: Stabilization and Study of [9]Cumulenes Angew. Chem., Int. Ed. 2015,
54, 6645– 6649 DOI: 10.1002/anie.201501810 [Crossref], [PubMed], [CAS], [Google Scholar]

672. Langton, M. J.; Matichak, J. D.; Thompson, A. L.; Anderson, H. L. Template-directed Synthesis of π-
Conjugated Porphyrin [2]Rotaxanes and a [4]Catenane Based on a Six-porphyrin Nanoring Chem. Sci. 2011, 2,
1897– 1901 DOI: 10.1039/c1sc00358e [Crossref], [CAS], [Google Scholar]

673. Lee, S.; Chen, C. H.; Flood, A. H. A Pentagonal Cyanostar Macrocycle with Cyanostilbene CH Donors Binds
Anions and Forms Dialkylphosphate [3]Rotaxanes Nat. Chem. 2013, 5, 704– 710 DOI: 10.1038/nchem.1668
[Crossref], [PubMed], [CAS], [Google Scholar]

674. Lahlali, H.; Jobe, K.; Watkinson, M.; Goldup, S. M. Macrocycle Size Matters: ″Small″ Functionalized
Rotaxanes in Excellent Yield Using the CuAAC Active Template Approach Angew. Chem., Int. Ed. 2011, 50, 4151–
4155 DOI: 10.1002/anie.201100415 [Crossref], [PubMed], [CAS], [Google Scholar]

675. Winn, J.; Pinczewska, A.; Goldup, S. M. Synthesis of a Rotaxane Cu(I) Triazolide under Aqueous Conditions
J. Am. Chem. Soc. 2013, 135, 13318– 13321 DOI: 10.1021/ja407446c [ACS Full Text ], [CAS], [Google Scholar]

676. Bordoli, R. J.; Goldup, S. M. An E cient Approach to Mechanically Planar Chiral Rotaxanes J. Am. Chem.
Soc. 2014, 136, 4817– 4820 DOI: 10.1021/ja412715m [ACS Full Text ], [CAS], [Google Scholar]

677. Hübner, G. M.; Gläser, J.; Seel, C.; Vögtle, F. High-Yielding Rotaxane Synthesis with an Anion Template
Angew. Chem., Int. Ed. 1999, 38, 383– 386
 DOI: 10.1002/(SICI)1521-3773(19990201)38:3<383::AID-ANIE383>3.0.CO;2-H [Crossref], [CAS], [Google Scholar]

678. Hiratani, K.; Suga, J.-i.; Nagawa, Y.; Houjou, H.; Tokuhisa, H.; Numata, M.; Watanabe, K. A new Synthetic
Method for Rotaxanes via Tandem Claisen Rearrangement, Diesteri cation, and Aminolysis Tetrahedron Lett.
2002, 43, 5747– 5750 DOI: 10.1016/S0040-4039(02)01201-7 [Crossref], [CAS], [Google Scholar]

679. Hannam, J. S.; Lacy, S. M.; Leigh, D. A.; Saiz, C. G.; Slawin, A. M. Z.; Stitchell, S. G. Controlled Submolecular
Translational Motion in Synthesis: A Mechanically Interlocking Auxiliary Angew. Chem., Int. Ed. 2004, 43, 3260–
3264 DOI: 10.1002/anie.200353606 [Crossref], [PubMed], [CAS], [Google Scholar]
680. Kameta, N.; Hiratani, K.; Nagawa, Y. A Novel Synthesis of Chiral Rotaxanes via Covalent Bond  
Formation
Chem. Commun. 2004, 466– 467 DOI: 10.1039/b314744d [Crossref], [PubMed], [CAS], [Google Scholar]

681. Reuter, C.; Wienand, W.; Hübner, G. M.; Seel, C.; Vögtle, F. High-Yield Synthesis of Ester, Carbonate, and
Acetal Rotaxanes by Anion Template Assistance and their Hydrolytic Dethreading Chem. - Eur. J. 1999, 5, 2692–
2697 DOI: 10.1002/(SICI)1521-3765(19990903)5:9<2692::AID-CHEM2692>3.3.CO;2-6 [Crossref], [CAS], [Google
Scholar]

682. Schalley, C. A.; Silva, G.; Nising, C. F.; Linnartz, P. Analysis and Improvement of an Anion-Templated
Rotaxane Synthesis Helv. Chim. Acta 2002, 85, 1578– 1596
 DOI: 10.1002/1522-2675(200206)85:6<1578::AID-HLCA1578>3.0.CO;2-L [Crossref], [CAS], [Google Scholar]

683. Mahoney, J. M.; Shukla, R.; Marshall, R. A.; Beatty, A. M.; Zajicek, J.; Smith, B. D. Templated Conversion of a
Crown Ether-Containing Macrobicycle into [2]Rotaxanes J. Org. Chem. 2002, 67, 1436– 1440
 DOI: 10.1021/jo0162787 [ACS Full Text ], [CAS], [Google Scholar]

684. Li, X.-Y.; Illigen, J.; Nieger, M.; Michel, S.; Schalley, C. A. Tetra- and Octalactam Macrocycles and Catenanes
with Exocyclic Metal Coordination Sites: Versatile Building Blocks for Supramolecular Chemistry Chem. - Eur. J.
2003, 9, 1332– 1347 DOI: 10.1002/chem.200390153 [Crossref], [PubMed], [CAS], [Google Scholar]

685. Nagawa, Y.; Suga, J.-i.; Hiratani, K.; Koyama, E.; Kanesato, M. [3]Rotaxane Synthesized via Covalent Bond
Formation Can Recognize Cations Forming a Sandwich Structure Chem. Commun. 2005, 749– 751
 DOI: 10.1039/b413715a [Crossref], [PubMed], [CAS], [Google Scholar]

686. Schill, G.; Zollenkopf, H.; Rotaxan-Verbindungen, I. Justus Liebigs Ann. Chem. 1969, 721, 53– 74
 DOI: 10.1002/jlac.19697210109 [Crossref], [CAS], [Google Scholar]

687. Langton, M. J.; Robinson, S. W.; Marques, I.; Felix, V.; Beer, P. D. Halogen Bonding in Water Results in
Enhanced Anion Recognition in Acyclic and Rotaxane Hosts Nat. Chem. 2014, 6, 1039– 1043
 DOI: 10.1038/nchem.2111 [Crossref], [PubMed], [CAS], [Google Scholar]

688. Aucagne, V.; Berna, J.; Crowley, J. D.; Goldup, S. M.; Hanni, K. D.; Leigh, D. A.; Lusby, P. J.; Ronaldson, V. E.;
Slawin, A. M.; Viterisi, A.; Walker, D. B. Catalytic ″Active-metal″ Template Synthesis of [2]Rotaxanes, [3]Rotaxanes,
and Molecular Shuttles, and Some Observations on the Mechanism of the Cu(I)-Catalyzed Azide-alkyne 1,3-
Cycloaddition J. Am. Chem. Soc. 2007, 129, 11950– 11963 DOI: 10.1021/ja073513f [ACS Full Text ],
[CAS], [Google Scholar]
689. Browne, C.; Ronson, T. K.; Nitschke, J. R. Palladium-Templated Subcomponent Self-Assemblyof  
Macrocycles, Catenanes, and Rotaxanes Angew. Chem., Int. Ed. 2014, 53, 10701– 10705
 DOI: 10.1002/anie.201406164 [Crossref], [PubMed], [CAS], [Google Scholar]

690. Anelli, P. L.; Spencer, N.; Stoddart, J. F. A Molecular Shuttle J. Am. Chem. Soc. 1991, 113, 5131– 5133
 DOI: 10.1021/ja00013a096 [ACS Full Text ], [CAS], [Google Scholar]

691. Ashton, P. R.; Johnston, M. R.; Stoddart, J. F.; Tolley, M. S.; Wheeler, J. W. The Template-directed Synthesis
of Porphyrin-stoppered [2]Rotaxanes J. Chem. Soc., Chem. Commun. 1992, 1128– 1131
 DOI: 10.1039/c39920001128 [Crossref], [CAS], [Google Scholar]

692. Ashton, P. R.; Philp, D.; Spencer, N.; Stoddart, J. F. A New Design Strategy for the Self-assembly of
Molecular Shuttles J. Chem. Soc., Chem. Commun. 1992, 1124– 1128 DOI: 10.1039/c39920001124 [Crossref],
[CAS], [Google Scholar]

693. Lane, A. S.; Leigh, D. A.; Murphy, A. Peptide-Based Molecular Shuttles J. Am. Chem. Soc. 1997, 119,
11092– 11093 DOI: 10.1021/ja971224t [ACS Full Text ], [CAS], [Google Scholar]

694. Leigh, D. A.; Murphy, A.; Smart, J. P.; Slawin, A. M. Z. Glycylglycine Rotaxanes—The Hydrogen Bond Directed
Assembly of Synthetic Peptide Rotaxanes Angew. Chem., Int. Ed. Engl. 1997, 36, 728– 732
 DOI: 10.1002/anie.199707281 [Crossref], [CAS], [Google Scholar]

695. Leigh, D. A.; Murphy, A.; Smart, J. P.; Deleuze, M. S.; Zerbetto, F. Controlling the Frequency of Macrocyclic
Ring Rotation in Benzylic Amide [2]Catenanes J. Am. Chem. Soc. 1998, 120, 6458– 6467
 DOI: 10.1021/ja974065m [ACS Full Text ], [CAS], [Google Scholar]

696. Leigh, D. A.; Troisi, A.; Zerbetto, F. A Quantum-Mechanical Description of Macrocyclic Ring Rotation in
Benzylic Amide [2]Catenanes Chem. - Eur. J. 2001, 7, 1450– 1454
 DOI: 10.1002/1521-3765(20010401)7:7<1450::AID-CHEM1450>3.0.CO;2-N [Crossref], [PubMed], [CAS], [Google
Scholar]

697. Deleuze, M. S.; Leigh, D. A.; Zerbetto, F. How Do Benzylic Amide [2]Catenane Rings Rotate? J. Am. Chem.
Soc. 1999, 121, 2364– 2379 DOI: 10.1021/ja9815273 [ACS Full Text ], [CAS], [Google Scholar]
698. Panman, M. R.; Bakker, B. H.; den Uyl, D.; Kay, E. R.; Leigh, D. A.; Buma, W. J.; Brouwer, A. M.; Geenevasen, J.
A. J.; Woutersen, S. Water Lubricates Hydrogen-bonded Molecular Machines Nat. Chem. 2013, 5, 929– 
 934 
 DOI: 10.1038/nchem.1744 [Crossref], [PubMed], [CAS], [Google Scholar]

699. Ghosh, P.; Federwisch, G.; Kogej, M.; Schalley, C. A.; Haase, D.; Saak, W.; Lutzen, A.; Gschwind, R. M.
Controlling the Rate of Shuttling Motions in [2]Rotaxanes by Electrostatic Interactions: A Cation as Solvent-
tunable Brake Org. Biomol. Chem. 2005, 3, 2691– 2700 DOI: 10.1039/b506756a [Crossref], [PubMed],
[CAS], [Google Scholar]

700. Cao, J.; Fyfe, M. C. T.; Stoddart, J. F.; Cousins, G. R. L.; Glink, P. T. Molecular Shuttles by the Protecting
Group Approach† J. Org. Chem. 2000, 65, 1937– 1946 DOI: 10.1021/jo991397w [ACS Full Text ],
[CAS], [Google Scholar]

701. Jiang, L.; Okano, J.; Orita, A.; Otera, J. Intermittent Molecular Shuttle as a Binary Switch Angew. Chem., Int.
Ed. 2004, 43, 2121– 2124 DOI: 10.1002/anie.200353534 [Crossref], [PubMed], [CAS], [Google Scholar]

702. Leigh, D. A.; Lusby, P. J.; Slawin, A. M. Z.; Walker, D. B. Rare and Diverse Binding Modes Introduced through
Mechanical Bonding Angew. Chem., Int. Ed. 2005, 44, 4557– 4564 DOI: 10.1002/anie.200500004 [Crossref],
[PubMed], [CAS], [Google Scholar]

703. Martinez-Cuezva, A.; Berna, J.; Orenes, R.-A.; Pastor, A.; Alajarin, M. Small-Molecule Recognition for
Controlling Molecular Motion in Hydrogen-Bond-Assembled Rotaxanes Angew. Chem., Int. Ed. 2014, 53, 6762–
6767 DOI: 10.1002/anie.201402962 [Crossref], [PubMed], [CAS], [Google Scholar]

704. Berna, J.; Alajarin, M.; Marin-Rodriguez, C.; Franco-Pujante, C. Redox Divergent Conversion of a [2]Rotaxane
into Two Distinct Degenerate Partners with Different Shuttling Dynamics Chem. Sci. 2012, 3, 2314– 2320
 DOI: 10.1039/c2sc20488f [Crossref], [CAS], [Google Scholar]

705. Hirose, K.; Shiba, Y.; Ishibashi, K.; Doi, Y.; Tobe, Y. A Shuttling Molecular Machine with Reversible Brake
Function Chem. - Eur. J. 2008, 14, 3427– 3433 DOI: 10.1002/chem.200702001 [Crossref], [PubMed],
[CAS], [Google Scholar]

706. Leigh, D. A.; Troisi, A.; Zerbetto, F. Reducing Molecular Shuttling to a Single Dimension Angew. Chem., Int.
Ed. 2000, 39, 350– 353 DOI: 10.1002/(SICI)1521-3773(20000117)39:2<350::AID-ANIE350>3.3.CO;2-4 [Crossref],
[PubMed], [CAS], [Google Scholar]
707. Günbaş, D. D.; Brouwer, A. M. Degenerate Molecular Shuttles with Flexible and Rigid Spacers J. Org. Chem.
2012, 77, 5724– 5735 DOI: 10.1021/jo300907r [ACS Full Text ], [CAS], [Google Scholar]   

708. Andersen, S. S.; Share, A. I.; Poulsen, B. L. C.; Kørner, M.; Duedal, T.; Benson, C. R.; Hansen, S. W.; Jeppesen,
J. O.; Flood, A. H. Mechanistic Evaluation of Motion in Redox-Driven Rotaxanes Reveals Longer Linkers Hasten
Forward Escapes and Hinder Backward Translations J. Am. Chem. Soc. 2014, 136, 6373– 6384
 DOI: 10.1021/ja5013596 [ACS Full Text ], [CAS], [Google Scholar]

709. Young, P. G.; Hirose, K.; Tobe, Y. Axle Length Does Not Affect Switching Dynamics in Degenerate Molecular
Shuttles with Rigid Spacers J. Am. Chem. Soc. 2014, 136, 7899– 7906 DOI: 10.1021/ja412671k [ACS Full Text
], [CAS], [Google Scholar]

710. Yoon, I.; Benítez, D.; Zhao, Y.-L.; Miljanić, O. Š.; Kim, S.-Y.; Tkatchouk, E.; Leung, K. C. F.; Khan, S. I.; Goddard,
W. A., III; Stoddart, J. F. Functionally Rigid and Degenerate Molecular Shuttles Chem. - Eur. J. 2009, 15, 1115–
1122 DOI: 10.1002/chem.200802096 [Crossref], [PubMed], [CAS], [Google Scholar]

711. Benniston, A. C.; Harriman, A. A Light-Induced Molecular Shuttle Based on a [2]Rotaxane-Derived Triad
Angew. Chem., Int. Ed. Engl. 1993, 32, 1459– 1461 DOI: 10.1002/anie.199314591 [Crossref], [Google Scholar]

712. Murakami, H.; Kawabuchi, A.; Kotoo, K.; Kunitake, M.; Nakashima, N. A Light-driven Molecular Shuttle
Based on a Rotaxane J. Am. Chem. Soc. 1997, 119, 7605– 7606 DOI: 10.1021/ja971438a [ACS Full Text ],
[CAS], [Google Scholar]

713. Murakami, H.; Kawabuchi, A.; Matsumoto, R.; Ido, T.; Nakashima, N. A Multi-mode-driven Molecular Shuttle:
Photochemically and Thermally Reactive Azobenzene Rotaxanes J. Am. Chem. Soc. 2005, 127, 15891– 15899
 DOI: 10.1021/ja053690l [ACS Full Text ], [CAS], [Google Scholar]

714. Clegg, W.; Gimenez-Saiz, C.; Leigh, D. A.; Murphy, A.; Slawin, A. M. Z.; Teat, S. J. ″Smart″ Rotaxanes: Shape
Memory and Control in Tertiary Amide Peptido[2]rotaxanes J. Am. Chem. Soc. 1999, 121, 4124– 4129
 DOI: 10.1021/ja9841310 [ACS Full Text ], [CAS], [Google Scholar]

715. Asakawa, M.; Brancato, G.; Fanti, M.; Leigh, D. A.; Shimizu, T.; Slawin, A. M. Z.; Wong, J. K. Y.; Zerbetto, F.;
Zhang, S. W. Switching ″On″ and ″Off″ the Expression of Chirality in Peptide Rotaxanes J. Am. Chem. Soc. 2002,
124, 2939– 2950 DOI: 10.1021/ja015995f [ACS Full Text ], [CAS], [Google Scholar]

716. Dong, S.; Yuan, J.; Huang, F. A Pillar[5]arene/imidazolium [2]Rotaxane: Solvent- and Thermo-driven
Molecular Motions and Supramolecular Gel Formation Chem. Sci. 2014, 5, 247– 252 DOI: 10.1039/C3SC52481G
[Crossref], [CAS], [Google Scholar]

  
717. Zhang, Z.; Han, C.; Yu, G.; Huang, F. A Solvent-driven Molecular Spring Chem. Sci. 2012, 3, 3026– 3031
 DOI: 10.1039/c2sc20728a [Crossref], [CAS], [Google Scholar]

718. Bissell, R. A.; Cordova, E.; Kaifer, A. E.; Stoddart, J. F. A Chemically and Electrochemically Switchable
Molecular Shuttle Nature 1994, 369, 133– 137 DOI: 10.1038/369133a0 [Crossref], [CAS], [Google Scholar]

719. Castro, R.; Berardi, M. J.; Cordova, E.; de Olza, M. O.; Kaifer, A. E.; Evanseck, J. D. Unexpected Roles of Guest
Polarizability and Maximum Hardness, and of Host Solvation in Supramolecular Inclusion Complexes: A Dual
Theoretical and Experimental Study J. Am. Chem. Soc. 1996, 118, 10257– 10268 DOI: 10.1021/ja960700x [ACS
Full Text ], [CAS], [Google Scholar]

720. Castro, R.; Nixon, K. R.; Evanseck, J. D.; Kaifer, A. E. Effects of Side Arm Length and Structure of para-
Substituted Phenyl Derivatives on Their Binding to the Host Cyclobis(paraquat-p-phenylene) J. Org. Chem. 1996,
61, 7298– 7303 DOI: 10.1021/jo9610321 [ACS Full Text ], [CAS], [Google Scholar]

721. Kaminski, G. A.; Jorgensen, W. L. Host-guest Chemistry of Rotaxanes and Catenanes: Application of a
Polarizable All-atom Force Field to Cyclobis(paraquat-p-phenylene) Complexes with Disubstituted Benzenes and
Biphenyls J. Chem. Soc., Perkin Trans. 2 1999, 2365– 2375 DOI: 10.1039/a905160k [Crossref], [CAS], [Google
Scholar]

722. Macias, A. T.; Kumar, K. A.; Marchand, A. P.; Evanseck, J. D. Computational Studies of Inclusion Phenomena
and Synthesis of a Novel and Selective Molecular Receptor for 1,4-Disubstituted Benzenes and 4,4′-Disubstituted
Biphenyls J. Org. Chem. 2000, 65, 2083– 2089 DOI: 10.1021/jo991669v [ACS Full Text ], [CAS], [Google
Scholar]

723. Zhang, K. C.; Liu, L.; Mu, T. W.; Guo, Q. X. Ab initio Calculations on the Inclusion Complexation of
Cyclobis(paraquat-p-phenylene) Chem. Phys. Lett. 2001, 333, 195– 198 DOI: 10.1016/S0009-2614(00)01337-3
[Crossref], [CAS], [Google Scholar]

724. Grabuleda, X.; Ivanov, P.; Jaime, C. Shuttling Process in [2]Rotaxanes. Modeling by Molecular Dynamics
and Free Energy Perturbation Simulations J. Phys. Chem. B 2003, 107, 7582– 7588 DOI: 10.1021/jp034658l
[ACS Full Text ], [CAS], [Google Scholar]

725. Grabuleda, X.; Jaime, C. Molecular Shuttles. A Computational Study (MM and MD) on the Translational
Isomerism in Some [2]Rotaxanes J. Org. Chem. 1998, 63, 9635– 9643 DOI: 10.1021/jo980400t [ACS Full Text
], [CAS], [Google Scholar]

  
726. Ercolani, G.; Mencarelli, P. Role of Face-to-face and Edge-to-face Aromatic Interactions in the Inclusion
Complexation of Cyclobis(paraquat-p-phenylene): A Theoretical Study J. Org. Chem. 2003, 68, 6470– 6473
 DOI: 10.1021/jo026887u [ACS Full Text ], [CAS], [Google Scholar]

727. Castro, R.; Davidov, P. D.; Kumar, K. A.; Marchand, A. P.; Evanseck, J. D.; Kaifer, A. E. Inclusion Complexation
of Cyclobis(paraquat-p-phenylene) and Related Cyclophane Derivatives with Substituted Aromatics: Cooperative
Non-covalent Cavity and External Interactions J. Phys. Org. Chem. 1997, 10, 369– 382
 DOI: 10.1002/(SICI)1099-1395(199705)10:5<369::AID-POC917>3.0.CO;2-I [Crossref], [CAS], [Google Scholar]

728. Grabuleda, X.; Ivanov, P.; Jaime, C. Computational Studies on Pseudorotaxanes by Molecular Dynamics
and Free Energy Perturbation Simulations J. Org. Chem. 2003, 68, 1539– 1547 DOI: 10.1021/jo0265636 [ACS
Full Text ], [CAS], [Google Scholar]

729. Martinez-Diaz, M. V.; Spencer, N.; Stoddart, J. F. The Self-assembly of a Switchable [2]Rotaxane Angew.
Chem., Int. Ed. Engl. 1997, 36, 1904– 1907 DOI: 10.1002/anie.199719041 [Crossref], [CAS], [Google Scholar]

730. Badjic, J. D.; Balzani, V.; Credi, A.; Silvi, S.; Stoddart, J. F. A Molecular Elevator Science 2004, 303, 1845–
1849 DOI: 10.1126/science.1094791 [Crossref], [PubMed], [CAS], [Google Scholar]

731. Leigh, D. A.; Thomson, A. R. Switchable Dual Binding Mode Molecular Shuttle Org. Lett. 2006, 8, 5377–
5379 DOI: 10.1021/ol062284j [ACS Full Text ], [CAS], [Google Scholar]

732. Vella, S. J.; Tiburcio, J.; Loeb, S. J. Optically Sensed, Molecular Shuttles Driven by Acid-base Chemistry
Chem. Commun. 2007, 4752– 4754 DOI: 10.1039/b710708k [Crossref], [PubMed], [CAS], [Google Scholar]

733. Zheng, H. Y.; Zhou, W. D.; Lv, J.; Yin, X. D.; Li, Y. J.; Liu, H. B.; Li, Y. L. A Dual-Response [2]Rotaxane Based on
a 1,2,3-Triazole Ring as a Novel Recognition Station Chem. - Eur. J. 2009, 15, 13253– 13262
 DOI: 10.1002/chem.200901841 [Crossref], [PubMed], [CAS], [Google Scholar]

734. Tuncel, D.; Katterle, M. pH-triggered Dethreading-rethreading and Switching of Cucurbit[6]uril on Bistable
[3]Pseudorotaxanes and [3]Rotaxanes Chem. - Eur. J. 2008, 14, 4110– 4116 DOI: 10.1002/chem.200702003
[Crossref], [PubMed], [CAS], [Google Scholar]
735. Kolchinski, A. G.; Busch, D. H.; Alcock, N. W. Gaining Control over Molecular Threading - Bene ts of 2nd

 1995,
Coordination Sites and Aqueous-Organic Interfaces in Rotaxane Synthesis J. Chem. Soc., Chem. Commun.
1289– 1291 DOI: 10.1039/c39950001289 [Crossref], [CAS], [Google Scholar]

736. Glink, P. T.; Schiavo, C.; Stoddart, J. F.; Williams, D. J. The Genesis of a New Range of Interlocked Molecules
Chem. Commun. 1996, 1483– 1490 DOI: 10.1039/cc9960001483 [Crossref], [CAS], [Google Scholar]

737. Cantrill, S. J.; Pease, A. R.; Stoddart, J. F. A Molecular Meccano Kit J. Chem. Soc. Dalton 2000, 3715– 3734
 DOI: 10.1039/b003769i [Crossref], [Google Scholar]

738. Ashton, P. R.; Glink, P. T.; Stoddart, J. F.; Tasker, P. A.; White, A. J. P.; Williams, D. J. Self-assembling [2]- and
[3]Rotaxanes from Secondary Dialkylammonium Salts and Crown Ethers Chem. - Eur. J. 1996, 2, 729– 736
 DOI: 10.1002/chem.19960020617 [Crossref], [CAS], [Google Scholar]

739. Ashton, P. R.; Ballardini, R.; Balzani, V.; Baxter, I.; Credi, A.; Fyfe, M. C. T.; Gandol , M. T.; Gomez-Lopez, M.;
Martinez-Diaz, M. V.; Piersanti, A., et al. Acid-base Controllable Molecular Shuttles J. Am. Chem. Soc. 1998, 120,
11932– 11942 DOI: 10.1021/ja982167m [ACS Full Text ], [CAS], [Google Scholar]

740. Garaudee, S.; Silvi, S.; Venturi, M.; Credi, A.; Flood, A. H.; Stoddart, J. F. Shuttling Dynamics in an Acid-base-
switchable [2]Rotaxane ChemPhysChem 2005, 6, 2145– 2152 DOI: 10.1002/cphc.200500295 [Crossref],
[PubMed], [CAS], [Google Scholar]

741. Frankfort, L.; Sohlberg, K. Semi-empirical Study of a pH-switchable [2] Rotaxane J. Mol. Struct.:
THEOCHEM 2003, 621, 253– 260 DOI: 10.1016/S0166-1280(02)00641-3 [Crossref], [CAS], [Google Scholar]

742. Elizarov, A. M.; Chiu, S. H.; Stoddart, J. F. An Acid-base Switchable [2]Rotaxane J. Org. Chem. 2002, 67,
9175– 9181 DOI: 10.1021/jo020373d [ACS Full Text ], [CAS], [Google Scholar]

743. Badjic, J. D.; Ronconi, C. M.; Stoddart, J. F.; Balzani, V.; Silvi, S.; Credi, A. Operating Molecular Elevators J.
Am. Chem. Soc. 2006, 128, 1489– 1499 DOI: 10.1021/ja0543954 [ACS Full Text ], [CAS], [Google Scholar]

744. Bruns, C. J.; Stoddart, J. F. Molecular Machines Muscle Up Nat. Nanotechnol. 2012, 8, 9– 10
 DOI: 10.1038/nnano.2012.239 [Crossref], [PubMed], [Google Scholar]

745. Clark, P. G.; Day, M. W.; Grubbs, R. H. Switching and Extension of a [c2]Daisy-Chain Dimer Polymer J. Am.
Chem. Soc. 2009, 131, 13631– 13633 DOI: 10.1021/ja905924u [ACS Full Text ], [CAS], [Google Scholar]
  
746. Du, G. Y.; Moulin, E.; Jouault, N.; Buhler, E.; Giuseppone, N. Muscle-like Supramolecular Polymers:
Integrated Motion from Thousands of Molecular Machines Angew. Chem., Int. Ed. 2012, 51, 12504– 12508
 DOI: 10.1002/anie.201206571 [Crossref], [PubMed], [CAS], [Google Scholar]

747. Keaveney, C. M.; Leigh, D. A. Shuttling Through Anion Recognition Angew. Chem., Int. Ed. 2004, 43, 1222–
1224 DOI: 10.1002/anie.200353248 [Crossref], [PubMed], [CAS], [Google Scholar]

748. Hesseler, B.; Zindler, M.; Herges, R.; Lüning, U. A Shuttle for the Transport of Protons Based on a
[2]Rotaxane Eur. J. Org. Chem. 2014, 2014, 3885– 3901 DOI: 10.1002/ejoc.201402249 [Crossref], [CAS], [Google
Scholar]

749. Hesseler, B.; Zindler, M.; Herges, R.; Lüning, U. A Shuttle for the Transport of Protons Based on a
[2]Rotaxane Eur. J. Org. Chem. 2014, 2014, 3885– 3901 DOI: 10.1002/ejoc.201402249 [Crossref], [CAS], [Google
Scholar]

750. Crowley, J. D.; Leigh, D. A.; Lusby, P. J.; McBurney, R. T.; Perret-Aebi, L. E.; Petzold, C.; Slawin, A. M. Z.;
Symes, M. D. A Switchable Palladium-complexed Molecular Shuttle and its Metastable Positional Isomers J. Am.
Chem. Soc. 2007, 129, 15085– 15090 DOI: 10.1021/ja076570h [ACS Full Text ], [CAS], [Google Scholar]

751. Buyukcakir, O.; Yasar, F. T.; Bozdemir, O. A.; Icli, B.; Akkaya, E. U. Autonomous Shuttling Driven by an
Oscillating Reaction: Proof of Principle in a Cucurbit[7]uril-Bodipy Pseudorotaxane Org. Lett. 2013, 15, 1012–
1015 DOI: 10.1021/ol303495n [ACS Full Text ], [CAS], [Google Scholar]

752. Joosten, A.; Trolez, Y.; Collin, J.-P.; Heitz, V.; Sauvage, J.-P. Copper(I)-Assembled [3]Rotaxane Whose Two
Rings Act as Flapping Wings J. Am. Chem. Soc. 2012, 134, 1802– 1809 DOI: 10.1021/ja210113y [ACS Full Text
], [CAS], [Google Scholar]

753. Ashton, P. R.; Bissell, R. A.; Spencer, N.; Stoddart, J. F.; Tolley, M. S. Towards Controllable Molecular Shuttles
- 1 Synlett 1992, 1992, 914– 918 DOI: 10.1055/s-1992-21540 [Crossref], [Google Scholar]

754. Ashton, P. R.; Bissell, R. A.; Gorski, R.; Philp, D.; Spencer, N.; Stoddart, J. F.; Tolley, M. S. Towards Controllable
Molecular Shuttles - 2 Synlett 1992, 1992, 919– 922 DOI: 10.1055/s-1992-21541 [Crossref], [Google Scholar]

755. Ashton, P. R.; Bissell, R. A.; Spencer, N.; Stoddart, J. F.; Tolley, M. S. Towards Controllable Molecular Shuttles
- 3 Synlett 1992, 1992, 923– 926 DOI: 10.1055/s-1992-21542 [Crossref], [Google Scholar]
756. Anelli, P. L.; Asakawa, M.; Ashton, P. R.; Bissell, R. A.; Clavier, G.; Gorski, R.; Kaifer, A. E.; Langford, S. J.;
Mattersteig, G.; Menzer, S., et al. Toward Controllable Molecular Shuttles Chem. - Eur. J. 1997, 3, 1113– 
 1135 
 DOI: 10.1002/chem.19970030719 [Crossref], [CAS], [Google Scholar]

757. Brouwer, A. M.; Frochot, C.; Gatti, F. G.; Leigh, D. A.; Mottier, L.; Paolucci, F.; Ro a, S.; Wurpel, G. W. H.
Photoinduction of Fast, Reversible Translational Motion in a Hydrogen-bonded Molecular Shuttle Science 2001,
291, 2124– 2128 DOI: 10.1126/science.1057886 [Crossref], [PubMed], [CAS], [Google Scholar]

758. Altieri, A.; Gatti, F. G.; Kay, E. R.; Leigh, D. A.; Martel, D.; Paolucci, F.; Slawin, A. M. Z.; Wong, J. K. Y.
Electrochemically Switchable Hydrogen-bonded Molecular Shuttles J. Am. Chem. Soc. 2003, 125, 8644– 8654
 DOI: 10.1021/ja0352552 [ACS Full Text ], [CAS], [Google Scholar]

759. Zheng, X. G.; Sohlberg, K. Modeling of a Rotaxane-based Molecular Device J. Phys. Chem. A 2003, 107,
1207– 1215 DOI: 10.1021/jp0267611 [ACS Full Text ], [CAS], [Google Scholar]

760. Raiteri, P.; Bussi, G.; Cucinotta, C. S.; Credi, A.; Stoddart, J. F.; Parrinello, M. Unravelling the Shuttling
Mechanism in a Photoswitchable Multicomponent Bistable Rotaxane Angew. Chem., Int. Ed. 2008, 47, 3536–
3539 DOI: 10.1002/anie.200705207 [Crossref], [PubMed], [CAS], [Google Scholar]

761. Tseng, H. R.; Vignon, S. A.; Stoddart, J. F. Toward Chemically Controlled Nanoscale Molecular Machinery
Angew. Chem., Int. Ed. 2003, 42, 1491– 1495 DOI: 10.1002/anie.200250453 [Crossref], [PubMed], [CAS], [Google
Scholar]

762. Fioravanti, G.; Haraszkiewicz, N.; Kay, E. R.; Mendoza, S. M.; Bruno, C.; Marcaccio, M.; Wiering, P. G.;
Paolucci, F.; Rudolf, P.; Brouwer, A. M., et al. Three State Redox-active Molecular Shuttle That Switches in Solution
and on a Surface J. Am. Chem. Soc. 2008, 130, 2593– 2601 DOI: 10.1021/ja077223a [ACS Full Text ],
[CAS], [Google Scholar]

763. Zhao, Y. L.; Dichtel, W. R.; Trabolsi, A.; Saha, S.; Aprahamian, I.; Stoddart, J. F. A Redox-switchable alpha-
Cyclodextrin-based [2]Rotaxane J. Am. Chem. Soc. 2008, 130, 11294– 11296 DOI: 10.1021/ja8036146 [ACS Full
Text ], [CAS], [Google Scholar]

764. Saha, S.; Flood, A. H.; Stoddart, J. F.; Impellizzeri, S.; Silvi, S.; Venturi, M.; Credi, A. A Redox-driven
Multicomponent Molecular Shuttle J. Am. Chem. Soc. 2007, 129, 12159– 12171 DOI: 10.1021/ja0724590 [ACS
Full Text ], [CAS], [Google Scholar]
765. Barnes, J. C.; Fahrenbach, A. C.; Dyar, S. M.; Frasconi, M.; Giesener, M. A.; Zhu, Z. X.; Liu, Z. C.; Hartlieb, K. J.;
  
Carmieli, R.; Wasielewski, M. R., et al. Mechanically Induced Intramolecular Electron Transfer in a Mixed-valence
Molecular Shuttle Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 11546– 11551 DOI: 10.1073/pnas.1201561109
[Crossref], [PubMed], [CAS], [Google Scholar]

766. Amabilino, D. B.; Ashton, P. R.; Boyd, S. E.; GomezLopez, M.; Hayes, W.; Stoddart, J. F. Translational
Isomerism in Some Two- and Three-station [2]Rotaxanes J. Org. Chem. 1997, 62, 3062– 3075
 DOI: 10.1021/jo9612584 [ACS Full Text ], [CAS], [Google Scholar]

767. Asakawa, M.; Ashton, P. R.; Balzani, V.; Credi, A.; Hamers, C.; Mattersteig, G.; Montalti, M.; Shipway, A. N.;
Spencer, N.; Stoddart, J. F., et al. A Chemically and Electrochemically Switchable [2]Catenane Incorporating a
Tetrathiafulvalene Unit Angew. Chem., Int. Ed. 1998, 37, 333– 337
 DOI: 10.1002/(SICI)1521-3773(19980216)37:3<333::AID-ANIE333>3.0.CO;2-P [Crossref], [PubMed],
[CAS], [Google Scholar]

768. Jeppesen, J. O.; Perkins, J.; Becher, J.; Stoddart, J. F. Slow Shuttling in an Amphiphilic Bistable [2]Rotaxane
Incorporating a Tetrathiafulvalene Unit Angew. Chem., Int. Ed. 2001, 40, 1216– 1221
 DOI: 10.1002/1521-3773(20010401)40:7<1216::AID-ANIE1216>3.0.CO;2-W [Crossref], [PubMed], [CAS], [Google
Scholar]

769. Tseng, H. R.; Vignon, S. A.; Celestre, P. C.; Perkins, J.; Jeppesen, J. O.; Di Fabio, A.; Ballardini, R.; Gandol , M.
T.; Venturi, M.; Balzani, V., et al. Redox-controllable Amphiphilic [2]Rotaxanes Chem. - Eur. J. 2004, 10, 155– 172
 DOI: 10.1002/chem.200305204 [Crossref], [PubMed], [CAS], [Google Scholar]

770. Flood, A. H.; Peters, A. J.; Vignon, S. A.; Steuerman, D. W.; Tseng, H. R.; Kang, S.; Heath, J. R.; Stoddart, J. F.
The Role of Physical Environment on Molecular Electromechanical Switching Chem. - Eur. J. 2004, 10, 6558–
6564 DOI: 10.1002/chem.200401052 [Crossref], [PubMed], [CAS], [Google Scholar]

771. Jeppesen, J. O.; Nygaard, S.; Vignon, S. A.; Stoddart, J. F. Honing up a Genre of Amphiphilic Bistable
[2]rotaxanes for Device Settings Eur. J. Org. Chem. 2005, 2005, 196– 220 DOI: 10.1002/ejoc.200400530
[Crossref], [Google Scholar]

772. Choi, J. W.; Flood, A. H.; Steuerman, D. W.; Nygaard, S.; Braunschweig, A. B.; Moonen, N. N. P.; Laursen, B.
W.; Luo, Y.; DeIonno, E.; Peters, A. J., et al. Ground-state Equilibrium Thermodynamics and Switching Kinetics of
Bistable [2]Rotaxanes Switched in Solution, Polymer Gels, and Molecular Electronic Devices Chem. - Eur. J. 2006,
12, 261– 279 DOI: 10.1002/chem.200500934 [Crossref], [CAS], [Google Scholar]
773. Kang, S. S.; Vignon, S. A.; Tseng, H. R.; Stoddart, J. F. Molecular Shuttles Based on Tetrathiafulvalene Units
  
and 1,5-Dioxynaphthalene Ring Systems Chem. - Eur. J. 2004, 10, 2555– 2564 DOI: 10.1002/chem.200305725
[Crossref], [PubMed], [CAS], [Google Scholar]

774. Jeppesen, J. O.; Nielsen, K. A.; Perkins, J.; Vignon, S. A.; Di Fabio, A.; Ballardini, R.; Gandol , M. T.; Venturi,
M.; Balzani, V.; Becher, J., et al. Amphiphilic Bistable Rotaxanes Chem. - Eur. J. 2003, 9, 2982– 3007
 DOI: 10.1002/chem.200204589 [Crossref], [CAS], [Google Scholar]

775. Ashton, P. R.; Ballardini, R.; Balzani, V.; Credi, A.; Dress, K. R.; Ishow, E.; Kleverlaan, C. J.; Kocian, O.; Preece, J.
A.; Spencer, N., et al. A Photochemically Driven Molecular-level Abacus Chem. - Eur. J. 2000, 6, 3558– 3574
 DOI: 10.1002/1521-3765(20001002)6:19<3558::AID-CHEM3558>3.0.CO;2-M [Crossref], [PubMed], [CAS], [Google
Scholar]

776. Ballardini, R.; Balzani, V.; Dehaen, W.; Dell’Erba, A. E.; Raymo, F. M.; Stoddart, J. F.; Venturi, M. Molecular
Meccano, 56 - Anthracene-containing [2]Rotaxanes: Synthesis, Spectroscopic, and Electrochemical Properties
Eur. J. Org. Chem. 2000, 2000, 591– 602
 DOI: 10.1002/(SICI)1099-0690(200002)2000:4<591::AID-EJOC591>3.0.CO;2-I [Crossref], [Google Scholar]

777. Kihara, N.; Hashimoto, M.; Takata, T. Redox Behavior of Ferrocene-containing Rotaxane: Transposition of
the Rotaxane Wheel by Redox Reaction of a Ferrocene Moiety Tethered at the End of the Axle Org. Lett. 2004, 6,
1693– 1696 DOI: 10.1021/ol049817d [ACS Full Text ], [CAS], [Google Scholar]

778. Armaroli, N.; Balzani, V.; Collin, J.-P.; Gavina, P.; Sauvage, J.-P.; Ventura, B. Rotaxanes Incorporating Two
Different Coordinating Units in Their Thread: Synthesis and Electrochemically and Photochemically Induced
Molecular Motions J. Am. Chem. Soc. 1999, 121, 4397– 4408 DOI: 10.1021/ja984051w [ACS Full Text ],
[CAS], [Google Scholar]

779. Trabolsi, A.; Khashab, N.; Fahrenbach, A. C.; Friedman, D. C.; Colvin, M. T.; Cotí, K. K.; Benítez, D.; Tkatchouk,
E.; Olsen, J.-C.; Belowich, M. E., et al. Radically Enhanced Molecular Recognition Nat. Chem. 2010, 2, 42– 49
 DOI: 10.1038/nchem.479 [Crossref], [PubMed], [CAS], [Google Scholar]

780. Vignon, S. A.; Jarrosson, T.; Iijima, T.; Tseng, H. R.; Sanders, J. K. M.; Stoddart, J. F. Switchable Neutral
Bistable Rotaxanes J. Am. Chem. Soc. 2004, 126, 9884– 9885 DOI: 10.1021/ja048080k [ACS Full Text ],
[CAS], [Google Scholar]

781. Marlin, D. S.; Cabrera, D. G.; Leigh, D. A.; Slawin, A. M. Z. Complexation-induced Translational Isomerism:
Shuttling Through Stepwise Competitive Binding Angew. Chem., Int. Ed. 2006, 45, 77– 83
 DOI: 10.1002/anie.200501761 [Crossref], [CAS], [Google Scholar]

  
782. Marlin, D. S.; Cabrera, D. G.; Leigh, D. A.; Slawin, A. M. Z. An Allosterically Regulated Molecular Shuttle
Angew. Chem., Int. Ed. 2006, 45, 1385– 1390 DOI: 10.1002/anie.200502624 [Crossref], [PubMed], [CAS], [Google
Scholar]

783. Barrell, M. J.; Leigh, D. A.; Lusby, P. J.; Slawin, A. M. Z. An Ion-Pair Template for Rotaxane Formation and its
Exploitation in an Orthogonal Interaction Anion-Switchahle Molecular Shuttle Angew. Chem., Int. Ed. 2008, 47,
8036– 8039 DOI: 10.1002/anie.200802745 [Crossref], [PubMed], [CAS], [Google Scholar]

784. Spence, G. T.; Pitak, M. B.; Beer, P. D. Anion-Induced Shuttling of a Naphthalimide Triazolium Rotaxane
Chem. - Eur. J. 2012, 18, 7100– 7108 DOI: 10.1002/chem.201200317 [Crossref], [PubMed], [CAS], [Google
Scholar]

785. Caballero, A.; Swan, L.; Zapata, F.; Beer, P. D. Iodide-Induced Shuttling of a Halogen- and Hydrogen-Bonding
Two-Station Rotaxane Angew. Chem., Int. Ed. 2014, 53, 11854– 11858 DOI: 10.1002/anie.201407580 [Crossref],
[PubMed], [CAS], [Google Scholar]

786. Collins, C. G.; Peck, E. M.; Kramer, P. J.; Smith, B. D. Squaraine Rotaxane Shuttle as a Ratiometric Deep-red
Optical Chloride Sensor Chem. Sci. 2013, 4, 2557– 2563 DOI: 10.1039/c3sc50535a [Crossref], [CAS], [Google
Scholar]

787. You, Y. C.; Tzeng, M. C.; Lai, C. C.; Chiu, S. H. Using Oppositely Charged Ions To Operate a Three-Station
[2]Rotaxane in Two Different Switching Modes Org. Lett. 2012, 14, 1046– 1049 DOI: 10.1021/ol203401d [ACS
Full Text ], [CAS], [Google Scholar]

788. Collin, J.-P.; Frey, J.; Heitz, V.; Sauvage, J.-P.; Tock, C.; Allouche, L. Adjustable Receptor Based on a
[3]Rotaxane Whose Two Threaded Rings Are Rigidly Attached to Two Porphyrinic Plates: Synthesis and
Complexation Studies J. Am. Chem. Soc. 2009, 131, 5609– 5620 DOI: 10.1021/ja900565p [ACS Full Text ],
[CAS], [Google Scholar]

789. Leigh, D. A.; Perez, E. M. Shuttling Through Reversible Covalent Chemistry Chem. Commun. 2004, 2262–
2263 DOI: 10.1039/b412570c [Crossref], [PubMed], [CAS], [Google Scholar]

790. Abraham, W.; Grubert, L.; Grummt, U. W.; Buck, K. A Photoswitchable Rotaxane with a Folded Molecular
Thread Chem. - Eur. J. 2004, 10, 3562– 3568 DOI: 10.1002/chem.200400243 [Crossref], [PubMed],
[CAS], [Google Scholar]
 with
791. Schmidt-Schaffer, S.; Grubert, L.; Grummt, U. W.; Buck, K.; Abraham, W. A Photoswitchable Rotaxane
an Unfolded Molecular Thread Eur. J. Org. Chem. 2006, 2006, 378– 398 DOI: 10.1002/ejoc.200500421
[Crossref], [Google Scholar]

792. Umehara, T.; Kawai, H.; Fujiwara, K.; Suzuki, T. Entropy- and Hydrolytic-driven Positional Switching of
Macrocycle Between Imine- and Hydrogen-bonding Stations in Rotaxane-based Molecular Shuttles J. Am. Chem.
Soc. 2008, 130, 13981– 13988 DOI: 10.1021/ja804888b [ACS Full Text ], [CAS], [Google Scholar]

793. Berna, J.; Alajarin, M.; Orenes, R. A. Azodicarboxamides as Template Binding Motifs for the Building of
Hydrogen-Bonded Molecular Shuttles J. Am. Chem. Soc. 2010, 132, 10741– 10747 DOI: 10.1021/ja101151t
[ACS Full Text ], [CAS], [Google Scholar]

794. Balzani, V.; Clemente-Leon, M.; Credi, A.; Ferrer, B.; Venturi, M.; Flood, A. H.; Stoddart, J. F. Autonomous
Arti cial Nanomotor Powered by Sunlight Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 1178– 1183
 DOI: 10.1073/pnas.0509011103 [Crossref], [PubMed], [CAS], [Google Scholar]

795. Wurpel, G. W. H.; Brouwer, A. M.; van Stokkum, I. H. M.; Farran, A.; Leigh, D. A. Enhanced Hydrogen Bonding
Induced by Optical Excitation: Unexpected Subnanosecond Photoinduced Dynamics in a Peptide-based
[2]Rotaxane J. Am. Chem. Soc. 2001, 123, 11327– 11328 DOI: 10.1021/ja015919c [ACS Full Text ],
[CAS], [Google Scholar]

796. Wang, Q. C.; Qu, D. H.; Ren, J.; Chen, K. C.; Tian, H. A Lockable Light-driven Molecular Shuttle with a
Fluorescent Signal Angew. Chem., Int. Ed. 2004, 43, 2661– 2665 DOI: 10.1002/anie.200453708 [Crossref],
[PubMed], [CAS], [Google Scholar]

797. Bottari, G.; Dehez, F.; Leigh, D. A.; Nash, P. J.; Perez, E. M.; Wong, J. K. Y.; Zerbetto, F. Entropy-driven
Translational Isomerism: A Tristable Molecular Shuttle Angew. Chem., Int. Ed. 2003, 42, 5886– 5889
 DOI: 10.1002/anie.200352176 [Crossref], [PubMed], [CAS], [Google Scholar]

798. Altieri, A.; Bottari, G.; Dehez, F.; Leigh, D. A.; Wong, J. K. Y.; Zerbetto, F. Remarkable Positional Discrimination
in Bistable Light- and Heat-switchable Hydrogen-bonded Molecular Shuttles Angew. Chem., Int. Ed. 2003, 42,
2296– 2300 DOI: 10.1002/anie.200250745 [Crossref], [PubMed], [CAS], [Google Scholar]

799. Qu, D. H.; Wang, G. C.; Ren, J.; Tian, H. A Light-driven Rotaxane Molecular Shuttle with Dual Fluorescence
Addresses Org. Lett. 2004, 6, 2085– 2088 DOI: 10.1021/ol049605g [ACS Full Text ], [CAS], [Google Scholar]
800. Zhou, W. D.; Li, J. B.; He, X. R.; Li, C. H.; Lv, J.; Li, Y. L.; Wang, S.; Liu, H. B.; Zhu, D. B. A Molecular Shuttle for
 
Driving a Multilevel Fluorescence Switch Chem. - Eur. J. 2008, 14, 754– 763 DOI: 10.1002/chem.200701105 
[Crossref], [PubMed], [CAS], [Google Scholar]

801. Qu, D. H.; Ji, F. Y.; Wang, Q. C.; Tian, H. A Double INHIBIT Logic Gate Employing Con guration and
Fluorescence Changes Adv. Mater. 2006, 18, 2035– 2038 DOI: 10.1002/adma.200600235 [Crossref],
[CAS], [Google Scholar]

802. Zhou, W. D.; Chen, D. G.; Li, J. B.; Xu, J. L.; Lv, J.; Liu, H. B.; Li, Y. L. Photoisomerization of Spiropyran for
Driving a Molecular Shuttle Org. Lett. 2007, 9, 3929– 3932 DOI: 10.1021/ol7015862 [ACS Full Text ],
[CAS], [Google Scholar]

803. Gunbas, D. D.; Zalewski, L.; Brouwer, A. M. Solvatochromic Rotaxane Molecular Shuttles Chem. Commun.
2011, 47, 4977– 4979 DOI: 10.1039/c0cc05755j [Crossref], [PubMed], [CAS], [Google Scholar]

804. Hanke, A.; Metzler, R. Towards the Molecular Workshop: Entropy-driven Designer Molecules, Entropy
Activation, and Nanomechanical Devices Chem. Phys. Lett. 2002, 359, 22– 26
 DOI: 10.1016/S0009-2614(02)00675-9 [Crossref], [CAS], [Google Scholar]

805. Gong, C. G.; Gibson, H. W. Controlling Microstructure in Polymeric Molecular Shuttles: Solvent-induced
Localization of Macrocycles in Poly(urethane/crown Ether) Rotaxanes Angew. Chem., Int. Ed. Engl. 1997, 36,
2331– 2333 DOI: 10.1002/anie.199723311 [Crossref], [CAS], [Google Scholar]

806. Meng, Z.; Xiang, J. F.; Chen, C. F. Tristable [n]Rotaxanes: From Molecular Shuttle to Molecular Cable Car
Chem. Sci. 2014, 5, 1520– 1525 DOI: 10.1039/c3sc53295j [Crossref], [CAS], [Google Scholar]

807. Busseron, E.; Coutrot, F. N-benzyltriazolium as Both Molecular Station and Barrier in [2]Rotaxane Molecular
Machines J. Org. Chem. 2013, 78, 4099– 106 DOI: 10.1021/jo400414f [ACS Full Text ], [CAS], [Google Scholar]

808. Serreli, V.; Lee, C. F.; Kay, E. R.; Leigh, D. A. A Molecular Information Ratchet Nature 2007, 445, 523– 527
 DOI: 10.1038/nature05452 [Crossref], [PubMed], [CAS], [Google Scholar]

809. Alvarez-Perez, M.; Goldup, S. M.; Leigh, D. A.; Slawin, A. M. Z. A Chemically-driven Molecular Information
Ratchet J. Am. Chem. Soc. 2008, 130, 1836– 1838 DOI: 10.1021/ja7102394 [ACS Full Text ], [CAS], [Google
Scholar]
810. Carlone, A.; Goldup, S. M.; Lebrasseur, N.; Leigh, D. A.; Wilson, A. A Three-compartment Chemically-driven
Molecular Information Ratchet J. Am. Chem. Soc. 2012, 134, 8321– 8323 DOI: 10.1021/ja302711z[ACS
Full 
Text ], [CAS], [Google Scholar]

811. Li, H.; Cheng, C. Y.; McGonigal, P. R.; Fahrenbach, A. C.; Frasconi, M.; Liu, W. G.; Zhu, Z. X.; Zhao, Y. L.; Ke, C.
F.; Lei, J. Y., et al. Relative Unidirectional Translation in an Arti cial Molecular Assembly Fueled by Light J. Am.
Chem. Soc. 2013, 135, 18609– 18620 DOI: 10.1021/ja4094204 [ACS Full Text ], [CAS], [Google Scholar]

812. Cheng, C.; McGonigal, P. R.; Schneebeli, S. T.; Li, H.; Vermeulen, N. A.; Ke, C.; Stoddart, J. F. An Arti cial
Molecular Pump Nat. Nanotechnol. 2015, 10, 547– 553 DOI: 10.1038/nnano.2015.96 [Crossref], [PubMed],
[CAS], [Google Scholar]

813. Cheng, C.; McGonigal, P. R.; Liu, W.-G.; Li, H.; Vermeulen, N. A.; Ke, C.; Frasconi, M.; Stern, C. L.; Goddard, W.
A., III; Stoddart, J. F. Energetically Demanding Transport in a Supramolecular Assembly J. Am. Chem. Soc. 2014,
136, 14702– 14705 DOI: 10.1021/ja508615f [ACS Full Text ], [CAS], [Google Scholar]

814. Baroncini, M.; Silvi, S.; Venturi, M.; Credi, A. Photoactivated Directionally Controlled Transit of a Non-
Symmetric Molecular Axle Through a Macrocycle Angew. Chem., Int. Ed. 2012, 51, 4223– 4226
 DOI: 10.1002/anie.201200555 [Crossref], [PubMed], [CAS], [Google Scholar]

815. Ragazzon, G.; Baroncini, M.; Silvi, S.; Venturi, M.; Credi, A. Light-powered Autonomous and Directional
Molecular Motion of a Dissipative Self-assembling System Nat. Nanotechnol. 2014, 10, 70– 75
 DOI: 10.1038/nnano.2014.260 [Crossref], [PubMed], [Google Scholar]

816. Martinez-Cuezva, A.; Pastor, A.; Cioncoloni, G.; Orenes, R.-A.; Alajarin, M.; Symes, M. D.; Berna, J. Versatile
Control of the Submolecular Motion of Di(acylamino)pyridine-based [2]Rotaxanes Chem. Sci. 2015, 6, 3087–
3094 DOI: 10.1039/C5SC00790A [Crossref], [PubMed], [CAS], [Google Scholar]

817. Leigh, D. A.; Lusby, P. J.; Slawin, A. M. Z.; Walker, D. B. Half-rotation in a Kinetically Locked [2] Catenane
Induced by Transition Metal Ion Substitution Chem. Commun. 2012, 48, 5826– 5828 DOI: 10.1039/c2cc32418k
[Crossref], [PubMed], [CAS], [Google Scholar]

818. Letinois-Halbes, U.; Hanss, D.; Beierle, J. M.; Collin, J.-P.; Sauvage, J.-P. A Fast-moving [2]Rotaxane Whose
Stoppers are Remote from the Copper Complex Core Org. Lett. 2005, 7, 5753– 5756 DOI: 10.1021/ol052051c
[ACS Full Text ], [CAS], [Google Scholar]
819. Nakatani, Y.; Furusho, Y.; Yashima, E. Amidinium Carboxylate Salt Bridges as a Recognition Motif for
Mechanically Interlocked Molecules: Synthesis of an Optically Active [2]Catenane and Control of Its 
Structure 
Angew. Chem., Int. Ed. 2010, 49, 5463– 5467 DOI: 10.1002/anie.201002382 [Crossref], [PubMed], [CAS], [Google
Scholar]

820. Bermudez, V.; Capron, N.; Gase, T.; Gatti, F. G.; Kajzar, F.; Leigh, D. A.; Zerbetto, F.; Zhang, S. W. In uencing
Intramolecular Motion with an Alternating Electric Field Nature 2000, 406, 608– 611 DOI: 10.1038/35020531
[Crossref], [PubMed], [CAS], [Google Scholar]

821. Brancato, G.; Coutrot, F.; Leigh, D. A.; Murphy, A.; Wong, J. K. Y.; Zerbetto, F. From Reactants to Products via
Simple Hydrogen-bonding Networks: Information Transmission in Chemical Reactions Proc. Natl. Acad. Sci. U. S.
A. 2002, 99, 4967– 4971 DOI: 10.1073/pnas.072695799 [Crossref], [PubMed], [CAS], [Google Scholar]

822. Lukin, O.; Kubota, T.; Okamoto, Y.; Schelhase, F.; Yoneva, A.; Muller, W. M.; Muller, U.; Vogtle, F. Knotaxanes -
Rotaxanes with Knots as Stoppers Angew. Chem., Int. Ed. 2003, 42, 4542– 4545 DOI: 10.1002/anie.200351981
[Crossref], [PubMed], [CAS], [Google Scholar]

823. Kapitulnik, A.; Casalnuovo, S.; Lim, K. C.; Heeger, A. J. Electric-Field Coupling to Slow Elastic Modes in Gels
of Conjugated Polymers Phys. Rev. Lett. 1984, 53, 469– 472 DOI: 10.1103/PhysRevLett.53.469 [Crossref],
[CAS], [Google Scholar]

824. Lim, K. C.; Kapitulnik, A.; Zacher, R.; Heeger, A. J. Conformation of Polydiacetylene Macromolecules in
Solution - Field-Induced Birefringence and Rotational Diffusion Constant J. Chem. Phys. 1985, 82, 516– 521
 DOI: 10.1063/1.448773 [Crossref], [CAS], [Google Scholar]

825. Gatti, F. G.; Lent, S.; Wong, J. K. Y.; Bottari, G.; Altieri, A.; Morales, M. A. F.; Teat, S. J.; Frochot, C.; Leigh, D. A.;
Brouwer, A. M., et al. Photoisomerization of a Rotaxane Hydrogen Bonding Template: Light-induced Acceleration
of a Large Amplitude Rotational Motion Proc. Natl. Acad. Sci. U. S. A. 2003, 100, 10– 14
 DOI: 10.1073/pnas.0134757100 [Crossref], [PubMed], [CAS], [Google Scholar]

826. Leigh, D. A.; Moody, K.; Smart, J. P.; Watson, K. J.; Slawin, A. M. Z. Catenane Chameleons: Environment-
sensitive Translational Isomerism in Amphiphilic Benzylic Amide [2]Catenanes Angew. Chem., Int. Ed. Engl. 1996,
35, 306– 310 DOI: 10.1002/anie.199603061 [Crossref], [CAS], [Google Scholar]

827. Kemsley, J. Water Lubricates Molecular Machines Chem. Eng. News 2013, 91, 43– 43 [Google Scholar]
828. Rijs, A. M.; Compagnon, I.; Oomens, J.; Hannam, J. S.; Leigh, D. A.; Buma, W. J. Stiff, and Sticky in the Right
Places: Binding Interactions in Isolated Mechanically Interlocked Molecules Probed by Mid-Infrared  
Spectroscopy
J. Am. Chem. Soc. 2009, 131, 2428– 2429 DOI: 10.1021/ja808788c [ACS Full Text ], [CAS], [Google Scholar]

829. Rijs, A. M.; Sändig, N.; Blom, M. N.; Oomens, J.; Hannam, J. S.; Leigh, D. A.; Zerbetto, F.; Buma, W. J.
Controlled Hydrogen-Bond Breaking in a Rotaxane by Discrete Solvation Angew. Chem., Int. Ed. 2010, 49, 3896–
3900 DOI: 10.1002/anie.201001231 [Crossref], [PubMed], [CAS], [Google Scholar]

830. Linke, M.; Chambron, J. C.; Heitz, V.; Sauvage, J.-P.; Semetey, V. Complete Rearrangement of a Multi-
porphyrinic Rotaxane by Metallation-demetallation of the Central Coordination Site Chem. Commun. 1998,
2469– 2470 DOI: 10.1039/a805746j [Crossref], [CAS], [Google Scholar]

831. Raehm, L.; Kern, J. M.; Sauvage, J.-P. A Transition Metal Containing Rotaxane in Motion: Electrochemically
Induced Pirouetting of the Ring on the Threaded Dumbbell Chem. - Eur. J. 1999, 5, 3310– 3317
 DOI: 10.1002/(SICI)1521-3765(19991105)5:11<3310::AID-CHEM3310>3.0.CO;2-R [Crossref], [CAS], [Google
Scholar]

832. Kern, J. M.; Raehm, L.; Sauvage, J.-P.; Divisia-Blohorn, B.; Vidal, P. L. Controlled Molecular Motions in
Copper-complexed Rotaxanes: An XAS Study Inorg. Chem. 2000, 39, 1555– 1560 DOI: 10.1021/ic991163v [ACS
Full Text ], [CAS], [Google Scholar]

833. Weber, N.; Hamann, C.; Kern, J. M.; Sauvage, J.-P. Synthesis of a Copper [3]Rotaxane Able to Function as an
Electrochemically Driven Oscillatory Machine in Solution, and to Form SAMs on a Metal Surface Inorg. Chem.
2003, 42, 6780– 6792 DOI: 10.1021/ic0347034 [ACS Full Text ], [CAS], [Google Scholar]

834. Poleschak, I.; Kern, J. M.; Sauvage, J.-P. A Copper-complexed Rotaxane in Motion: Pirouetting of the Ring
on the Millisecond Timescale Chem. Commun. 2004, 474– 476 DOI: 10.1039/b315080a [Crossref], [PubMed],
[CAS], [Google Scholar]

835. Ballesteros, B.; Faust, T. B.; Lee, C. F.; Leigh, D. A.; Muryn, C. A.; Pritchard, R. G.; Schultz, D.; Teat, S. J.;
Timco, G. A.; Winpenny, R. E. P. Synthesis, Structure, and Dynamic Properties of Hybrid Organic-Inorganic
Rotaxanes J. Am. Chem. Soc. 2010, 132, 15435– 15444 DOI: 10.1021/ja1074773 [ACS Full Text ],
[CAS], [Google Scholar]

836. Lee, C. F.; Leigh, D. A.; Pritchard, R. G.; Schultz, D.; Teat, S. J.; Timco, G. A.; Winpenny, R. E. P. Hybrid Organic-
inorganic Rotaxanes and Molecular Shuttles Nature 2009, 458, 314– 318 DOI: 10.1038/nature07847 [Crossref],
[PubMed], [CAS], [Google Scholar]
837. van Quaethem, A.; Lussis, P.; Leigh, D. A.; Duwez, A. S.; Fustin, C. A. Probing the Mobility of Catenane Rings
 
in Single Molecules Chem. Sci. 2014, 5, 1449– 1452 DOI: 10.1039/c3sc53113a [Crossref], [CAS], [Google 
Scholar]

838. Evans, N. H.; Serpell, C. J.; Beer, P. D. A [2]Catenane Displaying Pirouetting Motion Triggered by
Debenzylation and Locked by Chloride Anion Recognition Chem. - Eur. J. 2011, 17, 7734– 7738
 DOI: 10.1002/chem.201101033 [Crossref], [PubMed], [CAS], [Google Scholar]

839. Andrievsky, A.; Ahuis, F.; Sessler, J. L.; Vogtle, F.; Gudat, D.; Moini, M. Bipyrrole-based [2]Catenane: A New
Type of Anion Receptor J. Am. Chem. Soc. 1998, 120, 9712– 9713 DOI: 10.1021/ja980755u [ACS Full Text ],
[CAS], [Google Scholar]

840. Leontiev, A. V.; Serpell, C. J.; White, N. G.; Beer, P. D. Cation-induced Molecular Motion of Spring-like
[2]Catenanes Chem. Sci. 2011, 2, 922– 927 DOI: 10.1039/c1sc00034a [Crossref], [CAS], [Google Scholar]

841. Fang, L.; Wang, C.; Fahrenbach, A. C.; Trabolsi, A.; Botros, Y. Y.; Stoddart, J. F. Dual Stimulus Switching of a
[2]Catenane in Water Angew. Chem., Int. Ed. 2011, 50, 1805– 1809 DOI: 10.1002/anie.201006362 [Crossref],
[PubMed], [CAS], [Google Scholar]

842. Amabilino, D. B.; Dietrich-Buchecker, C. O.; Livoreil, A.; PerezGarcia, L.; Sauvage, J.-P.; Stoddart, J. F. A
Switchable Hybrid [2]-Catenane Based on Transition Metal Complexation and pi-Electron Donor-acceptor
Interactions J. Am. Chem. Soc. 1996, 118, 3905– 3913 DOI: 10.1021/ja954329+ [ACS Full Text ],
[CAS], [Google Scholar]

843. Balzani, V.; Credi, A.; Langford, S. J.; Raymo, F. M.; Stoddart, J. F.; Venturi, M. Constructing Molecular
Machinery: A Chemically-switchable [2]Catenane J. Am. Chem. Soc. 2000, 122, 3542– 3543
 DOI: 10.1021/ja994454b [ACS Full Text ], [CAS], [Google Scholar]

844. Grunder, S.; McGrier, P. L.; Whalley, A. C.; Boyle, M. M.; Stern, C.; Stoddart, J. F. A Water-Soluble pH-Triggered
Molecular Switch J. Am. Chem. Soc. 2013, 135, 17691– 17694 DOI: 10.1021/ja409006y [ACS Full Text ],
[CAS], [Google Scholar]

845. Korybut-Daszkiewicz, B.; Wieckowska, A.; Bilewicz, R.; Domagala, S.; Wozniak, K. An Electrochemically
Controlled Molecular Shuttle Angew. Chem., Int. Ed. 2004, 43, 1668– 1672 DOI: 10.1002/anie.200352528
[Crossref], [PubMed], [CAS], [Google Scholar]
846. Flamigni, L.; Talarico, A. M.; Serroni, S.; Puntoriero, F.; Gunter, M. J.; Johnston, M. R.; Jeynes, T. P.
Photoinduced Electron Transfer Between the Interlocked Components of Porphyrin Catenanes: Effect 
 of the 
Presence of Nonequivalent Reduction Sites on the Charge Recombination Rate Chem. - Eur. J. 2003, 9, 2649–
2659 DOI: 10.1002/chem.200204502 [Crossref], [PubMed], [CAS], [Google Scholar]

847. Zheng, X. G.; Sohlberg, K. Modeling Bistability and Switching in a [2]Catenane Phys. Chem. Chem. Phys.
2004, 6, 809– 815 DOI: 10.1039/b310816c [Crossref], [CAS], [Google Scholar]

848. Ceccarelli, M.; Mercuri, F.; Passerone, D.; Parrinello, M. The Microscopic Switching Mechanism of a
[2]Catenane J. Phys. Chem. B 2005, 109, 17094– 17099 DOI: 10.1021/jp051609v [ACS Full Text ],
[CAS], [Google Scholar]

849. Ashton, P. R.; Baldoni, V.; Balzani, V.; Credi, A.; Hoffmann, H. D. A.; Martinez-Diaz, M. V.; Raymo, F. M.;
Stoddart, J. F.; Venturi, M. Dual-mode ″Co-conformational″ Switching in Catenanes Incorporating Bipyridinium
and Dialkylammonium Recognition Sites Chem. - Eur. J. 2001, 7, 3482– 3493
 DOI: 10.1002/1521-3765(20010817)7:16<3482::AID-CHEM3482>3.0.CO;2-G [Crossref], [PubMed], [CAS], [Google
Scholar]

850. Hamilton, D. G.; Montalti, M.; Prodi, L.; Fontani, M.; Zanello, P.; Sanders, J. K. M. Photophysical and
Electrochemical Characterization of the Interactions Between Components in Neutral pi-Associated
[2]Catenanes Chem. - Eur. J. 2000, 6, 608– 617
 DOI: 10.1002/(SICI)1521-3765(20000218)6:4<608::AID-CHEM608>3.0.CO;2-E [Crossref], [PubMed],
[CAS], [Google Scholar]

851. Ashton, P. R.; Ballardini, R.; Balzani, V.; Credi, A.; Gandol , M. T.; Menzer, S.; Perezgarcia, L.; Prodi, L.;
Stoddart, J. F.; Venturi, M. Molecular Meccano 0.4. The Self-Assembly of [2]Catenanes Incorporating Photoactive
and Electroactive Pi-Extended Systems J. Am. Chem. Soc. 1995, 117, 11171– 11197 DOI: 10.1021/ja00150a015
[ACS Full Text ], [CAS], [Google Scholar]

852. Ashton, P. R.; Ballardini, R.; Balzani, V.; Gandol , M. T.; Marquis, D. J. F.; Perezgarcia, L.; Prodi, L.; Stoddart, J.
F.; Venturi, M. The Self-Assembly of Controllable [2]Catenanes J. Chem. Soc., Chem. Commun. 1994, 177– 180
 DOI: 10.1039/c39940000177 [Crossref], [CAS], [Google Scholar]

853. Livoreil, A.; Dietrich-Buchecker, C. O.; Sauvage, J.-P. Electrochemically Triggered Swinging of a [2]-Catenate
J. Am. Chem. Soc. 1994, 116, 9399– 9400 DOI: 10.1021/ja00099a095 [ACS Full Text ], [CAS], [Google Scholar]
854. Baumann, F.; Livoreil, A.; Kaim, W.; Sauvage, J.-P. Changeover in a Multimodal Copper(II) Catenate as
Monitored by EPR Spectroscopy Chem. Commun. 1997, 35– 36 DOI: 10.1039/a606282b [Crossref],
  
[CAS], [Google Scholar]

855. Livoreil, A.; Sauvage, J.-P.; Armaroli, N.; Balzani, V.; Flamigni, L.; Ventura, B. Electrochemically and
Photochemically Driven Ring Motions in a Disymmetrical Copper [2]-Catenate J. Am. Chem. Soc. 1997, 119,
12114– 12124 DOI: 10.1021/ja9720826 [ACS Full Text ], [CAS], [Google Scholar]

856. Cardenas, D. J.; Livoreil, A.; Sauvage, J.-P. Redox Control of the Ring-gliding Motion in a Cu-complexed
Catenane: A Process Involving Three Distinct Geometries J. Am. Chem. Soc. 1996, 118, 11980– 11981
 DOI: 10.1021/ja962774e [ACS Full Text ], [CAS], [Google Scholar]

857. Dietrich-Buchecker, C.; Sauvage, J.-P. Templated Synthesis of Interlocked Macrocyclic Ligands, the
Catenands - Preparation and Characterization of the Prototypical Bis-30 Membered Ring-System Tetrahedron
1990, 46, 503– 512 DOI: 10.1016/S0040-4020(01)85433-8 [Crossref], [CAS], [Google Scholar]

858. Cesario, M.; Dietrich-Buchecker, C. O.; Guilhem, J.; Pascard, C.; Sauvage, J.-P. Molecular-Structure of a
Catenand and Its Copper(I) Catenate - Complete Rearrangement of the Interlocked Macrocyclic Ligands by
Complexation J. Chem. Soc., Chem. Commun. 1985, 244– 247 DOI: 10.1039/c39850000244 [Crossref],
[CAS], [Google Scholar]

859. Cesario, M.; Dietrich, C. O.; Edel, A.; Guilhem, J.; Kintzinger, J. P.; Pascard, C.; Sauvage, J.-P. Topological
Enhancement of Basicity - Molecular-Structure and Solution Study of a Monoprotonated Catenand J. Am. Chem.
Soc. 1986, 108, 6250– 6254 DOI: 10.1021/ja00280a023 [ACS Full Text ], [CAS], [Google Scholar]

860. Albrechtgary, A. M.; Dietrich-Buchecker, C.; Saad, Z.; Sauvage, J.-P. Topological Kinetic Effects -
Complexation of Interlocked Macrocyclic Ligands by Cationic Species J. Am. Chem. Soc. 1988, 110, 1467– 1472
 DOI: 10.1021/ja00213a018 [ACS Full Text ], [Google Scholar]

861. Dietrich-Buchecker, C.; Sauvage, J.-P.; Kern, J. M. Synthesis and Electrochemical Studies of Catenates -
Stabilization of Low Oxidation-States by Interlocked Macrocyclic Ligands J. Am. Chem. Soc. 1989, 111, 7791–
7800 DOI: 10.1021/ja00202a020 [ACS Full Text ], [CAS], [Google Scholar]

862. Albrechtgary, A. M.; Dietrich-Buchecker, C.; Saad, Z.; Sauvage, J.-P. Formation of Li+ and Cd2+ Catenates, a
Reaction with a Negative Enthalpy of Activation J. Chem. Soc., Chem. Commun. 1992, 280– 282
 DOI: 10.1039/c39920000280 [Crossref], [Google Scholar]
863. Armaroli, N.; Decola, L.; Balzani, V.; Sauvage, J.-P.; Dietrich-Buchecker, C. O.; Kern, J. M.; Bailal, A. Absorption
and Emission Properties of a 2-Catenand, Its Protonated Forms, and Its Complexes with Li+, Cu+, Ag+ 2+ 2+
, Co, Ni ,
Zn2+, Pd2+ and Cd2+ - Tuning of the Luminescence over the Whole Visible Spectral Region J. Chem. Soc., Dalton
Trans. 1993, 3241– 3247 DOI: 10.1039/dt9930003241 [Crossref], [CAS], [Google Scholar]

864. Sauvage, J.-P.; Weiss, J. Synthesis of Dicopper(I) [3]Catenates - Multiring Interlocked Coordinating Systems
J. Am. Chem. Soc. 1985, 107, 6108– 6110 DOI: 10.1021/ja00307a049 [ACS Full Text ], [CAS], [Google Scholar]

865. Dietrich-Buchecker, C. O.; Khemiss, A.; Sauvage, J.-P. High-Yield Synthesis of Multiring Copper(I) Catenates
by Acetylenic Oxidative Coupling J. Chem. Soc., Chem. Commun. 1986, 1376– 1378
 DOI: 10.1039/c39860001376 [Crossref], [CAS], [Google Scholar]

866. Mohr, B.; Weck, M.; Sauvage, J. P.; Grubbs, R. H. High-yield synthesis of [2]catenanes by intramolecular
ring-closing metathesis Angew. Chem., Int. Ed. Engl. 1997, 36, 1308– 1310 DOI: 10.1002/anie.199713081
[Crossref], [CAS], [Google Scholar]

867. Koizumi, M.; Dietrich-Buchecker, C.; Sauvage, J.-P. A [2]Catenane Containing 1,1′-Binaphthyl Units and 1,10-
Phenanthroline Fragments: Synthesis and Intermolecular Energy Transfer Processes Eur. J. Org. Chem. 2004,
2004, 770– 775 DOI: 10.1002/ejoc.200300572 [Crossref], [Google Scholar]

868. Fuller, A. M. L.; Leigh, D. A.; Lusby, P. J.; Slawin, A. M. Z.; Walker, D. B. Selecting Topology and Connectivity
Through Metal-directed Macrocyclization Reactions: A Square Planar Palladium [2]Catenate and Two
Noninterlocked Isomers J. Am. Chem. Soc. 2005, 127, 12612– 12619 DOI: 10.1021/ja053005a [ACS Full Text
], [CAS], [Google Scholar]

869. Leigh, D. A.; Lusby, P. J.; Slawin, A. M. Z.; Walker, D. B. Half-rotation in a [2] Catenane via Interconvertible
Pd(II) Coordination Modes Chem. Commun. 2005, 4919– 4921 DOI: 10.1039/b510663j [Crossref], [PubMed],
[CAS], [Google Scholar]

870. Mobian, P.; Kern, J. M.; Sauvage, J.-P. Light-driven Machine Prototypes Based on Dissociative Excited
States: Photoinduced Decoordination and Thermal Recoordination of a Ring in a Ruthenium(II)-containing
[2]Catenane Angew. Chem., Int. Ed. 2004, 43, 2392– 2395 DOI: 10.1002/anie.200352522 [Crossref], [PubMed],
[CAS], [Google Scholar]

871. Baranoff, E.; Collin, J.-P.; Furusho, Y.; Laemmel, A. C.; Sauvage, J.-P. A Photochromic System Based on
Photochemical or Thermal Chelate Exchange on Ru(phen)2L2+ (L = Diimine or Dinitrile Ligand) Chem. Commun.
2000, 1935– 1936 DOI: 10.1039/b004982o [Crossref], [CAS], [Google Scholar]
872. Collin, J.-P.; Laemmel, A. C.; Sauvage, J.-P. Photochemical Expulsion of a Ru(phen)2 Unit From a
  
Macrocyclic Receptor and its Thermal Reco-ordination New J. Chem. 2001, 25, 22– 24 DOI: 10.1039/b007192g
[Crossref], [CAS], [Google Scholar]

873. Laemmel, A. C.; Collin, J.-P.; Sauvage, J.-P.; Accorsi, G.; Armaroli, N. Macrocyclic complexes of [Ru(N-N)2]2+
Units [N-N = 1,10 Phenanthroline or 4-(p-Anisyl)-1,10-Phenanthroline]: Synthesis and Photochemical Expulsion
Studies Eur. J. Inorg. Chem. 2003, 2003, 467– 474 DOI: 10.1002/ejic.200390066 [Crossref], [Google Scholar]

874. Bonnet, S.; Collin, J.-P.; Sauvage, J.-P. A Ru(terpy) (phen)-Incorporating Ring and its Light-induced
Geometrical Changes Chem. Commun. 2005, 3195– 3197 DOI: 10.1039/b503411f [Crossref], [PubMed],
[CAS], [Google Scholar]

875. Collin, J.-P.; Jouvenot, D.; Koizumi, M.; Sauvage, J.-P. Light-driven Expulsion of the Sterically Hindering
Ligand L in Tris-diimine Ruthenium(II) Complexes of the Ru(phen)2L2+ Family: A Pronounced Ring Effect Inorg.
Chem. 2005, 44, 4693– 4698 DOI: 10.1021/ic050246a [ACS Full Text ], [CAS], [Google Scholar]

876. Arico, F.; Mobian, P.; Kern, J. M.; Sauvage, J.-P. Synthesis of a [2]Catenane Around a Ru(diimine)32+ Scaffold
by Ring-closing Metathesis of ole ns Org. Lett. 2003, 5, 1887– 1890 DOI: 10.1021/ol0300367 [ACS Full Text ],
[CAS], [Google Scholar]

877. Pomeranc, D.; Jouvenot, D.; Chambron, J. C.; Collin, J.-P.; Heitz, V.; Sauvage, J.-P. Templated Synthesis of a
Rotaxane with a [Ru(diimine)3]2+ Core Chem. - Eur. J. 2003, 9, 4247– 4254 DOI: 10.1002/chem.200304716
[Crossref], [PubMed], [CAS], [Google Scholar]

878. Laemmel, A. C.; Collin, J.-P.; Sauvage, J.-P. E cient and Selective Photochemical Labilization of a Given
Bidentate Ligand in Mixed Ruthenium(II) Complexes of the Ru(phen)2L2+ and Ru(bipy)2L2+ Family (L = Sterically
Hindering Chelate) Eur. J. Inorg. Chem. 1999, 1999, 383– 386
 DOI: 10.1002/(SICI)1099-0682(199903)1999:3<383::AID-EJIC383>3.0.CO;2-9 [Crossref], [Google Scholar]

879. Vogtle, F.; Muller, W. M.; Muller, U.; Bauer, M.; Rissanen, K. Photoswitchable Catenanes Angew. Chem., Int.
Ed. Engl. 1993, 32, 1295– 1297 DOI: 10.1002/anie.199312951 [Crossref], [Google Scholar]

880. Ashton, P. R.; Blower, M.; Philp, D.; Spencer, N.; Stoddart, J. F.; Tolley, M. S.; Ballardini, R.; Ciano, M.; Balzani,
V.; Gandol , M. T., et al. The Control of Translational Isomerism in Catenated Structures New J. Chem. 1993, 17,
689– 695 [CAS], [Google Scholar]
881. Raymo, F. M.; Houk, K. N.; Stoddart, J. F. Origins of Selectivity in Molecular and Supramolecular Entities:
63, 
Solvent and Electrostatic Control of the Translational Isomerism in [2]Catenanes J. Org. Chem. 1998, 
6523–
6528 DOI: 10.1021/jo980543f [ACS Full Text ], [CAS], [Google Scholar]

882. Ballardini, R.; Balzani, V.; Gandol , M. T.; Gillard, R. E.; Stoddart, J. F.; Tabellini, E. The Synthesis and
Spectroscopic Properties of Macrocyclic Polyethers Containing Two Different Aromatic Moieties and Their
[2]Catenanes Incorporating Cyclobis(paraquat-p-phenylene) Chem. - Eur. J. 1998, 4, 449– 459
 DOI: 10.1002/(SICI)1521-3765(19980310)4:3<449::AID-CHEM449>3.3.CO;2-# [Crossref], [CAS], [Google Scholar]

883. Leigh, D. A.; Wong, J. K. Y.; Dehez, F.; Zerbetto, F. Unidirectional Rotation in a Mechanically Interlocked
Molecular Rotor Nature 2003, 424, 174– 179 DOI: 10.1038/nature01758 [Crossref], [PubMed], [CAS], [Google
Scholar]

884. Baranoff, E. D.; Voignier, J.; Yasuda, T.; Heitz, V.; Sauvage, J.-P.; Kato, T. A Liquid-Crystalline [2]Catenane and
Its Copper(I) Complex Angew. Chem., Int. Ed. 2007, 46, 4680– 4683 DOI: 10.1002/anie.200700308 [Crossref],
[PubMed], [CAS], [Google Scholar]

885. Hernandez, J. V.; Kay, E. R.; Leigh, D. A. A Reversible Synthetic Rotary Molecular Motor Science 2004, 306,
1532– 1537 DOI: 10.1126/science.1103949 [Crossref], [PubMed], [CAS], [Google Scholar]

886. Karczmarek, J.; Wright, J.; Corkum, P.; Ivanov, M. Optical Centrifuge for Molecules Phys. Rev. Lett. 1999, 82,
3420– 3423 DOI: 10.1103/PhysRevLett.82.3420 [Crossref], [CAS], [Google Scholar]

887. Villeneuve, D. M.; Aseyev, S. A.; Dietrich, P.; Spanner, M.; Ivanov, M. Y.; Corkum, P. B. Forced Molecular
Rotation in an Optical Centrifuge Phys. Rev. Lett. 2000, 85, 542– 545 DOI: 10.1103/PhysRevLett.85.542
[Crossref], [PubMed], [CAS], [Google Scholar]

888. Charlier, J. C.; Michenaud, J.-P. Energetics of Multilayered Carbon Tubules Phys. Rev. Lett. 1993, 70, 1858–
1861 DOI: 10.1103/PhysRevLett.70.1858 [Crossref], [PubMed], [CAS], [Google Scholar]

889. Merkle, R. C. A Proof About Molecular Bearings Nanotechnology 1993, 4, 86– 90


 DOI: 10.1088/0957-4484/4/2/004 [Crossref], [CAS], [Google Scholar]

890. Tuzun, R. E.; Noid, D. W.; Sumpter, B. G. The Dynamics of Molecular Bearings Nanotechnology 1995, 6, 64–
74 DOI: 10.1088/0957-4484/6/2/005 [Crossref], [CAS], [Google Scholar]
891. Han, J.; Globus, A.; Jaffe, R.; Deardorff, G. Molecular Dynamics Simulations of Carbon Nanotube-based
Gears Nanotechnology 1997, 8, 95– 102 DOI: 10.1088/0957-4484/8/3/001 [Crossref], [CAS], [Google  
 Scholar]

892. Sohlberg, K.; Tuzun, R. E.; Sumpter, B. G.; Noid, D. W. Application of Rigid-body dynamics and Semiclassical
Mechanics to Molecular Bearings Nanotechnology 1997, 8, 103– 111 DOI: 10.1088/0957-4484/8/3/002
[Crossref], [CAS], [Google Scholar]

893. Globus, A.; Bauschlicher, C. W., Jr.; Han, J.; Jaffe, R. L.; Levit, C.; Srivastava, D. Machine Phase Fullerene
Nanotechnology Nanotechnology 1998, 9, 192– 199 DOI: 10.1088/0957-4484/9/3/008 [Crossref], [CAS], [Google
Scholar]

894. Kwon, Y.-K.; Tomanek, D. Electronic and Structural Properties of Multiwall Carbon Nanotubes. Phys. Rev. B:
Condens. Matter Mater. Phys. 1998, 58, R16001,  DOI: 10.1103/PhysRevB.58.R16001 . [Crossref], [CAS], [Google
Scholar]

895. Tuzun, R. E.; Sohlberg, K.; Noid, D. W.; Sumpter, B. G. Docking Envelopes for the Assembly of Molecular
Bearings Nanotechnology 1998, 9, 37– 48 DOI: 10.1088/0957-4484/9/1/005 [Crossref], [CAS], [Google Scholar]

896. Palser, A. H. R. Interlayer Interactions in Graphite and Carbon Nanotubes Phys. Chem. Chem. Phys. 1999,
1, 4459– 4464 DOI: 10.1039/a905154f [Crossref], [CAS], [Google Scholar]

897. Cumings, J. Low-Friction Nanoscale Linear Bearing Realized from Multiwall Carbon Nanotubes Science
2000, 289, 602– 604 DOI: 10.1126/science.289.5479.602 [Crossref], [PubMed], [CAS], [Google Scholar]

898. Kolmogorov, A. N.; Crespi, V. H. The Smoothest Bearings: Interlayer Sliding in Multiwalled Carbon
Nanotubes Phys. Rev. Lett. 2000, 85, 4727– 4730 DOI: 10.1103/PhysRevLett.85.4727 [Crossref], [PubMed],
[CAS], [Google Scholar]

899. Yu, M.-F.; Yakobson, B. I.; Ruoff, R. S. Controlled Sliding and Pullout of Nested Shells in Individual
Multiwalled Carbon Nanotubes J. Phys. Chem. B 2000, 104, 8764– 8767 DOI: 10.1021/jp002828d [ACS Full Text
], [CAS], [Google Scholar]

900. Saito, R.; Matsuo, R.; Kimura, T.; Dresselhaus, G.; Dresselhaus, M. S. Anomalous Potential Barrier of Double-
wall Carbon Nanotubes Chem. Phys. Lett. 2001, 348, 187– 193 DOI: 10.1016/S0009-2614(01)01127-7
[Crossref], [CAS], [Google Scholar]
901. Damnjanović, M.; Vuković, T.; Milošević, I. Super-slippery Carbon Nanotubes Eur. Phys. J. B 2002, 25, 131–
134 DOI: 10.1140/epjb/e20020014 [Crossref], [CAS], [Google Scholar]   

902. Zheng, Q.; Jiang, Q. Multiwalled Carbon Nanotubes as Gigahertz Oscillators Phys. Rev. Lett. 2002, 88,
045503 DOI: 10.1103/PhysRevLett.88.045503 [Crossref], [CAS], [Google Scholar]

903. Zheng, Q.; Liu, J.; Jiang, Q. Excess van der Waals Interaction Energy of a Multiwalled Carbon Nanotube
with an Extruded Core and the Induced Core Oscillation Phys. Rev. B: Condens. Matter Mater. Phys. 2002, 65,
245409.1– 245409.6 DOI: 10.1103/PhysRevB.65.245409 [Crossref], [Google Scholar]

904. Damnjanović, M.; Dobardzic, E.; Milosevic, I.; Vukovic, T.; Nikolic, B. Lattice Dynamics and Symmetry of
Double Wall Carbon Nanotubes New J. Phys. 2003, 5, 148 DOI: 10.1088/1367-2630/5/1/148 [Crossref], [Google
Scholar]

905. Guo, W.; Guo, Y.; Gao, H.; Zheng, Q.; Zhong, W. Energy Dissipation in Gigahertz Oscillators from Multiwalled
Carbon Nanotubes Phys. Rev. Lett. 2003, 91, 125501 DOI: 10.1103/PhysRevLett.91.125501 [Crossref], [PubMed],
[CAS], [Google Scholar]

906. Legoas, S.; Coluci, V.; Braga, S.; Coura, P.; Dantas, S.; Galvão, D. Molecular-Dynamics Simulations of Carbon
Nanotubes as Gigahertz Oscillators Phys. Rev. Lett. 2003, 90, 055504 DOI: 10.1103/PhysRevLett.90.055504
[Crossref], [CAS], [Google Scholar]

907. Lozovik, Y. E.; Minogin, A. V.; Popov, A. M. Nanomachines Based on Carbon Nanotubes Phys. Lett. A 2003,
313, 112– 121 DOI: 10.1016/S0375-9601(03)00649-2 [Crossref], [CAS], [Google Scholar]

908. Lozovik, Y. E.; Minogin, A. V.; Popov, A. M. Possible Nanomachines: Nanotube Walls as Movable Elements
JETP Lett. 2003, 77, 631– 635 DOI: 10.1134/1.1600820 [Crossref], [CAS], [Google Scholar]

909. Servantie, J.; Gaspard, P. Methods of Calculation of a Friction Coe cient: Application to Nanotubes Phys.
Rev. Lett. 2003, 91, 185503 DOI: 10.1103/PhysRevLett.91.185503 [Crossref], [PubMed], [CAS], [Google Scholar]

910. Vukovic, T.; Damnjanovic, M.; Milosevic, I. Interaction Between Layers of the Multi-wall Carbon Nanotubes
Phys. E 2003, 16, 259– 268 DOI: 10.1016/S1386-9477(02)00685-9 [Crossref], [CAS], [Google Scholar]

911. Lin, J.-S.; Lin, J.-H.; Chang, C.-C. Molecular Dynamics Simulations of the Rotary Motor F0 Under External
Electric Fields Across the Membrane Biophys. J. 2010, 98, 1009– 1017 DOI: 10.1016/j.bpj.2009.11.025
[Crossref], [PubMed], [CAS], [Google Scholar]

  
912. Williams, P. A.; Papadakis, S. J.; Patel, A. M.; Falvo, M. R.; Washburn, S.; Super ne, R. Fabrication of
Nanometer-scale Mechanical Devices Incorporating Individual Multiwalled Carbon Nanotubes as Torsional
Springs Appl. Phys. Lett. 2003, 82, 805– 907 DOI: 10.1063/1.1538346 [Crossref], [CAS], [Google Scholar]

913. Zhao, Y.; Ma, C.-C.; Chen, G.; Jiang, Q. Energy Dissipation Mechanisms in Carbon Nanotube Oscillators
Phys. Rev. Lett. 2003, 91, 175504 DOI: 10.1103/PhysRevLett.91.175504 [Crossref], [PubMed], [CAS], [Google
Scholar]

914. Belikov, A. V.; Lozovik, Y. E.; Nikolaev, A. G.; Popov, A. M. Double-wall Nanotubes: Classi cation and Barriers
to Walls Relative Rotation, Sliding and Screwlike Motion Chem. Phys. Lett. 2004, 385, 72– 78
 DOI: 10.1016/j.cplett.2003.12.049 [Crossref], [CAS], [Google Scholar]

915. Bourlon, B.; Glattli, D. C.; Miko, C.; Forro, L.; Bachtold, A. Carbon Nanotube Based Bearing for Rotational
Motions Nano Lett. 2004, 4, 709– 712 DOI: 10.1021/nl035217g [ACS Full Text ], [CAS], [Google Scholar]

916. Kang, J. W.; Hwang, H. J. Gigahertz Actuator of Multiwall Marbon Nanotube Encapsulating Metallic Ions:
Molecular Dynamics Simulations J. Appl. Phys. 2004, 96, 3900– 3905 DOI: 10.1063/1.1785837 [Crossref],
[CAS], [Google Scholar]

917. Kang, J. W.; Hwang, H. J. Nanoscale Carbon Nanotube Motor Schematics and Simulations for Micro-
electro-mechanical Machines Nanotechnology 2004, 15, 1633– 1638 DOI: 10.1088/0957-4484/15/11/045
[Crossref], [CAS], [Google Scholar]

918. Legoas, S. B.; Coluci, V. R.; Braga, S. F.; Coura, P. Z.; Dantas, S. O.; Galvão, D. S. Gigahertz Nanomechanical
Oscillators Based on Carbon Nanotubes Nanotechnology 2004, 15, S184– S189
 DOI: 10.1088/0957-4484/15/4/012 [Crossref], [CAS], [Google Scholar]

919. Papadakis, S.; Hall, A.; Williams, P.; Vicci, L.; Falvo, M.; Super ne, R.; Washburn, S. Resonant Oscillators with
Carbon-Nanotube Torsion Springs Phys. Rev. Lett. 2004, 93, 146101 DOI: 10.1103/PhysRevLett.93.146101
[Crossref], [PubMed], [CAS], [Google Scholar]

920. Regan, B. C.; Aloni, S.; Ritchie, R. O.; Dahmen, U.; Zettl, A. Carbon Nanotubes as Nanoscale Mass Conveyors
Nature 2004, 428, 924– 927 DOI: 10.1038/nature02496 [Crossref], [PubMed], [CAS], [Google Scholar]
921. Tangney, P.; Louie, S.; Cohen, M. Dynamic Sliding Friction between Concentric Carbon Nanotubes Phys.
Rev. Lett. 2004, 93, 065530 DOI: 10.1103/PhysRevLett.93.065503 [Crossref], [Google Scholar]   

922. Tu, Z. C.; Ou-Yang, Z. C. A Molecular Motor Constructed from a Double-walled Carbon Nanotube Driven by
Temperature Variation J. Phys.: Condens. Matter 2004, 16, 1287– 1292 DOI: 10.1088/0953-8984/16/8/012
[Crossref], [CAS], [Google Scholar]

923. Bichoutskaia, E.; Popov, A.; El-Barbary, A.; Heggie, M.; Lozovik, Y. Ab initio Study of Relative Motion of Walls
in Carbon Nanotubes Phys. Rev. B: Condens. Matter Mater. Phys. 2005, 71, 113403
 DOI: 10.1103/PhysRevB.71.113403 [Crossref], [CAS], [Google Scholar]

924. Kang, J. W.; Song, K. O.; Kwon, O. K.; Hwang, H. J. Carbon Nanotube Oscillator Operated by Thermal
Expansion of Encapsulated Gases Nanotechnology 2005, 16, 2670– 2676 DOI: 10.1088/0957-4484/16/11/034
[Crossref], [CAS], [Google Scholar]

925. Meyer, J. C.; Paillet, M.; Roth, S. Single-molecule Torsional Pendulum Science 2005, 309, 1539– 1541
 DOI: 10.1126/science.1115067 [Crossref], [PubMed], [CAS], [Google Scholar]

926. Regan, B. C.; Aloni, S.; Jensen, K.; Ritchie, R. O.; Zettl, A. Nanocrystal-Powered Nanomotor Nano Lett. 2005,
5, 1730– 1733 DOI: 10.1021/nl0510659 [ACS Full Text ], [CAS], [Google Scholar]

927. Tu, Z.; Hu, X. Molecular Motor Constructed From a Double-walled Carbon Nanotube Driven by Axially
Varying Voltage Phys. Rev. B: Condens. Matter Mater. Phys. 2005, 72, 033404
 DOI: 10.1103/PhysRevB.72.033404 [Crossref], [CAS], [Google Scholar]

928. Tasis, D.; Tagmatarchis, N.; Bianco, A.; Prato, M. Chemistry of Carbon Nanotubes Chem. Rev. 2006, 106,
1105– 1136 DOI: 10.1021/cr050569o [ACS Full Text ], [CAS], [Google Scholar]

929. Yuzvinsky, T. D.; Fennimore, A. M.; Kis, A.; Zettl, A. Controlled Placement of Highly Aligned Carbon
Nanotubes for the Manufacture of Arrays of Nanoscale Torsional Actuators Nanotechnology 2006, 17, 434– 438
 DOI: 10.1088/0957-4484/17/2/015 [Crossref], [CAS], [Google Scholar]

930. Li, R.; Sun, D.; Zhang, B. Motion and Energy Dissipation of Single-walled Carbon Nanotube on Graphite by
Molecular Dynamics Simulation Mater. Res. Express 2014, 1, 025046 DOI: 10.1088/2053-1591/1/2/025046
[Crossref], [CAS], [Google Scholar]
931. Bedard, T. C.; Moore, J. S. Design and Synthesis of Molecular Turnstiles J. Am. Chem. Soc. 1995, 117,
10662– 10671 DOI: 10.1021/ja00148a008 [ACS Full Text ], [CAS], [Google Scholar]   

932. Hsu, L. Y.; Li, E. Y.; Rabitz, H. Single-molecule Electric Revolving Door Nano Lett. 2013, 13, 5020– 5025
 DOI: 10.1021/nl401340c [ACS Full Text ], [CAS], [Google Scholar]

933. Troisi, A.; Ratner, M. Molecular Recti cation Through Electric Field Induced Conformational Changes J.
Am. Chem. Soc. 2002, 124, 14528– 14529 DOI: 10.1021/ja028281t [ACS Full Text ], [CAS], [Google Scholar]

934. Troisi, A.; Ratner, M. A. Conformational Molecular Recti ers Nano Lett. 2004, 4, 591– 595
 DOI: 10.1021/nl0352088 [ACS Full Text ], [CAS], [Google Scholar]

935. Goychuk, I.; Haenggi, P. Quantum Recti ers From Harmonic Mixing Europhys. Lett. 1998, 43, 503– 509
 DOI: 10.1209/epl/i1998-00389-2 [Crossref], [CAS], [Google Scholar]

936. Kornilovitch, P.; Bratkovsky, A.; Williams, R. Bistable Molecular Conductors with a Field-switchable Dipole
Group Phys. Rev. B: Condens. Matter Mater. Phys. 2002, 66, 245413 DOI: 10.1103/PhysRevB.66.245413
[Crossref], [CAS], [Google Scholar]

937. Ghosh, A. W.; Rakshit, T.; Datta, S. Gating of a Molecular Transistor: Electrostatic and Conformational Nano
Lett. 2004, 4, 565– 568 DOI: 10.1021/nl035109u [ACS Full Text ], [CAS], [Google Scholar]

938. Tuzun, R. E.; Noid, D. W.; Sumpter, B. G. Dynamics of a Laser Driven Molecular Motor Nanotechnology
1995, 6, 52– 63 DOI: 10.1088/0957-4484/6/2/004 [Crossref], [CAS], [Google Scholar]

939. Hoki, K.; Gonzalez, L.; Fujimura, Y. Control of Molecular Handedness Using Pump-dump Laser Pulses J.
Chem. Phys. 2002, 116, 2433– 2438 DOI: 10.1063/1.1432996 [Crossref], [CAS], [Google Scholar]

940. Hoki, K.; Yamaki, M.; Fujimura, Y. Chiral Molecular Motors Driven by a Nonhelical Laser Pulse Angew.
Chem., Int. Ed. 2003, 42, 2976– 2978 DOI: 10.1002/anie.200250872 [Crossref], [PubMed], [CAS], [Google Scholar]

941. Hoki, K.; Yamaki, M.; Koseki, S.; Fujimura, Y. Mechanism of Unidirectional Motions of Chiral Molecular
Motors Driven by Linearly Polarized Pulses J. Chem. Phys. 2003, 119, 12393– 12398 DOI: 10.1063/1.1621622
[Crossref], [CAS], [Google Scholar]
942. Hoki, K.; Yamaki, M.; Koseki, S.; Fujimura, Y. Molecular Motors Driven by Laser Pulses: Role of Molecular
Chirality and Photon Helicity J. Chem. Phys. 2003, 118, 497– 504 DOI: 10.1063/1.1526834 [Crossref],
  
[CAS], [Google Scholar]

943. Hoki, K.; Sato, M.; Yamaki, M.; Sahnoun, R.; Gonzalez, L.; Koseki, S.; Fujimura, Y. Chiral Molecular Motors
Ignited by Femtosecond Pump-dump Laser Pulses J. Phys. Chem. B 2004, 108, 4916– 4921
 DOI: 10.1021/jp036437l [ACS Full Text ], [CAS], [Google Scholar]

944. Kono, H.; Sato, Y.; Tanaka, N.; Kato, T.; Nakai, K.; Koseki, S.; Fujimura, Y. Quantum Mechanical Study of
Electronic and Nuclear Dynamics of Molecules in Intense Laser Fields Chem. Phys. 2004, 304, 203– 226
 DOI: 10.1016/j.chemphys.2004.04.017 [Crossref], [CAS], [Google Scholar]

945. Hoki, K.; Sato, M.; Yamaki, M.; Sahnoun, R.; Gonzalez, L.; Koseki, S.; Fujimura, Y. Chiral Molecular Motors
Ignited by Femtosecond Pump–Dump Laser Pulses J. Phys. Chem. B 2004, 108, 4916– 4921
 DOI: 10.1021/jp036437l [ACS Full Text ], [CAS], [Google Scholar]

946. Yamaki, M.; Hoki, K.; Ohtsuki, Y.; Kono, H.; Fujimura, Y. Quantum Control of Unidirectional Rotations of a
Chiral Molecular Motor Phys. Chem. Chem. Phys. 2005, 7, 900– 904 DOI: 10.1039/b418231f [Crossref],
[PubMed], [Google Scholar]

947. Yamaki, M.; Hoki, K.; Ohtsuki, Y.; Kono, H.; Fujimura, Y. Quantum Control of a Chiral Molecular Motor Driven
by Laser Pulses J. Am. Chem. Soc. 2005, 127, 7300– 7301 DOI: 10.1021/ja0437757 [ACS Full Text ],
[CAS], [Google Scholar]

948. Yamaki, M.; Hoki, K.; Kono, H.; Fujimura, Y. Quantum Control of a Chiral Molecular Motor Driven by
Femtosecond Laser Pulses: Mechanisms of Regular and Reverse Rotations Chem. Phys. 2008, 347, 272– 274
 DOI: 10.1016/j.chemphys.2007.11.016 [Crossref], [CAS], [Google Scholar]

949. Yamaki, M.; Hoki, K.; Teranishi, T.; Chung, W.; Pichierri, F.; Kono, H.; Fujimura, Y. Theoretical Design of an
Aromatic Hydrocarbon Rotor Driven by a Circularly Polarized Electric Field J. Phys. Chem. A 2007, 111, 9374–
9378 DOI: 10.1021/jp073953t [ACS Full Text ], [CAS], [Google Scholar]

950. Yamaki, M.; Nakayama, S.; Hoki, K.; Kono, H.; Fujimura, Y. Quantum Dynamics of Light-driven Chiral
Molecular Motors Phys. Chem. Chem. Phys. 2009, 11, 1662– 1678 DOI: 10.1039/b815047h [Crossref],
[PubMed], [CAS], [Google Scholar]
951. Fujimura, Y.; González, L.; Kröner, D.; Manz, J.; Mehdaoui, I.; Schmidt, B. Quantum Ignition of Intramolecular
Rotation by Means of IR+UV Laser Pulses Chem. Phys. Lett. 2004, 386, 248– 253   
 DOI: 10.1016/j.cplett.2004.01.070 [Crossref], [CAS], [Google Scholar]

952. Jian, H.; Tour, J. En Route to Surface-bound Electric Field-driven Molecular Motors J. Org. Chem. 2003, 68,
5091– 5103 DOI: 10.1021/jo034169h [ACS Full Text ], [CAS], [Google Scholar]

953. Caskey, D.; Shoemaker, R.; Michl, J. Toward Self-assembled Surface-mounted Prismatic Altitudinal Rotors.
A Test Case: Molecular Rectangle Org. Lett. 2004, 6, 2093– 2096 DOI: 10.1021/ol049539i [ACS Full Text ],
[CAS], [Google Scholar]

954. Zheng, X.; Mulcahy, M.; Horinek, D.; Galeotti, F.; Magnera, T.; Michl, J. Dipolar and Nonpolar Altitudinal
Molecular Rotors Mounted on an Au(111) Surface J. Am. Chem. Soc. 2004, 126, 4540– 4542
 DOI: 10.1021/ja039482f [ACS Full Text ], [CAS], [Google Scholar]

955. Magnera, T.; Michl, J.; Kelly, T. Altitudinal Surface-mounted Molecular Rotors Molecular Machines 2005,
262, 63– 97 DOI: 10.1007/128_014 [Crossref], [Google Scholar]

956. Caskey, D.; Michl, J. Toward Self-assembled Surface-mounted Prismatic Altitudinal Rotors. A Test Case:
Trigonal and Tetragonal Prisms J. Org. Chem. 2005, 70, 5442– 5448 DOI: 10.1021/jo050409c [ACS Full Text ],
[CAS], [Google Scholar]

957. Kottas, G.; Clarke, L.; Horinek, D.; Michl, J. Arti cial Molecular Rotors Chem. Rev. 2005, 105, 1281– 1376
 DOI: 10.1021/cr0300993 [ACS Full Text ], [CAS], [Google Scholar]

958. Mulcahy, M.; Magnera, T.; Michl, J. Molecular Rotors on Au(111): Rotator Orientation from IR Spectroscopy
J. Phys. Chem. C 2009, 113, 20698– 20704 DOI: 10.1021/jp906809b [ACS Full Text ], [CAS], [Google Scholar]

959. Mulcahy, M.; Bastl, Z.; Stensrud, K.; Magnera, T.; Michl, J. Mercury-Mediated Attachment of Metal-
Sandwich-Based Altitudinal Molecular Rotors to Gold Surfaces J. Phys. Chem. C 2010, 114, 14050– 14060
 DOI: 10.1021/jp100923f [ACS Full Text ], [CAS], [Google Scholar]

960. Kobr, L.; Zhao, K.; Shen, Y.; Comotti, A.; Bracco, S.; Shoemaker, R.; Sozzani, P.; Clark, N.; Price, J.; Rogers, C.,
et al. Inclusion Compound Based Approach to Arrays of Arti cial Dipolar Molecular Rotors. A Surface Inclusion J.
Am. Chem. Soc. 2012, 134, 10122– 10131 DOI: 10.1021/ja302173y [ACS Full Text ], [CAS], [Google Scholar]
961. Casher, D. L.; Kobr, L.; Michl, J. Average Orientation of a Molecular Rotor Embedded in a Langmuir-Blodgett
Monolayer Langmuir 2012, 28, 1625– 1637 DOI: 10.1021/la2037789 [ACS Full Text ], [CAS], [Google  
 Scholar]

962. Kobr, L.; Zhao, K.; Shen, Y.; Polivkova, K.; Shoemaker, R.; Clark, N.; Price, J.; Rogers, C.; Michl, J. Inclusion
Compound Based Approach to Arrays of Arti cial Dipolar Molecular Rotors: Bulk Inclusions J. Org. Chem. 2013,
78, 1768– 1777 DOI: 10.1021/jo3009897 [ACS Full Text ], [CAS], [Google Scholar]

963. Clarke, L. I.; Horinek, D.; Kottas, G. S.; Varaksa, N.; Magnera, T. F.; Hinderer, T. P.; Horansky, R. D.; Michl, J.;
Price, J. C. The Dielectric Response of Chloromethylsilyl and Dichloromethylsilyl Dipolar Rotors on Fused Silica
Surfaces Nanotechnology 2002, 13, 533– 540 DOI: 10.1088/0957-4484/13/4/317 [Crossref], [CAS], [Google
Scholar]

964. Michl, J.; Magnera, T. F. Two-dimensional Supramolecular Chemistry with Molecular Tinkertoys Proc. Natl.
Acad. Sci. U. S. A. 2002, 99, 4788– 4792 DOI: 10.1073/pnas.052016299 [Crossref], [PubMed], [CAS], [Google
Scholar]

965. Horinek, D.; Michl, J. Molecular Dynamics Simulation of an Electric Field Driven Dipolar Molecular Rotor
Attached to a Quartz Glass Surface J. Am. Chem. Soc. 2003, 125, 11900– 11910 DOI: 10.1021/ja0348851 [ACS
Full Text ], [CAS], [Google Scholar]

966. Horinek, D.; Michl, J. Surface-mounted Altitudinal Molecular Rotors in Alternating Electric Field: Single-
molecule Parametric Oscillator Molecular Dynamics Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 14175– 14180
 DOI: 10.1073/pnas.0506183102 [Crossref], [PubMed], [CAS], [Google Scholar]

967. Vacek, J.; Michl, J. Molecular Dynamics of a Grid-mounted Molecular Dipolar Rotor in a Rotating Electric
Field Proc. Natl. Acad. Sci. U. S. A. 2001, 98, 5481– 5486 DOI: 10.1073/pnas.091100598 [Crossref], [PubMed],
[CAS], [Google Scholar]

968. Vacek, J.; Michl, J. Arti cial Surface-mounted Molecular Rotors: Molecular Dynamics Simulations Adv.
Funct. Mater. 2007, 17, 730– 739 DOI: 10.1002/adfm.200601225 [Crossref], [CAS], [Google Scholar]

969. Prokop, A.; Vacek, J.; Michl, J. Friction in Carborane-Based Molecular Rotors Driven by Gas Flow or Electric
Field: Classical Molecular Dynamics ACS Nano 2012, 6, 1901– 1914 DOI: 10.1021/nn300003x [ACS Full Text ],
[CAS], [Google Scholar]

970. Dunitz, J. D.; Maverick, E. F.; Trueblood, K. N. Atomic Motions in Molecular Crystals from Diffraction
Measurements Angew. Chem., Int. Ed. Engl. 1988, 27, 880– 895 DOI: 10.1002/anie.198808801
[Crossref], [Google Scholar]

  
971. Akutagawa, T.; Shitagami, K.; Nishihara, S.; Takeda, S.; T, H.; Nakamura, T.; Hosokoshi, Y.; Inoue, K.; Ikeuchi,
S.; Miyazaki, Y.; Saito, K. Molecular Rotor of Cs2([18]crown-6)3 in the Solid State Coupled with the Magnetism of
[Ni(dmit)2] J. Am. Chem. Soc. 2005, 127, 4397– 4402 DOI: 10.1021/ja043527a [ACS Full Text ], [CAS], [Google
Scholar]

972. Khuong, T.; Nunez, J.; Godinez, C.; Garcia-Garibay, M. Crystalline Molecular Machines: A Quest Toward
Solid-state Dynamics and Function Acc. Chem. Res. 2006, 39, 413– 422 DOI: 10.1021/ar0680217 [ACS Full Text
], [CAS], [Google Scholar]

973. Karlen, S.; Reyes, H.; Taylor, R.; Khan, S.; Hawthorne, M.; Garcia-Garibay, M. Symmetry and Dynamics of
Molecular Rotors in Amphidynamic Molecular Cystals Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 14973– 14977
 DOI: 10.1073/pnas.1008213107 [Crossref], [PubMed], [CAS], [Google Scholar]

974. Lemouchi, C.; Iliopoulos, K.; Zorina, L.; Simonov, S.; Wzietek, P.; Cauchy, T.; Rodríguez-Fortea, A.; Canadell,
E.; Kaleta, J.; Michl, J., et al. Crystalline Arrays of Pairs of Molecular Rotors: Correlated Motion, Rotational
Barriers, and Space-Inversion Symmetry Breaking Due to Conformational Mutations J. Am. Chem. Soc. 2013,
135, 9366– 9376 DOI: 10.1021/ja4044517 [ACS Full Text ], [CAS], [Google Scholar]

975. Zhang, Q.-C.; Wu, F.-T.; Hao, H.-M.; Xu, H.; Zhao, H.-X.; Long, L.-S.; Huang, R.-B.; Zheng, L.-S. Modulating the
Rotation of a Molecular Rotor through Hydrogen-Bonding Interactions between the Rotator and Stator Angew.
Chem., Int. Ed. 2013, 52, 12602– 12605 DOI: 10.1002/anie.201306193 [Crossref], [PubMed], [CAS], [Google
Scholar]

976. Karlen, S. D.; Garcia-Garibay, M. A. Amphidynamic Crystals: Structural Blueprints for Molecular Machines
2005, 262, 179– 227 DOI: 10.1007/128_012 [Crossref], [Google Scholar]

977. Shima, T.; Hampel, F.; Gladysz, J. A. Molecular gyroscopes: [Fe(CO)3] and [Fe(CO)2(NO)]+ rotators encased
in three-spoke stators; facile assembly by alkene metatheses Angew. Chem., Int. Ed. 2004, 43, 5537– 5540
 DOI: 10.1002/anie.200460534 [Crossref], [PubMed], [CAS], [Google Scholar]

978. Skopek, K.; Hershberger, M.; Gladysz, J. Gyroscopes and the chemical literature: 1852–2002 Coord. Chem.
Rev. 2007, 251, 1723– 1733 DOI: 10.1016/j.ccr.2006.12.015 [Crossref], [CAS], [Google Scholar]

979. Commins, P.; Nunez, J.; Garcia-Garibay, M. Synthesis of Bridged Molecular Gyroscopes with Closed
Topologies: Triple One-Pot Macrocyclization J. Org. Chem. 2011, 76, 8355– 8363 DOI: 10.1021/jo201513y [ACS
Full Text ], [CAS], [Google Scholar]

  
980. Karlen, S. D.; Godinez, C. E.; Garcia-Garibay, M. A. Improved physical properties and rotational dynamics in
a molecular gyroscope with an asymmetric stator structure Org. Lett. 2006, 8, 3417– 3420
 DOI: 10.1021/ol060894d [ACS Full Text ], [CAS], [Google Scholar]

981. Horansky, R.; Clarke, L.; Winston, E.; Price, J.; Karlen, S.; Jarowski, P.; Santillan, R.; Garcia-Garibay, M. Dipolar
rotor-rotor interactions in a di uorobenzene molecular rotor crystal Phys. Rev. B: Condens. Matter Mater. Phys.
2006, 74, 054306 DOI: 10.1103/PhysRevB.74.054306 [Crossref], [CAS], [Google Scholar]

982. Setaka, W.; Yamaguchi, K. Order-disorder transition of dipolar rotor in a crystalline molecular gyrotop and
its optical change J. Am. Chem. Soc. 2013, 135, 14560– 14563 DOI: 10.1021/ja408405f [ACS Full Text ],
[CAS], [Google Scholar]

983. Shima, T.; Hampel, F.; Gladysz, J. Molecular gyroscopes: {Fe(CO)3} and {Fe(CO)2(NO)}+ rotators encased in
three-spoke stators; facile assembly by alkene metatheses Angew. Chem., Int. Ed. 2004, 43, 5537– 5540
 DOI: 10.1002/anie.200460534 [Crossref], [PubMed], [CAS], [Google Scholar]

984. Zeits, P.; Rachiero, G.; Harnpel, F.; Reibenspies, J.; Gladysz, J. Gyroscope-Like Platinum and Palladium
Complexes with Trans-Spanning Bis(pyridine) Ligands Organometallics 2012, 31, 2854– 2877
 DOI: 10.1021/om201145x [ACS Full Text ], [CAS], [Google Scholar]

985. Dominguez, Z.; Dang, H.; Strouse, M. J.; Garcia-Garibay, M. A. Molecular “Compasses” and “Gyroscopes.” III.
Dynamics of a Phenylene Rotor and Clathrated Benzene in a Slipping-Gear Crystal Lattice J. Am. Chem. Soc.
2002, 124, 7719– 7727 DOI: 10.1021/ja025753v [ACS Full Text ], [CAS], [Google Scholar]

986. Dominguez, Z.; Dang, H.; Strouse, M.; Garcia-Garibay, M. Molecular ″compasses″ and ″gyroscopes″. I.
Expedient synthesis and solid state dynamics of an open rotor with a bis(triarylmethyl) frame J. Am. Chem. Soc.
2002, 124, 2398– 2399 DOI: 10.1021/ja0119447 [ACS Full Text ], [CAS], [Google Scholar]

987. Marahatta, A. B.; Kanno, M.; Hoki, K.; Setaka, W.; Irle, S., II; Kono, H. Theoretical Investigation of the
Structures and Dynamics of Crystalline Molecular Gyroscopes J. Phys. Chem. C 2012, 116, 24845– 24854
 DOI: 10.1021/jp308974j [ACS Full Text ], [CAS], [Google Scholar]

988. Dominguez, Z.; Khuong, T.; Dang, H.; Sanrame, C.; Nunez, J.; Garcia-Garibay, M. Molecular compasses and
gyroscopes with polar rotors: Synthesis and characterization of crystalline forms J. Am. Chem. Soc. 2003, 125,
8827– 8837 DOI: 10.1021/ja035274b [ACS Full Text ], [CAS], [Google Scholar]
 
 parts:
989. Godinez, C.; Zepeda, G.; Mortko, C.; Dang, H.; Garcia-Garibay, M. Molecular crystals with moving
Synthesis, characterization, and crystal packing of molecular gyroscopes with methyl-substituted triptycyl
frames J. Org. Chem. 2004, 69, 1652– 1662 DOI: 10.1021/jo035517i [ACS Full Text ], [CAS], [Google Scholar]

990. Khuong, T.; Zepeda, G.; Ruiz, R.; Khan, S.; Garcia-Garibay, M. Molecular compasses and gyroscopes:
Engineering molecular crystals with fast internal rotation Cryst. Growth Des. 2004, 4, 15– 18
 DOI: 10.1021/cg034144b [ACS Full Text ], [CAS], [Google Scholar]

991. Binhi, V. N.; Savin, A. V. Molecular Gyroscopes and Biological Effects of Weak Extremely Low-Frequency
Magnetic Fields Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 2002, 65, 051912
 DOI: 10.1103/PhysRevE.65.051912 [Crossref], [Google Scholar]

992. Karlen, S.; Ortiz, R.; Chapman, O.; Garcia-Garibay, M. Effects of rotational symmetry order on the solid state
dynamics of phenylene and diamantane rotators J. Am. Chem. Soc. 2005, 127, 6554– 6555
 DOI: 10.1021/ja042512+ [ACS Full Text ], [CAS], [Google Scholar]

993. Gardinier, J. R.; Pellechia, P. J.; Smith, M. D. Ionic rotors. preparation, structure, and dynamic solid-state 2D
NMR study of the 1,4-diethynylbenzenebis(triphenylborate) dianion J. Am. Chem. Soc. 2005, 127, 12448– 12449
 DOI: 10.1021/ja053256j [ACS Full Text ], [CAS], [Google Scholar]

994. Karlen, S. D.; Garcia-Garibay, M. A. Highlighting gyroscopic motion in crystals in 13C CPMAS spectra by
speci c isotopic substitution and restricted cross polarization Chem. Commun. 2005, 189– 191
 DOI: 10.1039/b409744k [Crossref], [PubMed], [CAS], [Google Scholar]

995. Karlen, S. D.; Khan, S. I.; Garcia-Garibay, M. A. Removal of Con icting Molecular Symmetries Restores a
HexagonalArray of Six-Fold Phenyl Embraces in a bis(Trityl)-Containing Compound. I. Crystals of 1,1,1,6,6,6-
Hexaphenyl-2,4-hexadiyne Cryst. Growth Des. 2005, 5, 53– 55 DOI: 10.1021/cg049898k [ACS Full Text ],
[CAS], [Google Scholar]

996. Karlen, S.; Khan, S.; Garcia-Garibay, M. Crystalline molecular gyroscopes: The effects of subtle molecular
differences on the crystal packing of triphenylmethyl and triphenylsilyl stators Mol. Cryst. Liq. Cryst. 2006, 456,
221– 230 DOI: 10.1080/15421400600788757 [Crossref], [CAS], [Google Scholar]

997. Nunez, J.; Khuong, T.; Campos, L.; Farfan, N.; Dang, H.; Karlen, S.; Garcia-Garibay, M. Crystal phases and
phase transitions in a highly polymorphogenic solid-state molecular gyroscope with meta-methoxytrityl frames
Cryst. Growth Des. 2006, 6, 866– 873 DOI: 10.1021/cg050155o [ACS Full Text ], [CAS], [Google Scholar]

998. Garcia-Garibay, M.; Godinez, C. Engineering Crystal Packing and Internal Dynamics in Molecular  
Gyroscopes by Re ning their Components. Fast Exchange of a Phenylene Rotator by H-2 NMR Cryst. Growth
Des. 2009, 9, 3124– 3128 DOI: 10.1021/cg801065a [ACS Full Text ], [CAS], [Google Scholar]

999. Rodriguez-Molina, B.; Ochoa, M.; Farfan, N.; Santillan, R.; Garcia-Garibay, M. Synthesis, Characterization,
and Rotational Dynamics of Crystalline Molecular Compasses with N-Heterocyclic Rotators J. Org. Chem. 2009,
74, 8554– 8565 DOI: 10.1021/jo901261j [ACS Full Text ], [CAS], [Google Scholar]

1000. O’Brien, Z.; Natarajan, A.; Khan, S.; Garcia-Garibay, M. Synthesis and Solid-State Rotational Dynamics of
Molecular Gyroscopes with a Robust and Low Density Structure Built with a Phenylene Rotator and a Tri(meta-
terphenyl)methyl Stator Cryst. Growth Des. 2011, 11, 2654– 2659 DOI: 10.1021/cg200373g [ACS Full Text ],
[CAS], [Google Scholar]

1001. O’Brien, Z.; Karlen, S.; Khan, S.; Garcia-Garibay, M. Solid-State Molecular Rotors with Perdeuterated
Stators: Mechanistic Insights from Biphenylene Rotational Dynamics in Ordered and Disordered Crystal Forms J.
Org. Chem. 2010, 75, 2482– 2491 DOI: 10.1021/jo9025176 [ACS Full Text ], [CAS], [Google Scholar]

1002. Escalante-Sanchez, E.; Rodriguez-Molina, B.; Garcia-Garibay, M. Toward Crystalline Molecular Rotors with
Linearly Conjugated Diethynyl-Phenylene Rotators and Pentiptycene Stators J. Org. Chem. 2012, 77, 7428– 7434
 DOI: 10.1021/jo301223q [ACS Full Text ], [CAS], [Google Scholar]

1003. Horansky, R.; Clarke, L.; Price, J.; Khuong, T.; Jarowski, P.; Garcia-Garibay, M. Dielectric response of a
dipolar molecular rotor crystal. Phys. Rev. B: Condens. Matter Mater. Phys. 2005, 72,
 DOI: 10.1103/PhysRevB.72.014302 . [Crossref], [Google Scholar]

1004. Ozin, G. A.; Manners, I.; Fournier-Bidoz, S.; Arsenault, A. Dream Nanomachines Adv. Mater. 2005, 17,
3011– 3018 DOI: 10.1002/adma.200501767 [Crossref], [CAS], [Google Scholar]

1005. Ebbens, S. J.; Howse, J. R. In pursuit of propulsion at the nanoscale Soft Matter 2010, 6, 726– 738
 DOI: 10.1039/b918598d [Crossref], [CAS], [Google Scholar]

1006. Young, N. O.; G, J. S.; Block, M. J. The motion of bubbles in a vertical temperature gradient J. Fluid Mech.
1959, 6, 350– 356 DOI: 10.1017/S0022112059000684 [Crossref], [Google Scholar]
1007. Scriven, L. E.; Sternling, C. V. The Marangoni Effects Nature 1960, 187, 186– 188 DOI: 10.1038/187186a0
[Crossref], [Google Scholar]   

1008. Cazabat, A. M.; Heslot, F.; Troian, S. M.; Carles, P. Fingering instability of thin spreading lms driven by
temperature gradients Nature 1990, 346, 824– 826 DOI: 10.1038/346824a0 [Crossref], [CAS], [Google Scholar]

1009. Barton, K. D.; Shankar Subramanian, R. The migration of liquid drops in a vertical temperature gradient J.
Colloid Interface Sci. 1989, 133, 211– 222 DOI: 10.1016/0021-9797(89)90294-4 [Crossref], [CAS], [Google
Scholar]

1010. Burns, M. A.; Mastrangelo, C. H.; Sammarco, T. S.; Man, F. P.; Webster, J. R.; Johnsons, B. N.; Foerster, B.;
Jones, D.; Fields, Y.; Kaiser, A. R.; Burke, D. T. Microfabricated structures for integrated DNA analysis Proc. Natl.
Acad. Sci. U. S. A. 1996, 93, 5556– 5561 DOI: 10.1073/pnas.93.11.5556 [Crossref], [PubMed], [CAS], [Google
Scholar]

1011. Kataoka, D. E.; Troian, S. M. Patterning liquid ow on the microscopic scale Nature 1999, 402, 794– 797
 DOI: 10.1038/45521 [Crossref], [CAS], [Google Scholar]

1012. Kumar, G.; Prabhu, K. N. Review of non-reactive and reactive wetting of liquids on surfaces Adv. Colloid
Interface Sci. 2007, 133, 61– 89 DOI: 10.1016/j.cis.2007.04.009 [Crossref], [PubMed], [CAS], [Google Scholar]

1013. Gennes, P. G. d. Forced wetting by a reactive uid Europhys. Lett. 1997, 39, 407– 412
 DOI: 10.1209/epl/i1997-00369-6 [Crossref], [Google Scholar]

1014. de Gennes, P. G. The dynamics of reactive wetting on solid surfaces Phys. A 1998, 249, 196– 205
 DOI: 10.1016/S0378-4371(97)00466-4 [Crossref], [CAS], [Google Scholar]

1015. Bain, C. D.; Burnett-Hall, G. D.; Montgomerie, R. R. Rapid motion of liquid drops Nature 1994, 372, 414–
415 DOI: 10.1038/372414a0 [Crossref], [CAS], [Google Scholar]

1016. Dos Santos, F. D.; Ondarçuhu, T. Free-Running Droplets Phys. Rev. Lett. 1995, 75, 2972– 2975
 DOI: 10.1103/PhysRevLett.75.2972 [Crossref], [PubMed], [CAS], [Google Scholar]

1017. Lee, S.-W.; Laibinis, P. E. Directed Movement of Liquids on Patterned Surfaces Using Noncovalent
Molecular Adsorption J. Am. Chem. Soc. 2000, 122, 5395– 5396 DOI: 10.1021/ja994076a [ACS Full Text ],
[CAS], [Google Scholar]
73,
1998,
1018. Mitsumata, T.; Ikeda, K.; Gong, J. P.; Osada, Y. Solvent-driven chemical motor Appl. Phys. Lett.
2366– 2368 DOI: 10.1063/1.122505 [Crossref], [CAS], [Google Scholar]

1019. Ismagilov, R. F.; Schwartz, A.; Bowden, N.; Whitesides, G. M. Autonomous Movement and Self-Assembly
Angew. Chem., Int. Ed. 2002, 41, 652– 654
 DOI: 10.1002/1521-3773(20020215)41:4<652::AID-ANIE652>3.0.CO;2-U [Crossref], [CAS], [Google Scholar]

1020. Paxton, W. F.; Kistler, K. C.; Olmeda, C. C.; Sen, A.; St. Angelo, S. K.; Cao, Y.; Mallouk, T. E.; Lammert, P. E.;
Crespi, V. H. Catalytic Nanomotors: Autonomous Movement of Striped Nanorods J. Am. Chem. Soc. 2004, 126,
13424– 13431 DOI: 10.1021/ja047697z [ACS Full Text ], [CAS], [Google Scholar]

1021. Golestanian, R.; Liverpool, T. B.; Ajdari, A. Propulsion of a Molecular Machine by Asymmetric Distribution
of Reaction Products Phys. Rev. Lett. 2005, 94, 220801 DOI: 10.1103/PhysRevLett.94.220801 [Crossref],
[PubMed], [CAS], [Google Scholar]

1022. Dhar, P.; Fischer, T. M.; Wang, Y.; Mallouk, T. E.; Paxton, W. F.; Sen, A. Autonomously Moving Nanorods at a
Viscous Interface Nano Lett. 2006, 6, 66– 72 DOI: 10.1021/nl052027s [ACS Full Text ], [CAS], [Google Scholar]

1023. Paxton, W. F.; Sen, A.; Mallouk, T. E. Motility of Catalytic Nanoparticles through Self-Generated Forces
Chem. - Eur. J. 2005, 11, 6462– 6470 DOI: 10.1002/chem.200500167 [Crossref], [PubMed], [CAS], [Google
Scholar]

1024. Kline, T. R.; Paxton, W. F.; Wang, Y.; Velegol, D.; Mallouk, T. E.; Sen, A. Catalytic Micropumps: Microscopic
Convective Fluid Flow and Pattern Formation J. Am. Chem. Soc. 2005, 127, 17150– 17151
 DOI: 10.1021/ja056069u [ACS Full Text ], [CAS], [Google Scholar]

1025. Mano, N.; Heller, A. Bioelectrochemical Propulsion J. Am. Chem. Soc. 2005, 127, 11574– 11575
 DOI: 10.1021/ja053937e [ACS Full Text ], [CAS], [Google Scholar]

1026. Catchmark, J. M.; Subramanian, S.; Sen, A. Directed Rotational Motion of Microscale Objects Using
Interfacial Tension Gradients Continually Generated via Catalytic Reactions Small 2005, 1, 202– 206
 DOI: 10.1002/smll.200400061 [Crossref], [PubMed], [CAS], [Google Scholar]

1027. Fournier-Bidoz, S.; Arsenault, A. C.; Manners, I.; Ozin, G. A. Synthetic self-propeled nanorotors Chem.
Commun. 2005, 441– 443 DOI: 10.1039/b414896g [Crossref], [PubMed], [CAS], [Google Scholar]
1028. Valadares, L. F.; Tao, Y.-G.; Zacharia, N. S.; Kitaev, V.; Galembeck, F.; Kapral, R.; Ozin, G. A. Catalytic
Nanomotors: Self-Propeled Sphere Dimers Small 2010, 6, 565– 572 DOI: 10.1002/smll.200901976 
[Crossref], 
[PubMed], [CAS], [Google Scholar]

1029. Qin, L.; Banholzer, M. J.; Xu, X.; Huang, L.; Mirkin, C. A. Rational Design and Synthesis of Catalytically
Driven Nanorotors J. Am. Chem. Soc. 2007, 129, 14870– 14871 DOI: 10.1021/ja0772391 [ACS Full Text ],
[CAS], [Google Scholar]

1030. Wang, Y.; Fei, S.-t.; Byun, Y.-M.; Lammert, P. E.; Crespi, V. H.; Sen, A.; Mallouk, T. E. Dynamic Interactions
between Fast Microscale Rotors J. Am. Chem. Soc. 2009, 131, 9926– 9927 DOI: 10.1021/ja904827j [ACS Full
Text ], [CAS], [Google Scholar]

1031. Vicario, J.; Eelkema, R.; Browne, W. R.; Meetsma, A.; La Crois, R. M.; Feringa, B. L. Catalytic molecular
motors: fuelling autonomous movement by a surface bound synthetic manganese catalase Chem. Commun.
2005, 3936– 3938 DOI: 10.1039/b505092h [Crossref], [PubMed], [CAS], [Google Scholar]

1032. Lee, T.-C.; Alarcón-Correa, M.; Miksch, C.; Hahn, K.; Gibbs, J. G.; Fischer, P. Self-Propeling Nanomotors in
the Presence of Strong Brownian Forces Nano Lett. 2014, 14, 2407– 2412 DOI: 10.1021/nl500068n [ACS Full
Text ], [CAS], [Google Scholar]

1033. Pantarotto, D.; Browne, W. R.; Feringa, B. L. Autonomous propulsion of carbon nanotubes powered by a
multienzyme ensemble Chem. Commun. 2008, 1533– 1535 DOI: 10.1039/B715310D [Crossref], [PubMed],
[CAS], [Google Scholar]

1034. Gao, W.; Feng, X.; Pei, A.; Gu, Y.; Li, J.; Wang, J. Seawater-driven magnesium based Janus micromotors
for environmental remediation Nanoscale 2013, 5, 4696– 4700 DOI: 10.1039/c3nr01458d [Crossref], [PubMed],
[CAS], [Google Scholar]

1035. Pavlick, R. A.; Sengupta, S.; McFadden, T.; Zhang, H.; Sen, A. A Polymerization-Powered Motor Angew.
Chem., Int. Ed. 2011, 50, 9374– 9377 DOI: 10.1002/anie.201103565 [Crossref], [PubMed], [CAS], [Google Scholar]

1036. Taylor, G. Analysis of the Swimming of Microscopic Organisms Proc. R. Soc. London, Ser. A 1951, 209,
447– 461 DOI: 10.1098/rspa.1951.0218 [Crossref], [Google Scholar]

1037. Shapere, A.; Wilczek, F. Self-Propulsion at Low Reynolds Number Phys. Rev. Lett. 1987, 58, 2051– 2054
 DOI: 10.1103/PhysRevLett.58.2051 [Crossref], [PubMed], [Google Scholar]
1038. Alfred Shapere, F. W. Geometry of self-propulsion at low Reynolds number J. Fluid Mech. 1989, 198,
557– 585 DOI: 10.1017/S002211208900025X [Crossref], [Google Scholar]   

1039. Stone, H. A.; Samuel, A. D. T. Propulsion of Microorganisms by Surface Distortions Phys. Rev. Lett. 1996,
77, 4102– 4104 DOI: 10.1103/PhysRevLett.77.4102 [Crossref], [PubMed], [CAS], [Google Scholar]

1040. Ajdari, A.; Stone, H. A. A note on swimming using internally generated traveling waves Phys. Fluids 1999,
11, 1275– 1277 DOI: 10.1063/1.869991 [Crossref], [CAS], [Google Scholar]

1041. Dreyfus, R.; Baudry, J.; Stone, H. A. Purcell’s “rotator”: mechanical rotation at low Reynolds number Eur.
Phys. J. B 2005, 47, 161– 164 DOI: 10.1140/epjb/e2005-00302-5 [Crossref], [CAS], [Google Scholar]

1042. Purcell, E. M. The e ciency of propulsion by a rotating agellum Proc. Natl. Acad. Sci. U. S. A. 1997, 94,
11307– 11311 DOI: 10.1073/pnas.94.21.11307 [Crossref], [PubMed], [CAS], [Google Scholar]

1043. Wiggins, C. H.; Goldstein, R. E. Flexive and Propulsive Dynamics of Elastica at Low Reynolds Number
Phys. Rev. Lett. 1998, 80, 3879– 3882 DOI: 10.1103/PhysRevLett.80.3879 [Crossref], [CAS], [Google Scholar]

1044. Camalet, S.; Jülicher, F.; Prost, J. Self-Organized Beating and Swimming of Internally Driven Filaments
Phys. Rev. Lett. 1999, 82, 1590– 1593 DOI: 10.1103/PhysRevLett.82.1590 [Crossref], [CAS], [Google Scholar]

1045. Becker, L. E.; Koehler, S. A.; Stone, H. A. On self-propulsion of micro-machines at low Reynolds number:
Purcell’s three-link swimmer J. Fluid Mech. 2003, 490, 15– 35 DOI: 10.1017/S0022112003005184
[Crossref], [Google Scholar]

1046. Naja , A.; Golestanian, R. Simple swimmer at low Reynolds number: Three linked spheres Phys. Rev. E
2004, 69, 062901 DOI: 10.1103/PhysRevE.69.062901 [Crossref], [CAS], [Google Scholar]

1047. Avron, J. E.; Gat, O.; Kenneth, O. Optimal Swimming at Low Reynolds Numbers Phys. Rev. Lett. 2004, 93,
186001 DOI: 10.1103/PhysRevLett.93.186001 [Crossref], [PubMed], [CAS], [Google Scholar]

1048. Avron, J. E.; Kenneth, O.; Oaknin, D. H. Pushmepullyou: an e cient micro-swimmer New J. Phys. 2005, 7,
234 DOI: 10.1088/1367-2630/7/1/234 [Crossref], [Google Scholar]
1049. Porto, M.; Urbakh, M.; Klafter, J. Atomic Scale Engines: Cars and Wheels Phys. Rev. Lett. 2000, 84, 6058–
6061 DOI: 10.1103/PhysRevLett.84.6058 [Crossref], [PubMed], [CAS], [Google Scholar]   

1050. Wang, Z. Bioinspired laser-operated molecular locomotive Phys. Rev. E 2004, 70, 031903
 DOI: 10.1103/PhysRevE.70.031903 [Crossref], [CAS], [Google Scholar]

1051. Dreyfus, R.; Baudry, J.; Roper, M. L.; Fermigier, M.; Stone, H. A.; Bibette, J. Microscopic arti cial swimmers
Nature 2005, 437, 862– 865 DOI: 10.1038/nature04090 [Crossref], [PubMed], [CAS], [Google Scholar]

1052. Baranova, N. B.; Zel’dovich, B. Y. Separation of mirror isomeric molecules by radio-frequency electric eld
of rotating polarization Chem. Phys. Lett. 1978, 57, 435– 437 DOI: 10.1016/0009-2614(78)85543-2 [Crossref],
[CAS], [Google Scholar]

1053. Space, B.; Rabitz, H.; Lörincz, A.; Moore, P. Feasibility of using photophoresis to create a concentration
gradient of solvated molecules J. Chem. Phys. 1996, 105, 9515– 9524 DOI: 10.1063/1.472785 [Crossref],
[CAS], [Google Scholar]

1054. Arai, Y.; Yasuda, R.; Akashi, K.; Harada, Y.; Miyata, H.; Kinosita, K.; Itoh, H. Tying a molecular knot with
optical tweezers Nature 1999, 399, 446– 448 DOI: 10.1038/20894 [Crossref], [PubMed], [CAS], [Google Scholar]

1055. Neuman, K. C.; Block, S. M. Optical trapping Rev. Sci. Instrum. 2004, 75, 2787– 2809
 DOI: 10.1063/1.1785844 [Crossref], [PubMed], [CAS], [Google Scholar]

1056. Zhuang, X. W. Unraveling DNA condensation with optical tweezers Science 2004, 305, 188– 190
 DOI: 10.1126/science.1100603 [Crossref], [PubMed], [CAS], [Google Scholar]

1057. Allemand, J. F.; Bensimon, D.; Croquette, V. Stretching DNA and RNA to probe their interactions with
proteins Curr. Opin. Struct. Biol. 2003, 13, 266– 274 DOI: 10.1016/S0959-440X(03)00067-8 [Crossref], [PubMed],
[CAS], [Google Scholar]

1058. Bustamante, C.; Macosko, J. C.; Wuite, G. J. Grabbing the cat by the tail: manipulating molecules one by
one Nat. Rev. Mol. Cell Biol. 2000, 1, 130– 136 DOI: 10.1038/35040072 [Crossref], [PubMed], [CAS], [Google
Scholar]

1059. Mehta, A. D.; Rief, M.; Spudich, J. A.; Smith, D. A.; Simmons, R. M. Single-molecule biomechanics with
optical methods Science 1999, 283, 1689– 1695 DOI: 10.1126/science.283.5408.1689 [Crossref], [PubMed],
[CAS], [Google Scholar]

  
1060. Yanagida, T.; Harada, Y.; Ishijima, A. Nano-manipulation of actomyosin molecular motors in vitro: a new
working principle Trends Biochem. Sci. 1993, 18, 319– 324 DOI: 10.1016/0968-0004(93)90064-T [Crossref],
[PubMed], [CAS], [Google Scholar]

1061. Lensen, D.; Elemans, J. A. A. W. Arti cial molecular rotors and motors on surfaces: STM reveals and
triggers Soft Matter 2012, 8, 9053– 9063 DOI: 10.1039/c2sm26235e [Crossref], [CAS], [Google Scholar]

1062. Binning, G.; Rohrer, H.; Gerber, C.; Weibel, E. Surface Studies by Scanning Tunneling Microscopy Phys.
Rev. Lett. 1982, 49, 57– 61 DOI: 10.1103/PhysRevLett.49.57 [Crossref], [CAS], [Google Scholar]

1063. Binnig, G.; Rohrer, H. Scanning Tunneling Microscopy - from Birth to Adolescence Usp. Fiz. Nauk 1988,
154, 261– 278 DOI: 10.3367/UFNr.0154.198802d.0261 [Crossref], [Google Scholar]

1064. Binnig, G.; Rohrer, H. Scanning Tunneling Microscopy - from Birth to Adolescence Angew. Chem., Int. Ed.
Engl. 1987, 26, 606– 614 DOI: 10.1002/anie.198706061 [Crossref], [Google Scholar]

1065. Binnig, G.; Rohrer, H. Scanning Tunneling Microscopy - from Birth to Adolescence Rev. Mod. Phys. 1987,
59, 615– 625 DOI: 10.1103/RevModPhys.59.615 [Crossref], [CAS], [Google Scholar]

1066. Block, S. M. Making light work with optical tweezers Nature 1992, 360, 493– 495 DOI: 10.1038/360493a0
[Crossref], [PubMed], [CAS], [Google Scholar]

1067. Kuo, S. C.; Sheetz, M. P. Optical tweezers in cell biology Trends Cell Biol. 1992, 2, 116– 118
 DOI: 10.1016/0962-8924(92)90016-G [Crossref], [PubMed], [CAS], [Google Scholar]

1068. Iancu, V.; Hla, S.-W. Realization of a four-step molecular switch in scanning tunneling microscope
manipulation of single chlorophyll-a molecules Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 13718– 13721
 DOI: 10.1073/pnas.0603643103 [Crossref], [PubMed], [CAS], [Google Scholar]

1069. Chiang, S. Introduction: Force and Tunneling Microscopy Chem. Rev. 1997, 97, 1015– 1016
 DOI: 10.1021/cr970041m [ACS Full Text ], [Google Scholar]
1070. Gimzewski, J. K.; Joachim, C. Nanoscale science of single molecules using local probes Science 1999,
283, 1683– 1688 DOI: 10.1126/science.283.5408.1683 [Crossref], [PubMed], [CAS], [Google Scholar]
  

1071. Stroscio, J. A.; Eigler, D. M. Atomic and molecular manipulation with the scanning tunneling microscope
Science 1991, 254, 1319– 1326 DOI: 10.1126/science.254.5036.1319 [Crossref], [PubMed], [CAS], [Google
Scholar]

1072. Avouris, P. Manipulation of Matter at the Atomic and Molecular-Levels Acc. Chem. Res. 1995, 28, 95–
102 DOI: 10.1021/ar00051a002 [ACS Full Text ], [CAS], [Google Scholar]

1073. Gimzewski, J. K.; Jung, T. A.; Cuberes, M. T.; Schlittler, R. R. Scanning tunneling microscopy of individual
molecules: beyond imaging Surf. Sci. 1997, 386, 101– 114 DOI: 10.1016/S0039-6028(97)00301-4 [Crossref],
[CAS], [Google Scholar]

1074. Rosei, F.; Schunack, M.; Naitoh, Y.; Jiang, P.; Gourdon, A.; Laegsgaard, E.; Stensgaard, I.; Joachim, C.;
Besenbacher, F. Properties of large organic molecules on metal surfaces Prog. Surf. Sci. 2003, 71, 95– 146
 DOI: 10.1016/S0079-6816(03)00004-2 [Crossref], [CAS], [Google Scholar]

1075. Takano, H.; Kenseth, J. R.; Wong, S. S.; O’Brien, J. C.; Porter, M. D. Chemical and biochemical analysis
using scanning force microscopy Chem. Rev. 1999, 99, 2845– 2890 DOI: 10.1021/cr9801317 [ACS Full Text ],
[CAS], [Google Scholar]

1076. Guthold, M.; Falvo, M.; Matthews, W. G.; Paulson, S.; Mullin, J.; Lord, S.; Erie, D.; Washburn, S.; Super ne,
R.; Brooks, F. P., Jr.; Taylor, R. M., 2nd Investigation and modi cation of molecular structures with the
nanoManipulator J. Mol. Graphics Modell. 1999, 17, 187– 197 DOI: 10.1016/S1093-3263(99)00030-3 [Crossref],
[CAS], [Google Scholar]

1077. Kawakami, M.; Taniguchi, Y. Recent Advances in Single-Molecule Biophysics with the Use of Atomic
Force Microscopy Adv. Chem. Phys. 2012, 146, 89– 132 [CAS], [Google Scholar]

1078. Giessibl, F. J. Advances in atomic force microscopy Rev. Mod. Phys. 2003, 75, 949– 983
 DOI: 10.1103/RevModPhys.75.949 [Crossref], [CAS], [Google Scholar]

1079. Hansma, P. K.; Elings, V. B.; Marti, O.; Bracker, C. E. Scanning tunneling microscopy and atomic force
microscopy: application to biology and technology Science 1988, 242, 209– 216 DOI: 10.1126/science.3051380
[Crossref], [PubMed], [CAS], [Google Scholar]
1080. Engel, A.; Gaub, H. E.; Muller, D. J. Atomic force microscopy: a forceful way with single molecules Curr.
Biol. 1999, 9, R133– R136 DOI: 10.1016/S0960-9822(99)80081-5 [Crossref], [PubMed], [CAS], [Google  
 Scholar]

1081. Clausen-Schaumann, H.; Seitz, M.; Krautbauer, R.; Gaub, H. E. Force spectroscopy with single bio-
molecules Curr. Opin. Chem. Biol. 2000, 4, 524– 530 DOI: 10.1016/S1367-5931(00)00126-5 [Crossref], [PubMed],
[CAS], [Google Scholar]

1082. Samori, B. Stretching single molecules along unbinding and unfolding pathways with the scanning force
microscope Chem. - Eur. J. 2000, 6, 4249– 4255
 DOI: 10.1002/1521-3765(20001201)6:23<4249::AID-CHEM4249>3.3.CO;2-P [Crossref], [PubMed], [CAS], [Google
Scholar]

1083. Janshoff, A.; Neitzert, M.; Oberdorfer, Y.; Fuchs, H. Force Spectroscopy of Molecular Systems-Single
Molecule Spectroscopy of Polymers and Biomolecules Angew. Chem., Int. Ed. 2000, 39, 3212– 3237
 DOI: 10.1002/1521-3773(20000915)39:18<3212::AID-ANIE3212>3.0.CO;2-X [Crossref], [PubMed], [CAS], [Google
Scholar]

1084. Rief, M.; Grubmuller, H. Force spectroscopy of single biomolecules ChemPhysChem 2002, 3, 255– 261
 DOI: 10.1002/1439-7641(20020315)3:3<255::AID-CPHC255>3.0.CO;2-M [Crossref], [PubMed], [CAS], [Google
Scholar]

1085. Best, R. B.; Clarke, J. What can atomic force microscopy tell us about protein folding? Chem. Commun.
2002, 183– 192 DOI: 10.1039/b108159b [Crossref], [PubMed], [CAS], [Google Scholar]

1086. Hla, S. W.; Meyer, G.; Rieder, K. H. Inducing single-molecule chemical reactions with a UHV-STM: a new
dimension for nano-science and technology ChemPhysChem 2001, 2, 361– 366
 DOI: 10.1002/1439-7641(20010618)2:6<361::AID-CPHC361>3.0.CO;2-N [Crossref], [PubMed], [CAS], [Google
Scholar]

1087. Joachim, C.; Gimzewski, J. K. A nanoscale single-molecule ampli er and its consequences Proc. IEEE
1998, 86, 184– 190 DOI: 10.1109/5.658770 [Crossref], [CAS], [Google Scholar]

1088. Joachim, C.; Gimzewski, J. K.; Aviram, A. Electronics using hybrid-molecular and mono-molecular devices
Nature 2000, 408, 541– 548 DOI: 10.1038/35046000 [Crossref], [PubMed], [CAS], [Google Scholar]

1089. Moresco, F. Manipulation of large molecules by low-temperature STM: model systems for molecular
electronics Phys. Rep. 2004, 399, 175– 225 DOI: 10.1016/j.physrep.2004.08.001 [Crossref], [CAS], [Google
Scholar]

  
1090. Bhushan, B.; Israelachvili, J. N.; Landman, U. Nanotribology - Friction, Wear and Lubrication at the Atomic-
Scale Nature 1995, 374, 607– 616 DOI: 10.1038/374607a0 [Crossref], [CAS], [Google Scholar]

1091. Luan, B. Q.; Robbins, M. O. The breakdown of continuum models for mechanical contacts Nature 2005,
435, 929– 932 DOI: 10.1038/nature03700 [Crossref], [PubMed], [CAS], [Google Scholar]

1092. Buldum, A.; Lu, J. P. Atomic scale sliding and rolling of carbon nanotubes Phys. Rev. Lett. 1999, 83,
5050– 5053 DOI: 10.1103/PhysRevLett.83.5050 [Crossref], [CAS], [Google Scholar]

1093. Kang, J. W.; Hwang, H. J. Fullerene nano ball bearings: an atomistic study Nanotechnology 2004, 15,
614– 621 DOI: 10.1088/0957-4484/15/5/036 [Crossref], [CAS], [Google Scholar]

1094. Sasaki, N.; Miura, K. Key issues of nanotribology for successful nanofabrication - From basis to C-60
molecular bearings Jpn. J. Appl. Phys. 1 2004, 43, 4486– 4491 DOI: 10.1143/JJAP.43.4486 [Crossref],
[CAS], [Google Scholar]

1095. Falvo, M. R.; Taylor, R. M., 2nd; Helser, A.; Chi, V.; Brooks, F. P., Jr.; Washburn, S.; Super ne, R. Nanometre-
scale rolling and sliding of carbon nanotubes Nature 1999, 397, 236– 238 DOI: 10.1038/16662 [Crossref],
[PubMed], [CAS], [Google Scholar]

1096. Falvo, M. R.; Steele, J.; Taylor, R. M. Super ne, R. Gearlike rolling motion mediated by commensurate
contact: Carbon nanotubes on HOPG Phys. Rev. B: Condens. Matter Mater. Phys. 2000, 62, 10665– 10667
 DOI: 10.1103/PhysRevB.62.R10665 [Crossref], [CAS], [Google Scholar]

1097. Miura, K.; Takagi, T.; Kamiya, S.; Sahashi, T.; Yamauchi, M. Natural rolling of zigzag multiwalled carbon
nanotubes on graphite Nano Lett. 2001, 1, 161– 163 DOI: 10.1021/nl015513y [ACS Full Text ], [CAS], [Google
Scholar]

1098. Moriarty, P.; Ma, Y. R.; Upward, M. D.; Beton, P. H. Translation, rotation and removal of C-60 on Si(100)-2 ×
1 using anisotropic molecular manipulation Surf. Sci. 1998, 407, 27– 35 DOI: 10.1016/S0039-6028(98)00082-X
[Crossref], [CAS], [Google Scholar]

1099. Miura, K.; Kamiya, S.; Sasaki, N. C-60 molecular bearings. Phys. Rev. Lett. 2003, 90,
 DOI: 10.1103/PhysRevLett.90.055509 . [Crossref], [Google Scholar]
1100. Keeling, D. L.; Humphry, M. J.; Fawcett, R. H. J.; Beton, P. H.; Hobbs, C.; Kantorovich, L. Bond  
breaking
coupled with translation in rolling of covalently bound molecules Phys. Rev. Lett. 2005, 94, 94
 DOI: 10.1103/PhysRevLett.94.146104 [Crossref], [Google Scholar]

1101. Otero, R.; Hummelink, F.; Sato, F.; Legoas, S. B.; Thostrup, P.; Laegsgaard, E.; Stensgaard, I.; Galvao, D. S.;
Besenbacher, F. Lock-and-key effect in the surface diffusion of large organic molecules probed by STM Nat.
Mater. 2004, 3, 779– 782 DOI: 10.1038/nmat1243 [Crossref], [PubMed], [CAS], [Google Scholar]

1102. Eigler, D. M.; Lutz, C. P.; Rudge, W. E. An Atomic Switch Realized with the Scanning Tunneling Microscope
Nature 1991, 352, 600– 603 DOI: 10.1038/352600a0 [Crossref], [CAS], [Google Scholar]

1103. Stipe, B. C.; Rezaei, M. A.; Ho, W. Inducing and viewing the rotational motion of a single molecule Science
1998, 279, 1907– 1909 DOI: 10.1126/science.279.5358.1907 [Crossref], [PubMed], [CAS], [Google Scholar]

1104. Stipe, B. C.; Rezaei, M. A.; Ho, W. Coupling of vibrational excitation to the rotational motion of a single
adsorbed molecule Phys. Rev. Lett. 1998, 81, 1263– 1266 DOI: 10.1103/PhysRevLett.81.1263 [Crossref],
[CAS], [Google Scholar]

1105. Wintjes, N.; Bonifazi, D.; Cheng, F.; Kiebele, A.; Stöhr, M.; Jung, T.; Spillmann, H.; Diederich, F. A
Supramolecular Multiposition Rotary Device Angew. Chem., Int. Ed. 2007, 46, 4089– 4092
 DOI: 10.1002/anie.200700285 [Crossref], [PubMed], [CAS], [Google Scholar]

1106. Jung, T. A.; Schlittler, R. R.; Gimzewski, J. K.; Tang, H.; Joachim, C. Controlled room-temperature
positioning of individual molecules: Molecular exure and motion Science 1996, 271, 181– 184
 DOI: 10.1126/science.271.5246.181 [Crossref], [CAS], [Google Scholar]

1107. Jung, T. A.; Schlittler, R. R.; Gimzewski, J. K. Conformational identi cation of individual adsorbed
molecules with the STM Nature 1997, 386, 696– 698 DOI: 10.1038/386696a0 [Crossref], [CAS], [Google Scholar]

1108. Moresco, F.; Meyer, G.; Rieder, K. H.; Ping, H.; Tang, H.; Joachim, C. TBPP molecules on copper surfaces: a
low temperature scanning tunneling microscope investigation Surf. Sci. 2002, 499, 94– 102
 DOI: 10.1016/S0039-6028(01)01803-9 [Crossref], [CAS], [Google Scholar]

1109. Moresco, F.; Meyer, G.; Rieder, K. H.; Tang, H.; Gourdon, A.; Joachim, C. Low temperature manipulation of
big molecules in constant height mode Appl. Phys. Lett. 2001, 78, 306– 308 DOI: 10.1063/1.1339251 [Crossref],
[CAS], [Google Scholar]
  
1110. Moresco, F.; Meyer, G.; Rieder, K. H.; Tang, H.; Gourdon, A.; Joachim, C. Conformational changes of single
molecules induced by scanning tunneling microscopy manipulation: A route to molecular switching Phys. Rev.
Lett. 2001, 86, 672– 675 DOI: 10.1103/PhysRevLett.86.672 [Crossref], [PubMed], [CAS], [Google Scholar]

1111. Moresco, F.; Meyer, G.; Rieder, K. H.; Tang, H.; Gourdon, A.; Joachim, C. Recording intramolecular
mechanics during the manipulation of a large molecule. Phys. Rev. Lett. 2001, 87,
 DOI: 10.1103/PhysRevLett.87.088302 . [Crossref], [PubMed], [Google Scholar]

1112. Cuberes, M. T.; Schlittler, R. R.; Gimzewski, J. K. Room-temperature repositioning of individual C-60
molecules at Cu steps: Operation of a molecular counting device Appl. Phys. Lett. 1996, 69, 3016– 3018
 DOI: 10.1063/1.116824 [Crossref], [CAS], [Google Scholar]

1113. Gimzewski, J. K.; Joachim, C.; Schlittler, R. R.; Langlais, V.; Tang, H.; Johannsen, I. Rotation of a single
molecule within a supramolecular bearing Science 1998, 281, 531– 533 DOI: 10.1126/science.281.5376.531
[Crossref], [PubMed], [CAS], [Google Scholar]

1114. Griessl, S. J. H.; Lackinger, M.; Jamitzky, F.; Markert, T.; Hietschold, M.; Heckl, W. A. Incorporation and
manipulation of coronene in an organic template structure Langmuir 2004, 20, 9403– 9407
 DOI: 10.1021/la049441c [ACS Full Text ], [CAS], [Google Scholar]

1115. Joachim, C.; Gimzewski, J. K. Single molecular rotor at the nanoscale Struct. Bonding (Berlin) 2001, 99,
1– 18 [Crossref], [CAS], [Google Scholar]

1116. Stranick, S. J.; Kamna, M. M.; Weiss, P. S. Atomic-scale dynamics of a two-dimensional gas-solid
interface Science 1994, 266, 99– 102 DOI: 10.1126/science.266.5182.99 [Crossref], [PubMed], [CAS], [Google
Scholar]

1117. Kwon, K. Y.; Wong, K. L.; Pawin, G.; Bartels, L.; Stolbov, S.; Rahman, T. S. Unidirectional adsorbate motion
on a high-symmetry surface: “Walking” molecules can stay the course. Phys. Rev. Lett. 2005, 95,
 DOI: 10.1103/PhysRevLett.95.166101 . [Crossref], [Google Scholar]

1118. Das, B.; Sebastian, K. L. Adsorbed hypostrophene: can it roll on a surface by rearrangement of bonds?
Chem. Phys. Lett. 2000, 330, 433– 439 DOI: 10.1016/S0009-2614(00)01100-3 [Crossref], [CAS], [Google Scholar]
1119. Das, B.; Sebastian, K. L. Molecular wheels on surfaces Chem. Phys. Lett. 2002, 357, 25– 31
 DOI: 10.1016/S0009-2614(02)00435-9 [Crossref], [CAS], [Google Scholar]   

1120. Fujikawa, Y.; Sadowski, J. T.; Kelly, K. F.; Nakayama, K. S.; Mickelson, E. T.; Hauge, R. H.; Margrave, J. L.;
Sakurai, T. Adsorption of uorinated C-60 on the Si(111)-(7 × 7) surface studied by scanning tunneling
microscopy and high-resolution electron energy loss spectroscopy Jpn. J. Appl. Phys. 1 2002, 41, 245– 249
 DOI: 10.1143/JJAP.41.245 [Crossref], [CAS], [Google Scholar]

1121. Joachim, C.; Tang, H.; Moresco, F.; Rapenne, G.; Meyer, G. The design of a nanoscale molecular barrow
Nanotechnology 2002, 13, 330– 335 DOI: 10.1088/0957-4484/13/3/318 [Crossref], [CAS], [Google Scholar]

1122. Jimenez-Bueno, G.; Rapenne, G. Technomimetic molecules: synthesis of a molecular wheelbarrow


Tetrahedron Lett. 2003, 44, 6261– 6263 DOI: 10.1016/S0040-4039(03)01519-3 [Crossref], [CAS], [Google
Scholar]

1123. Rapenne, G. Synthesis of technomimetic molecules: towards rotation control in single-molecular


machines and motors Org. Biomol. Chem. 2005, 3, 1165– 1169 DOI: 10.1039/b419282f [Crossref], [PubMed],
[CAS], [Google Scholar]

1124. Grill, L.; Rieder, K. H.; Moresco, F.; Jimenez-Bueno, G.; Wang, C.; Rapenne, G.; Joachim, C. Imaging of a
molecular wheelbarrow by scanning tunneling microscopy Surf. Sci. 2005, 584, L153– L158
 DOI: 10.1016/j.susc.2005.03.062 [Crossref], [CAS], [Google Scholar]

1125. Grill, L.; Rieder, K. H.; Moresco, F.; Rapenne, G.; Stojkovic, S.; Bouju, X.; Joachim, C. Rolling a single
molecular wheel at the atomic scale Nat. Nanotechnol. 2007, 2, 95– 98 DOI: 10.1038/nnano.2006.210
[Crossref], [PubMed], [CAS], [Google Scholar]

1126. Vaughan, O. P.; Williams, F. J.; Bampos, N.; Lambert, R. M. A chemically switchable molecular pinwheel
Angew. Chem., Int. Ed. 2006, 45, 3779– 3781 DOI: 10.1002/anie.200600683 [Crossref], [PubMed], [CAS], [Google
Scholar]

1127. Wong, K. L.; Pawin, G.; Kwon, K. Y.; Lin, X.; Jiao, T.; Solanki, U.; Fawcett, R. H. J.; Bartels, L.; Stolbov, S.;
Rahman, T. S. A molecule carrier Science 2007, 315, 1391– 1393 DOI: 10.1126/science.1135302 [Crossref],
[PubMed], [CAS], [Google Scholar]

1128. Chiaravalloti, F.; Gross, L.; Rieder, K. H.; Stojkovic, S. M.; Gourdon, A.; Joachim, C.; Moresco, F. A rack-and-
pinion device at the molecular scale Nat. Mater. 2007, 6, 30– 33 DOI: 10.1038/nmat1802 [Crossref], [PubMed],
[CAS], [Google Scholar]

  
1129. Shirai, Y.; Osgood, A. J.; Zhao, Y.; Kelly, K. F.; Tour, J. M. Directional control in Thermally Driven Single-
molecule Nanocars Nano Lett. 2005, 5, 2330– 2334 DOI: 10.1021/nl051915k [ACS Full Text ], [CAS], [Google
Scholar]

1130. Chu, P. L. E.; Wang, L. Y.; Khatua, S.; Kolomeisky, A. B.; Link, S.; Tour, J. M. Synthesis and Single-Molecule
Imaging of Highly Mobile Adamantane-Wheeled Nanocars ACS Nano 2013, 7, 35– 41 DOI: 10.1021/nn304584a
[ACS Full Text ], [CAS], [Google Scholar]

1131. Khatua, S.; Guerrero, J. M.; Claytor, K.; Vives, G.; Kolomeisky, A. B.; Tour, J. M.; Link, S. Micrometer-Scale
Translation and Monitoring of Individual Nanocars on Glass ACS Nano 2009, 3, 351– 356
 DOI: 10.1021/nn800798a [ACS Full Text ], [CAS], [Google Scholar]

1132. Morin, J. F.; Shirai, Y.; Tour, J. M. En route to a motorized nanocar Org. Lett. 2006, 8, 1713– 1716
 DOI: 10.1021/ol060445d [ACS Full Text ], [CAS], [Google Scholar]

1133. Tierney, H. L.; Murphy, C. J.; Jewell, A. D.; Baber, A. E.; Iski, E. V.; Khodaverdian, H. Y.; McGuire, A. F.;
Klebanov, N.; Sykes, E. C. H. Experimental demonstration of a single-molecule electric motor Nat. Nanotechnol.
2011, 6, 625– 629 DOI: 10.1038/nnano.2011.142 [Crossref], [PubMed], [CAS], [Google Scholar]

1134. Tanaka, H.; Ikeda, T.; Takeuchi, M.; Sada, K.; Shinkai, S.; Kawai, T. Molecular Rotation in Self-Assembled
Multidecker Porphyrin Complexes ACS Nano 2011, 5, 9575– 9582 DOI: 10.1021/nn203773p [ACS Full Text ],
[CAS], [Google Scholar]

1135. Kudernac, T.; Ruangsupapichat, N.; Parschau, M.; Macia, B.; Katsonis, N.; Harutyunyan, S. R.; Ernst, K. H.;
Feringa, B. L. Electrically driven directional motion of a four-wheeled molecule on a metal surface Nature 2011,
479, 208– 211 DOI: 10.1038/nature10587 [Crossref], [PubMed], [CAS], [Google Scholar]

1136. Hao, X.; Zhu, N.; Gschneidtner, T.; Jonsson, E. O.; Zhang, J. D.; Moth-Poulsen, K.; Wang, H. D.; Thygesen, K.
S.; Jacobsen, K. W.; Ulstrup, J., et al. Direct measurement and modulation of single-molecule coordinative
bonding forces in a transition metal complex Nat. Commun. 2013, 4, 2121 DOI: 10.1038/ncomms3121
[Crossref], [PubMed], [CAS], [Google Scholar]

1137. Biscarini, F.; Gebauer, W.; Di Domenico, D.; Zamboni, R.; Pascual, J. I.; Leigh, D. A.; Murphy, A.; Tetard, D.
STM investigation of exible supramolecules: Benzylic amide [2] catenanes Synth. Met. 1999, 102, 1466– 1467
 DOI: 10.1016/S0379-6779(98)00539-6 [Crossref], [CAS], [Google Scholar]
1138. Samori, P.; Jackel, F.; Unsal, O.; Godt, A.; Rabe, J. P. Ordered nanostructures of a [2]catenane  
through self-
assembly at surfaces - An STM study with sub-molecular resolution ChemPhysChem 2001, 2, 461– 464
 DOI: 10.1002/1439-7641(20010716)2:7<461::AID-CPHC461>3.3.CO;2-A [Crossref], [PubMed], [CAS], [Google
Scholar]

1139. Shigekawa, H.; Miyake, K.; Sumaoka, J.; Harada, A.; Komiyama, M. The molecular abacus: STM
manipulation of cyclodextrin necklace J. Am. Chem. Soc. 2000, 122, 5411– 5412 DOI: 10.1021/ja000037j [ACS
Full Text ], [CAS], [Google Scholar]

1140. Feng, M.; Guo, X. F.; Lin, X.; He, X. B.; Ji, W.; Du, S. X.; Zhang, D. Q.; Zhu, D. B.; Gao, H. J. Stable, reproducible
nanorecording on rotaxane thin lms J. Am. Chem. Soc. 2005, 127, 15338– 15339 DOI: 10.1021/ja054836j [ACS
Full Text ], [CAS], [Google Scholar]

1141. Cavallini, M.; Biscarini, F.; Leon, S.; Zerbetto, F.; Bottari, G.; Leigh, D. A. Information storage using
supramolecular surface patterns Science 2003, 299, 531– 531 DOI: 10.1126/science.1078012 [Crossref],
[PubMed], [CAS], [Google Scholar]

1142. Brough, B.; Northrop, B. H.; Schmidt, J. J.; Tseng, H. R.; Houk, K. N.; Stoddart, J. F.; Ho, C. M. Evaluation of
synthetic linear motor-molecule actuation energetics Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 8583– 8588
 DOI: 10.1073/pnas.0509645103 [Crossref], [PubMed], [CAS], [Google Scholar]

1143. Lussis, P.; Svaldo-Lanero, T.; Bertocco, A.; Fustin, C. A.; Leigh, D. A.; Duwez, A. S. A single synthetic small
molecule that generates force against a load Nat. Nanotechnol. 2011, 6, 553– 557 DOI: 10.1038/nnano.2011.132
[Crossref], [PubMed], [CAS], [Google Scholar]

1144. Luo, Y.; Collier, C. P.; Jeppesen, J. O.; Nielsen, K. A.; DeIonno, E.; Ho, G.; Perkins, J.; Tseng, H. R.;
Yamamoto, T.; Stoddart, J. F.; Heath, J. R. Two-dimensional molecular electronics circuits ChemPhysChem 2002,
3, 519– 525 DOI: 10.1002/1439-7641(20020617)3:6<519::AID-CPHC519>3.0.CO;2-2 [Crossref], [PubMed],
[CAS], [Google Scholar]

1145. Huang, T. J.; Flood, A.; Chu, C. W.; Kang, S.; Guo, T. F.; Yamamoto, T.; Tseng, H. R.; Yu, B. D.; Yang, Y.;
Stoddart, J. F.; Ho, C. M. In situ infrared spectroscopic studies of molecular behavior in nanoelectronic devices
IEEE-NANO 2003, 2, 698– 701 DOI: 10.1109/NANO.2003.1231008 [Crossref], [Google Scholar]

1146. Collier, C. P.; Wong, E. W.; Belohradsky, M.; Raymo, F. M.; Stoddart, J. F.; Kuekes, P. J.; Williams, R. S.;
Heath, J. R. Electronically con gurable molecular-based logic gates Science 1999, 285, 391– 394
 DOI: 10.1126/science.285.5426.391 [Crossref], [PubMed], [CAS], [Google Scholar]

  
1147. Kitagawa, K.; Morita, T.; Kimura, S. A helical molecule that exhibits two lengths in response to an applied
potential Angew. Chem., Int. Ed. 2005, 44, 6330– 6333 DOI: 10.1002/anie.200502240 [Crossref], [PubMed],
[CAS], [Google Scholar]

1148. Carella, A.; Rapenne, G.; Launay, J. P. Design and synthesis of the active part of a potential molecular
motor New J. Chem. 2005, 29, 288– 290 DOI: 10.1039/b415214j [Crossref], [CAS], [Google Scholar]

1149. Carella, A.; Jaud, J.; Rapenne, G.; Launay, J. P. Technomimetic molecules: synthesis of ruthenium(II)
1,2,3,4,5-penta(p-bromophenyl)cyclopentadienyl hydrotris(indazolyl)borate, an organometallic molecular turnstile
Chem. Commun. 2003, 2434– 2435 DOI: 10.1039/b307577j [Crossref], [PubMed], [CAS], [Google Scholar]

1150. DeIonno, E.; Tseng, H. R.; Harvey, D. D.; Stoddart, J. F.; Heath, J. R. Infrared spectroscopic characterization
of [2]rotaxane molecular switch tunnel junction devices J. Phys. Chem. B 2006, 110, 7609– 7612
 DOI: 10.1021/jp0607723 [ACS Full Text ], [CAS], [Google Scholar]

1151. Flood, A. H.; Wong, E. W.; Stoddart, J. F. Models of charge transport and transfer in molecular switch
tunnel junctions of bistable catenanes and rotaxanes Chem. Phys. 2006, 324, 280– 290
 DOI: 10.1016/j.chemphys.2005.12.031 [Crossref], [CAS], [Google Scholar]

1152. Flood, A. H.; Nygaard, S.; Laursen, B. W.; Jeppesen, J. O.; Stoddart, J. F. Locking down the electronic
structure of (monopyrrolo)tetrathiafulvalene in [2]rotaxanes Org. Lett. 2006, 8, 2205– 2208
 DOI: 10.1021/ol060319+ [ACS Full Text ], [CAS], [Google Scholar]

1153. Loppacher, C.; Guggisberg, M.; Pfeiffer, O.; Meyer, E.; Bammerlin, M.; Luthi, R.; Schlittler, R.; Gimzewski, J.
K.; Tang, H.; Joachim, C. Direct determination of the energy required to operate a single molecule switch. Phys.
Rev. Lett. 2003, 90,  DOI: 10.1103/PhysRevLett.90.066107 . [Crossref], [PubMed], [Google Scholar]

1154. Viala, C.; Secchi, A.; Gourdon, A. Synthesis of polyaromatic hydrocarbons with a central rotor Eur. J. Org.
Chem. 2002, 2002, 4185– 4189 DOI: 10.1002/1099-0690(200212)2002:24<4185::AID-EJOC4185>3.0.CO;2-Z
[Crossref], [Google Scholar]

1155. Gourdon, A. Synthesis of ″molecular landers″ Eur. J. Org. Chem. 1998, 1998, 2797– 2801
 DOI: 10.1002/(SICI)1099-0690(199812)1998:12<2797::AID-EJOC2797>3.0.CO;2-D [Crossref], [Google Scholar]
1156. Langlais, V. J.; Schlittler, R. R.; Tang, H.; Gourdon, A.; Joachim, C.; Gimzewski, J. K. Spatially resolved
 
tunneling along a molecular wire Phys. Rev. Lett. 1999, 83, 2809– 2812 DOI: 10.1103/PhysRevLett.83.2809 
[Crossref], [CAS], [Google Scholar]

1157. Moresco, F.; Gross, L.; Grill, L.; Alemani, M.; Gourdon, A.; Joachim, C.; Rieder, K. H. Contacting a single
molecular wire by STM manipulation Appl. Phys. A: Mater. Sci. Process. 2005, 80, 913– 920
 DOI: 10.1007/s00339-004-3116-x [Crossref], [CAS], [Google Scholar]

1158. Moresco, F.; Gourdon, A. Scanning tunneling microscopy experiments on single molecular landers Proc.
Natl. Acad. Sci. U. S. A. 2005, 102, 8809– 8814 DOI: 10.1073/pnas.0500915102 [Crossref], [PubMed],
[CAS], [Google Scholar]

1159. Soukiassian, L.; Mayne, A. J.; Comtet, G.; Hellner, L.; Dujardin, G.; Gourdon, A. Selective internal
manipulation of a single molecule by scanning tunneling microscopy J. Chem. Phys. 2005, 122, 134704
 DOI: 10.1063/1.1874972 [Crossref], [PubMed], [CAS], [Google Scholar]

1160. Lastapis, M.; Martin, M.; Riedel, D.; Hellner, L.; Comtet, G.; Dujardin, G. Picometer-scale electronic control
of molecular dynamics inside a single molecule Science 2005, 308, 1000– 1003 DOI: 10.1126/science.1108048
[Crossref], [PubMed], [CAS], [Google Scholar]

1161. Metiu, H. Special topic: Single-molecule physics and chemistry - Preface J. Chem. Phys. 2002, 117,
10923– 10923 [CAS], [Google Scholar]

1162. Barbara, P. F. Single-molecule spectroscopy Acc. Chem. Res. 2005, 38, 503– 503
 DOI: 10.1021/ar050120h [ACS Full Text ], [CAS], [Google Scholar]

1163. Chiou, P. Y.; Ohta, A. T.; Wu, M. C. Massively parallel manipulation of single cells and microparticles using
optical images Nature 2005, 436, 370– 372 DOI: 10.1038/nature03831 [Crossref], [PubMed], [CAS], [Google
Scholar]

1164. Dholakia, K. Micromanipulation - Optoelectronic tweezers Nat. Mater. 2005, 4, 579– 580
 DOI: 10.1038/nmat1436 [Crossref], [PubMed], [CAS], [Google Scholar]

1165. Moerner, W. E.; Basche, T. Optical Spectroscopy of Single Impurity Molecules in Solids Angew. Chem., Int.
Ed. Engl. 1993, 32, 457– 476 DOI: 10.1002/anie.199304573 [Crossref], [Google Scholar]
1166. Moerner, W. E. Examining Nanoenvironments in Solids on the Scale of a Single, Isolated Impurity
 
Molecule Science 1994, 265, 46– 53 DOI: 10.1126/science.265.5168.46 [Crossref], [PubMed], [CAS], [Google 
Scholar]

1167. Moerner, W. E. High-resolution optical spectroscopy of single molecules in solids Acc. Chem. Res. 1996,
29, 563– 571 DOI: 10.1021/ar950245u [ACS Full Text ], [CAS], [Google Scholar]

1168. Xie, X. S. Single-molecule spectroscopy and dynamics at room temperature Acc. Chem. Res. 1996, 29,
598– 606 DOI: 10.1021/ar950246m [ACS Full Text ], [CAS], [Google Scholar]

1169. Goodwin, P. M.; Ambrose, W. P.; Keller, R. A. Single-molecule detection in liquids by laser-induced
uorescence Acc. Chem. Res. 1996, 29, 607– 613 DOI: 10.1021/ar950250y [ACS Full Text ], [CAS], [Google
Scholar]

1170. Nie, S. M.; Zare, R. N. Optical detection of single molecules Annu. Rev. Biophys. Biomol. Struct. 1997, 26,
567– 596 DOI: 10.1146/annurev.biophys.26.1.567 [Crossref], [PubMed], [CAS], [Google Scholar]

1171. Plakhotnik, T.; Donley, E. A.; Wild, U. P. Single-molecule spectroscopy Annu. Rev. Phys. Chem. 1997, 48,
181– 212 DOI: 10.1146/annurev.physchem.48.1.181 [Crossref], [PubMed], [CAS], [Google Scholar]

1172. Xie, X. S.; Trautman, J. K. Optical studies of single molecules at room temperature Annu. Rev. Phys.
Chem. 1998, 49, 441– 480 DOI: 10.1146/annurev.physchem.49.1.441 [Crossref], [PubMed], [CAS], [Google
Scholar]

1173. Moerner, W. E.; Orrit, M. Illuminating single molecules in condensed matter Science 1999, 283, 1670–
1676 DOI: 10.1126/science.283.5408.1670 [Crossref], [PubMed], [CAS], [Google Scholar]

1174. Weiss, S. Fluorescence spectroscopy of single biomolecules Science 1999, 283, 1676– 1683
 DOI: 10.1126/science.283.5408.1676 [Crossref], [PubMed], [CAS], [Google Scholar]

1175. Ambrose, W. P.; Goodwin, P. M.; Jett, J. H.; Van Orden, A.; Werner, J. H.; Keller, R. A. Single molecule
uorescence spectroscopy at ambient temperature Chem. Rev. 1999, 99, 2929– 2956 DOI: 10.1021/cr980132z
[ACS Full Text ], [CAS], [Google Scholar]

1176. Deniz, A. A.; Laurence, T. A.; Dahan, M.; Chemla, D. S.; Schultz, P. G.; Weiss, S. Ratiometric single-molecule
studies of freely diffusing biomolecules Annu. Rev. Phys. Chem. 2001, 52, 233– 253
 DOI: 10.1146/annurev.physchem.52.1.233 [Crossref], [PubMed], [CAS], [Google Scholar]

  
1177. Moerner, W. E. A dozen years of single-molecule spectroscopy in physics, chemistry, and biophysics J.
Phys. Chem. B 2002, 106, 910– 927 DOI: 10.1021/jp012992g [ACS Full Text ], [CAS], [Google Scholar]

1178. Keller, R. A.; Ambrose, W. P.; Arias, A. A.; Gai, H.; Emory, S. R.; Goodwin, P. M.; Jett, J. H. Analytical
applications of single-molecule detection Anal. Chem. 2002, 74, 316a– 324a DOI: 10.1021/ac022035i [ACS Full
Text ], [CAS], [Google Scholar]

1179. Bohmer, M.; Enderlein, J. Fluorescence spectroscopy of single molecules under ambient conditions:
Methodology and technology ChemPhysChem 2003, 4, 793– 808 DOI: 10.1002/cphc.200200565 [Crossref],
[PubMed], [CAS], [Google Scholar]

1180. Moerner, W. E.; Fromm, D. P. Methods of single-molecule uorescence spectroscopy and microscopy
Rev. Sci. Instrum. 2003, 74, 3597– 3619 DOI: 10.1063/1.1589587 [Crossref], [CAS], [Google Scholar]

1181. Tinnefeld, P.; Sauer, M. Branching out of single-molecule uorescence spectroscopy: Challenges for
chemistry and in uence on biology Angew. Chem., Int. Ed. 2005, 44, 2642– 2671 DOI: 10.1002/anie.200300647
[Crossref], [PubMed], [CAS], [Google Scholar]

1182. Fang, N.; Lee, H.; Sun, C.; Zhang, X. Sub-diffraction-limited optical imaging with a silver superlens Science
2005, 308, 534– 537 DOI: 10.1126/science.1108759 [Crossref], [PubMed], [CAS], [Google Scholar]

1183. Nishimura, D.; Takashima, Y.; Aoki, H.; Takahashi, T.; Yamaguchi, H.; Ito, S.; Harada, A. Single-Molecule
Imaging of Rotaxanes Immobilized on Glass Substrates: Observation of Rotary Movement Angew. Chem., Int. Ed.
2008, 47, 6077– 6079 DOI: 10.1002/anie.200801431 [Crossref], [PubMed], [CAS], [Google Scholar]

1184. Crooks, R. M.; Ricco, A. J. New organic materials suitable for use in chemical sensor arrays Acc. Chem.
Res. 1998, 31, 219– 227 DOI: 10.1021/ar970246h [ACS Full Text ], [CAS], [Google Scholar]

1185. Goldenberg, L. M.; Bryce, M. R.; Petty, M. C. Chemosensor devices: voltammetric molecular recognition at
solid interfaces J. Mater. Chem. 1999, 9, 1957– 1974 DOI: 10.1039/a901825e [Crossref], [CAS], [Google Scholar]

1186. Flink, S.; van Veggel, F. C. J. M.; Reinhoudt, D. N. Sensor functionalities in self-assembled monolayers
Adv. Mater. 2000, 12, 1315– 1328 DOI: 10.1002/1521-4095(200009)12:18<1315::AID-ADMA1315>3.3.CO;2-B
[Crossref], [CAS], [Google Scholar]
 
sensor
1187. Shipway, A. N.; Katz, E.; Willner, I. Nanoparticle arrays on surfaces for electronic, optical, and
applications ChemPhysChem 2000, 1, 18– 52
 DOI: 10.1002/1439-7641(20000804)1:1<18::AID-CPHC18>3.0.CO;2-L [Crossref], [PubMed], [CAS], [Google
Scholar]

1188. Katz, E.; Willner, I. Probing biomolecular interactions at conductive and semiconductive surfaces by
impedance spectroscopy: Routes to impedimetric immunosensors, DNA-Sensors, and enzyme biosensors
Electroanalysis 2003, 15, 913– 947 DOI: 10.1002/elan.200390114 [Crossref], [CAS], [Google Scholar]

1189. Cooke, G. Electrochemical and photochemical control of host-guest complexation at surfaces Angew.
Chem., Int. Ed. 2003, 42, 4860– 4870 DOI: 10.1002/anie.200301636 [Crossref], [PubMed], [CAS], [Google Scholar]

1190. Astruc, D.; Daniel, M. C.; Ruiz, J. Dendrimers and gold nanoparticles as exo-receptors sensing biologically
important anions Chem. Commun. 2004, 2637– 2649 DOI: 10.1039/b410399h [Crossref], [PubMed],
[CAS], [Google Scholar]

1191. Drechsler, U.; Erdogan, B.; Rotello, V. M. Nanoparticles: Scaffolds for molecular recognition Chem. - Eur. J.
2004, 10, 5570– 5579 DOI: 10.1002/chem.200306076 [Crossref], [PubMed], [CAS], [Google Scholar]

1192. Mallouk, T. E.; Gavin, J. A. Molecular recognition in lamellar solids and thin lms Acc. Chem. Res. 1998,
31, 209– 217 DOI: 10.1021/ar970038p [ACS Full Text ], [CAS], [Google Scholar]

1193. Chia, S. Y.; Cao, J. G.; Stoddart, J. F.; Zink, J. I. Working supramolecular machines trapped in glass and
mounted on a lm surface Angew. Chem., Int. Ed. 2001, 40, 2447– 2451
 DOI: 10.1002/1521-3773(20010702)40:13<2447::AID-ANIE2447>3.0.CO;2-P [Crossref], [PubMed], [CAS], [Google
Scholar]

1194. Gase, T.; Grando, D.; Chollet, P. A.; Kajzar, F.; Murphy, A.; Leigh, D. A. Linear and unanticipated second-order
nonlinear optical properties of benzylic amide [2]Catenane thin lms: Evidence of partial rotation of the
interlocked molecular rings in the solid state Adv. Mater. 1999, 11, 1303– 1306
 DOI: 10.1002/(SICI)1521-4095(199910)11:15<1303::AID-ADMA1303>3.0.CO;2-4 [Crossref], [CAS], [Google
Scholar]

1195. Deleuze, M. S. Can benzylic amide [2]catenane rings rotate on graphite? J. Am. Chem. Soc. 2000, 122,
1130– 1143 DOI: 10.1021/ja992458a [ACS Full Text ], [CAS], [Google Scholar]
1196. Bidan, G.; Billon, M.; Divisia-Blohorn, B.; Kern, J. M.; Raehm, L.; Sauvage, J.-P. Electrode-deposited lms of
polyrotaxanes: electrochemically induced gliding motion New J. Chem. 1998, 22, 1139– 1141   
 DOI: 10.1039/a804969f [Crossref], [CAS], [Google Scholar]

1197. Raehm, L.; Kern, J. M.; Sauvage, J.-P.; Hamann, C.; Palacin, S.; Bourgoin, J. P. Disul de- and thiol-
incorporating copper catenanes: Synthesis deposition onto gold, and surface studies Chem. - Eur. J. 2002, 8,
2153– 2162 DOI: 10.1002/1521-3765(20020503)8:9<2153::AID-CHEM2153>3.0.CO;2-E [Crossref], [PubMed],
[CAS], [Google Scholar]

1198. Tseng, H. R.; Wu, D. M.; Fang, N. X. L.; Zhang, X.; Stoddart, J. F. The metastability of an electrochemically
controlled nanoscale machine on gold surfaces ChemPhysChem 2004, 5, 111– 116
 DOI: 10.1002/cphc.200300992 [Crossref], [PubMed], [CAS], [Google Scholar]

1199. Huang, T. J.; Tseng, H. R.; Sha, L.; Lu, W. X.; Brough, B.; Flood, A. H.; Yu, B. D.; Celestre, P. C.; Chang, J. P.;
Stoddart, J. F.; Ho, C. M. Mechanical shuttling of linear motor-molecules in condensed phases on solid
substrates Nano Lett. 2004, 4, 2065– 2071 DOI: 10.1021/nl035099x [ACS Full Text ], [CAS], [Google Scholar]

1200. Long, B.; Nikitin, K.; Fitzmaurice, D. Assembly of an electronically switchable rotaxane on the surface of a
titanium dioxide nanoparticle J. Am. Chem. Soc. 2003, 125, 15490– 15498 DOI: 10.1021/ja037592g [ACS Full
Text ], [CAS], [Google Scholar]

1201. Nikitin, K.; Fitzmaurice, D. The oxidation-state dependent structural conformation and supramolecular
function of a redox-active [2]rotaxane in solution J. Am. Chem. Soc. 2005, 127, 8067– 8076
 DOI: 10.1021/ja050694h [ACS Full Text ], [CAS], [Google Scholar]

1202. van Delden, R. A.; ter Wiel, M. K. J.; Pollard, M. M.; Vicario, J.; Koumura, N.; Feringa, B. L. Unidirectional
molecular motor on a gold surface Nature 2005, 437, 1337– 1340 DOI: 10.1038/nature04127 [Crossref],
[PubMed], [CAS], [Google Scholar]

1203. Hou, S. M.; Sagara, T.; Xu, D. C.; Kelly, T. R.; Ganz, E. Investigation of triptycene-based surface-mounted
rotors Nanotechnology 2003, 14, 566– 570 DOI: 10.1088/0957-4484/14/5/316 [Crossref], [CAS], [Google
Scholar]

1204. Jensen, K.; Girit, Ç.; Mickelson, W.; Zettl, A. Tunable Nanoresonators Constructed from Telescoping
Nanotubes Phys. Rev. Lett. 2006, 96, 215503 DOI: 10.1103/PhysRevLett.96.215503 [Crossref], [PubMed],
[CAS], [Google Scholar]
1205. Deshpande, V. V.; Chiu, H. Y.; Postma, H. W. C.; Mikó, C.; Forró, L.; Bockrath, M. Carbon Nanotube Linear
Bearing Nanoswitches Nano Lett. 2006, 6, 1092– 1095 DOI: 10.1021/nl052513f [ACS Full Text ],  
[CAS], [Google
Scholar]

1206. Ebron, V. H.; Yang, Z.; Seyer, D. J.; Kozlov, M. E.; Oh, J.; Xie, H.; Razal, J.; Hall, L. J.; Ferraris, J. P.;
MacDiarmid, A. G.; Baughman, R. H. Fuel-Powered Arti cial Muscles Science 2006, 311, 1580– 1583
 DOI: 10.1126/science.1120182 [Crossref], [PubMed], [CAS], [Google Scholar]

1207. Huang, T. J.; Flood, A. H.; Brough, B.; Liu, Y.; Bonvallet, P. A.; Kang, S. S.; Chu, C. W.; Guo, T. F.; Lu, W. X.;
Yang, Y., et al. Understanding and harnessing biomimetic molecular machines for NEMS actuation materials
IEEE Trans. Autom. Sci. Eng. 2006, 3, 254– 259 DOI: 10.1109/TASE.2006.875543 [Crossref], [Google Scholar]

1208. Yamaguchi, H.; Kobayashi, Y.; Kobayashi, R.; Takashima, Y.; Hashidzume, A.; Harada, A. Photoswitchable
gel assembly based on molecular recognition Nat. Commun. 2012, 3, 603 DOI: 10.1038/ncomms1617
[Crossref], [PubMed], [CAS], [Google Scholar]

1209. Zheng, Y.; Hashidzume, A.; Takashima, Y.; Yamaguchi, H.; Harada, A. Switching of macroscopic molecular
recognition selectivity using a mixed solvent system Nat. Commun. 2012, 3, 831 DOI: 10.1038/ncomms1841
[Crossref], [PubMed], [CAS], [Google Scholar]

1210. Abelow, A. E.; Zharov, I. Reversible nanovalves in inorganic materials J. Mater. Chem. 2012, 22, 21810–
21818 DOI: 10.1039/c2jm33437b [Crossref], [CAS], [Google Scholar]

1211. Adiga, S. P.; Brenner, D. W. Toward designing smart nanovalves: Modeling of ow control through
nanopores via the helix-coil transition of grafted polypeptide chains Macromolecules 2007, 40, 1342– 1348
 DOI: 10.1021/ma0617522 [ACS Full Text ], [CAS], [Google Scholar]

1212. Angelos, S.; Yang, Y. W.; Patel, K.; Stoddart, J. F.; Zink, J. I. pH-responsive supramolecular nanovalves
based on cucurbit[6]uril pseudorotaxanes Angew. Chem., Int. Ed. 2008, 47, 2222– 2226
 DOI: 10.1002/anie.200705211 [Crossref], [PubMed], [CAS], [Google Scholar]

1213. Chen, T.; Fu, J. J. pH-responsive nanovalves based on hollow mesoporous silica spheres for controlled
release of corrosion inhibitor Nanotechnology 2012, 23, 235605 DOI: 10.1088/0957-4484/23/23/235605
[Crossref], [PubMed], [CAS], [Google Scholar]

1214. Croissant, J.; Chaix, A.; Mongin, O.; Wang, M.; Clement, S.; Raehm, L.; Durand, J. O.; Hugues, V.; Blanchard-
Desce, M.; Maynadier, M., et al. Two-Photon-Triggered Drug Delivery via Fluorescent Nanovalves Small 2014, 10,
1752– 1755 DOI: 10.1002/smll.201400042 [Crossref], [PubMed], [CAS], [Google Scholar]

  
1215. Hwang, A. A.; Lu, J.; Tamanoi, F.; Zink, J. I. Functional Nanovalves on Protein-Coated Nanoparticles for In
vitro and In vivo Controlled Drug Delivery Small 2015, 11, 319– 328 DOI: 10.1002/smll.201400765 [Crossref],
[PubMed], [CAS], [Google Scholar]

1216. Dong, J. Y.; Xue, M.; Zink, J. I. Functioning of nanovalves on polymer coated mesoporous silica
Nanoparticles Nanoscale 2013, 5, 10300– 10306 DOI: 10.1039/c3nr03442a [Crossref], [PubMed], [CAS], [Google
Scholar]

1217. Lau, Y. A.; Ferris, D. P.; Zink, J. I. Photo-driven nano-impellers and nanovalves for on-command drug
release. Proc. SPIE 2010, 7574, 75740P,  DOI: 10.1117/12.841175 . [Crossref], [CAS], [Google Scholar]

1218. Li, Q. L.; Wang, L.; Qiu, X. L.; Sun, Y. L.; Wang, P. X.; Liu, Y.; Li, F.; Qi, A. D.; Gao, H.; Yang, Y. W. Stimuli-
responsive biocompatible nanovalves based on beta-cyclodextrin modi ed poly(glycidyl methacrylate) Polym.
Chem. 2014, 5, 3389– 3395 DOI: 10.1039/c4py00041b [Crossref], [CAS], [Google Scholar]

1219. Liu, J. S.; Du, X. Z. pH- and competitor-driven nanovalves of cucurbit[7]uril pseudorotaxanes based on
mesoporous silica supports for controlled release J. Mater. Chem. 2010, 20, 3642– 3649
 DOI: 10.1039/b915510d [Crossref], [CAS], [Google Scholar]

1220. Liu, J. S.; Du, X. Z.; Zhang, X. F. Enzyme-Inspired Controlled Release of Cucurbit[7]uril Nanovalves by
Using Magnetic Mesoporous Silica Chem. - Eur. J. 2011, 17, 810– 815 DOI: 10.1002/chem.201002899
[Crossref], [PubMed], [CAS], [Google Scholar]

1221. Luo, G. F.; Chen, W. H.; Liu, Y.; Zhang, J.; Cheng, S. X.; Zhuo, R. X.; Zhang, X. Z. Charge-reversal plug gate
nanovalves on peptide-functionalized mesoporous silica nanoparticles for targeted drug delivery J. Mater. Chem.
B 2013, 1, 5723– 5732 DOI: 10.1039/c3tb20792g [Crossref], [CAS], [Google Scholar]

1222. Meng, H. A.; Xue, M.; Xia, T. A.; Zhao, Y. L.; Tamanoi, F.; Stoddart, J. F.; Zink, J. I.; Nel, A. E. Autonomous in
Vitro Anticancer Drug Release from Mesoporous Silica Nanoparticles by pH-Sensitive Nanovalves J. Am. Chem.
Soc. 2010, 132, 12690– 12697 DOI: 10.1021/ja104501a [ACS Full Text ], [CAS], [Google Scholar]

1223. Saha, S.; Leung, K. C. F.; Nguyen, T. D.; Stoddart, J. F.; Zink, J. I. Nanovalves Adv. Funct. Mater. 2007, 17,
685– 693 DOI: 10.1002/adfm.200600989 [Crossref], [CAS], [Google Scholar]
1224. Yang, Y. W. Towards biocompatible nanovalves based on mesoporous silica nanoparticles
MedChemComm 2011, 2, 1033– 1049 DOI: 10.1039/c1md00158b [Crossref], [CAS], [Google Scholar]
  

1225. Zhou, Y.; Tan, L. L.; Li, Q. L.; Qiu, X. L.; Qi, A. D.; Tao, Y. C.; Yang, Y. W. Acetylcholine-Triggered Cargo
Release from Supramolecular Nanovalves Based on Different Macrocyclic Receptors Chem. - Eur. J. 2014, 20,
2998– 3004 DOI: 10.1002/chem.201304864 [Crossref], [PubMed], [CAS], [Google Scholar]

1226. Hernandez, R.; Tseng, H. R.; Wong, J. W.; Stoddart, J. F.; Zink, J. I. An operational supramolecular
nanovalve J. Am. Chem. Soc. 2004, 126, 3370– 3371 DOI: 10.1021/ja039424u [ACS Full Text ], [CAS], [Google
Scholar]

1227. Unimolecular and Supramolecular Electronics II: Chemistry and Physics Meet at Metal-Molecule
Interfaces; 2012; Vol. 313, pp 1– 272. [Google Scholar]

1228. Joachim, C. Molecular wires and logic circuits integration in a single molecule? Nanotechnology 2003, 2,
80– 81 [Google Scholar]

1229. Joachim, C. Molecular and intramolecular electronics Superlattices Microstruct. 2000, 28, 305– 315
 DOI: 10.1006/spmi.2000.0918 [Crossref], [CAS], [Google Scholar]

1230. Joachim, C.; Ratner, M. A. Molecular electronics Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 8800– 8800
 DOI: 10.1073/pnas.0504046102 [Crossref], [Google Scholar]

1231. Schenning, A. P. H. J.; Jonkheijm, P.; Hoeben, F. J. M.; van Herrikhuyzen, J.; Meskers, S. C. J.; Meijer, E. W.;
Herz, L. M.; Daniel, C.; Silva, C.; Phillips, R. T., et al. Towards supramolecular electronics Synth. Met. 2004, 147,
43– 48 DOI: 10.1016/j.synthmet.2004.06.038 [Crossref], [CAS], [Google Scholar]

1232. Weiss, J. Supramolecular approaches to nano and molecular electronics Coord. Chem. Rev. 2010, 254,
2247– 2248 DOI: 10.1016/j.ccr.2010.06.002 [Crossref], [CAS], [Google Scholar]

1233. Flood, A. H.; Stoddart, J. F.; Steuerman, D. W.; Heath, J. R. Chemistry. Whence molecular electronics?
Science 2004, 306, 2055– 2056 DOI: 10.1126/science.1106195 [Crossref], [PubMed], [CAS], [Google Scholar]

1234. Blanco, M. J.; Jimenez, M. C.; Chambron, J. C.; Heitz, V.; Linke, M.; Sauvage, J.-P. Rotaxanes as new
architectures for photoinduced electron transfer and molecular motions Chem. Soc. Rev. 1999, 28, 293– 305
 DOI: 10.1039/a901205b [Crossref], [CAS], [Google Scholar]
 -
muscles
1235. Huang, T. J. Towards arti cial molecular motor-based electroactive/photoactive biomimetic
art. no. 65240H P. Soc. Photo. Opt. Ins. 2007, 6524, H5240– H5240 DOI: 10.1117/12.718483 [Crossref], [Google
Scholar]

1236. Juluri, B. K.; Kumar, A. S.; Liu, Y.; Ye, T.; Yang, Y. W.; Flood, A. H.; Fang, L.; Stoddart, J. F.; Weiss, P. S.;
Huang, T. J. A Mechanical Actuator Driven Electrochemically by Arti cial Molecular Muscles ACS Nano 2009, 3,
291– 300 DOI: 10.1021/nn8002373 [ACS Full Text ], [CAS], [Google Scholar]

1237. Li, D. B.; Paxton, W. F.; Baughman, R. H.; Huang, T. J.; Stoddart, J. F.; Weiss, P. S. Molecular,
Supramolecular, and Macromolecular Motors and Arti cial Muscles MRS Bull. 2009, 34, 671– 681
 DOI: 10.1557/mrs2009.179 [Crossref], [CAS], [Google Scholar]

1238. Liu, Y.; Flood, A. H.; Bonvallett, P. A.; Vignon, S. A.; Northrop, B. H.; Tseng, H. R.; Jeppesen, J. O.; Huang, T.
J.; Brough, B.; Baller, M.; Magonov, S.; Solares, S. D.; Goddard, W. A., III; Ho, C. M.; Stoddart, J. F. Linear arti cial
molecular muscles J. Am. Chem. Soc. 2005, 127, 9745– 9759 DOI: 10.1021/ja051088p [ACS Full Text ],
[CAS], [Google Scholar]

1239. Park, J. K.; Carr, J.; Calhoun, B.; Moore, R. B. Enhanced actuation in arti cial muscles through supra-
molecular orientation of ionic domains Polym. Prepr. 2006, 47, 484– 485 [CAS], [Google Scholar]

1240. Otero, T. F.; Grande, H.; Rodriguez, J. Reversible electrochemical reactions in conducting polymers: A
molecular approach to arti cial muscles J. Phys. Org. Chem. 1996, 9, 381– 386
 DOI: 10.1002/(SICI)1099-1395(199606)9:6<381::AID-POC796>3.0.CO;2-N [Crossref], [CAS], [Google Scholar]

1241. Valero, L.; Arias-Pardilla, J.; Cauich-Rodriguez, J.; Smit, M. A.; Otero, T. F. Characterization of the
movement of polypyrrole-dodecylbenzenesulfonate-perchlorate/tape arti cial muscles. Faradaic control of
reactive arti cial molecular motors and muscles Electrochim. Acta 2011, 56, 3721– 3726
 DOI: 10.1016/j.electacta.2010.11.058 [Crossref], [CAS], [Google Scholar]

1242. Takashima, Y.; Hatanaka, S.; Otsubo, M.; Nakahata, M.; Kakuta, T.; Hashidzume, A.; Yamaguchi, H.;
Harada, A. Expansion-contraction of photoresponsive arti cial muscle regulated by host-guest interactions Nat.
Commun. 2012, 3, 1270 DOI: 10.1038/ncomms2280 [Crossref], [PubMed], [CAS], [Google Scholar]

1243. Ceroni, P.; Credi, A.; Venturi, M. Light to investigate (read) and operate (write) molecular devices and
machines Chem. Soc. Rev. 2014, 43, 4068– 4083 DOI: 10.1039/c3cs60400d [Crossref], [PubMed], [CAS], [Google
Scholar]
1244. Hutchison, K. A.; Parakka, J. P.; Kesler, B. S.; Schumaker, R. R. Chiropticenes: Molecular chiroptical
  
dipole
switches for optical data storage Micro-Nano-Photonic Mater. Dev. 2000, 3937, 64– 72 DOI: 10.1117/12.382795
[Crossref], [Google Scholar]

1245. Berna, J.; Leigh, D. A.; Lubomska, M.; Mendoza, S. M.; Perez, E. M.; Rudolf, P.; Teobaldi, G.; Zerbetto, F.
Macroscopic transport by synthetic molecular machines Nat. Mater. 2005, 4, 704– 710 DOI: 10.1038/nmat1455
[Crossref], [PubMed], [CAS], [Google Scholar]

1246. Silvi, S.; Arduini, A.; Pochini, A.; Secchi, A.; Tomasulo, M.; Raymo, F. M.; Baroncini, M.; Credi, A. A simple
molecular machine operated by photoinduced proton transfer J. Am. Chem. Soc. 2007, 129, 13378– 13379
 DOI: 10.1021/ja0753851 [ACS Full Text ], [CAS], [Google Scholar]

1247. Silvi, S.; Constable, E. C.; Housecroft, C. E.; Beves, J. E.; Dunphy, E. L.; Tomasulo, M.; Raymo, F. M.; Credi, A.
Photochemical switching of luminescence and singlet oxygen generation by chemical signal communication
Chem. Commun. 2009, 1484– 1486 DOI: 10.1039/b900712a [Crossref], [PubMed], [CAS], [Google Scholar]

1248. Tatum, L. A.; Foy, J. T.; Aprahamian, I. Waste Management of Chemically Activated Switches: Using a
Photoacid To Eliminate Accumulation of Side Products J. Am. Chem. Soc. 2014, 136, 17438– 17441
 DOI: 10.1021/ja511135k [ACS Full Text ], [CAS], [Google Scholar]

1249. Jousselme, B.; Blanchard, P.; Levillain, E.; Delaunay, J.; Allain, M.; Richomme, P.; Rondeau, D.; Gallego-
Planas, N.; Roncali, J. Crown-annelated oligothiophenes as model compounds for molecular actuation J. Am.
Chem. Soc. 2003, 125, 1363– 1370 DOI: 10.1021/ja026819p [ACS Full Text ], [CAS], [Google Scholar]

1250. Roncali, J.; Garreau, R.; Delabouglise, D.; Garnier, F.; Lemaire, M. Modi cation of the Structure and
Electrochemical Properties of Poly(Thiophene) by Ether Groups J. Chem. Soc., Chem. Commun. 1989, 679– 681
 DOI: 10.1039/c39890000679 [Crossref], [CAS], [Google Scholar]

1251. Roncali, J.; Shi, L. H.; Garnier, F. Effects of Environmental-Factors on the Electrooptical Properties of
Conjugated Polymers Containing Oligo(Oxyethylene) Substituents J. Phys. Chem. 1991, 95, 8983– 8989
 DOI: 10.1021/j100175a102 [ACS Full Text ], [CAS], [Google Scholar]

1252. Marsella, M. J.; Swager, T. M. Designing Conducting Polymer-Based Sensors - Selective Ionochromic
Response in Crown-Ether Containing Polythiophenes J. Am. Chem. Soc. 1993, 115, 12214– 12215
 DOI: 10.1021/ja00078a090 [ACS Full Text ], [CAS], [Google Scholar]
1253. Wang, B.; Wasielewski, M. R. Design and synthesis of metal ion-recognition-induced conjugated
polymers: An approach to metal ion sensory materials J. Am. Chem. Soc. 1997, 119, 12– 21   
 DOI: 10.1021/ja962229d [ACS Full Text ], [CAS], [Google Scholar]

1254. Marsella, M. J.; Newland, R. J.; Carroll, P. J.; Swager, T. M. Ionoresistivity as a Highly Sensitive Sensory
Probe - Investigations of Polythiophenes Functionalized with Calix[4]Arene-Based Ion Receptors J. Am. Chem.
Soc. 1995, 117, 9842– 9848 DOI: 10.1021/ja00144a009 [ACS Full Text ], [CAS], [Google Scholar]

1255. Roncali, J. Electrogenerated functional conjugated polymers as advanced electrode materials J. Mater.
Chem. 1999, 9, 1875– 1893 DOI: 10.1039/a902747e [Crossref], [CAS], [Google Scholar]

1256. McQuade, D. T.; Pullen, A. E.; Swager, T. M. Conjugated polymer-based chemical sensors Chem. Rev.
2000, 100, 2537– 2574 DOI: 10.1021/cr9801014 [ACS Full Text ], [CAS], [Google Scholar]

1257. Barbarella, G.; Melucci, M.; Sotgiu, G. The versatile thiophene: An overview of recent research on
thiophene-based materials Adv. Mater. 2005, 17, 1581– 1593 DOI: 10.1002/adma.200402020 [Crossref],
[CAS], [Google Scholar]

1258. Jousselme, B.; Blanchard, P.; Gallego-Planas, N.; Delaunay, J.; Allain, M.; Richomme, P.; Levillain, E.;
Roncali, J. Photomechanical actuation and manipulation of the electronic properties of linear π-conjugated
systems J. Am. Chem. Soc. 2003, 125, 2888– 2889 DOI: 10.1021/ja029754z [ACS Full Text ], [CAS], [Google
Scholar]

1259. Jousselme, B.; Blanchard, P.; Gallego-Planas, N.; Levillain, E.; Delaunay, J.; Allain, M.; Richomme, P.;
Roncali, J. Photomechanical control of the electronic properties of linear π-conjugated systems Chem. - Eur. J.
2003, 9, 5297– 5306 DOI: 10.1002/chem.200305010 [Crossref], [PubMed], [CAS], [Google Scholar]

1260. Clayden, J.; Lund, A.; Vallverdu, L. S.; Helliwell, M. Ultra-remote stereocontrol by conformational
communication of information along a carbon chain Nature 2004, 431, 966– 971 DOI: 10.1038/nature02933
[Crossref], [PubMed], [CAS], [Google Scholar]

1261. Barton, D. H. R.; Cookson, R. C. The Principles of Conformational Analysis Q. Rev., Chem. Soc. 1956, 10,
44– 82 DOI: 10.1039/qr9561000044 [Crossref], [CAS], [Google Scholar]

1262. Browne, W. R.; Pollard, M. M.; de Lange, B.; Meetsma, A.; Feringa, B. L. Reversible three-state switching of
luminescence: a new twist to electro- and photochromic behavior J. Am. Chem. Soc. 2006, 128, 12412– 12413
 DOI: 10.1021/ja064423y [ACS Full Text ], [CAS], [Google Scholar]
  
1263. Nilsson, J. R.; O’Sullivan, M. C.; Li, S.; Anderson, H. L.; Andreasson, J. A photoswitchable supramolecular
complex with release-and-report capabilities Chem. Commun. 2015, 51, 847– 850 DOI: 10.1039/C4CC08513B
[Crossref], [PubMed], [CAS], [Google Scholar]

1264. Zigon, N.; Larpent, P.; Jouaiti, A.; Kyritsakas, N.; Hosseini, M. W. Optical reading of the open and closed
states of a molecular turnstile Chem. Commun. 2014, 50, 5040– 5042 DOI: 10.1039/c4cc01071j [Crossref],
[PubMed], [CAS], [Google Scholar]

1265. Dube, H.; Ams, M. R.; Rebek, J. Supramolecular Control of Fluorescence through Reversible
Encapsulation J. Am. Chem. Soc. 2010, 132, 9984– 9985 DOI: 10.1021/ja103912a [ACS Full Text ],
[CAS], [Google Scholar]

1266. Tzeli, D.; Theodorakopoulos, G.; Petsalakis, I. D.; Ajami, D.; Rebek, J. Conformations and Fluorescence of
Encapsulated Stilbene J. Am. Chem. Soc. 2012, 134, 4346– 4354 DOI: 10.1021/ja211164b [ACS Full Text ],
[CAS], [Google Scholar]

1267. Masar, M. S.; Gianneschi, N. C.; Oliveri, C. G.; Stern, C. L.; Nguyen, S. T.; Mirkin, C. A. Allosterically regulated
supramolecular catalysis of acyl transfer reactions for signal ampli cation and detection of small molecules J.
Am. Chem. Soc. 2007, 129, 10149– 10158 DOI: 10.1021/ja0711516 [ACS Full Text ], [CAS], [Google Scholar]

1268. Riddle, J. A.; Jiang, X.; Huffman, J.; Lee, D. Signal-amplifying resonance energy transfer: A dynamic
multichromophore array for allosteric switching Angew. Chem., Int. Ed. 2007, 46, 7019– 7022
 DOI: 10.1002/anie.200701410 [Crossref], [PubMed], [CAS], [Google Scholar]

1269. Jiang, X.; Bollinger, J. C.; Lee, D. Two-dimensional electronic conjugation: Cooperative folding and
uorescence switching J. Am. Chem. Soc. 2006, 128, 11732– 11733 DOI: 10.1021/ja063413u [ACS Full Text ],
[CAS], [Google Scholar]

1270. Riddle, J. A.; Jiang, X.; Lee, D. W. Conformational dynamics for chemical sensing: simplicity and diversity
Analyst 2008, 133, 417– 422 DOI: 10.1039/b715673c [Crossref], [PubMed], [CAS], [Google Scholar]

1271. Opsitnick, E.; Lee, D. Two-dimensional electronic conjugation: Statics and dynamics at structural
domains beyond molecular wires Chem. - Eur. J. 2007, 13, 7041– 7049 DOI: 10.1002/chem.200700813
[Crossref], [Google Scholar]
1272. Onagi, H.; Rebek, J. Fluorescence resonance energy transfer across a mechanical bond of a rotaxane
 
Chem. Commun. 2005, 4604– 4606 DOI: 10.1039/b506177f [Crossref], [PubMed], [CAS], [Google Scholar] 

1273. Azov, V. A.; Schlegel, A.; Diederich, F. Geometrically precisely de ned multinanometer
expansion/contraction motions in a resorcin[4]arene cavitand based molecular switch Angew. Chem., Int. Ed.
2005, 44, 4635– 4638 DOI: 10.1002/anie.200500970 [Crossref], [PubMed], [CAS], [Google Scholar]

1274. Li, Y. J.; Li, H.; Li, Y. L.; Liu, H. B.; Wang, S.; He, X. R.; Wang, N.; Zhu, D. B. Energy transfer switching in a
bistable molecular machine Org. Lett. 2005, 7, 4835– 4838 DOI: 10.1021/ol051567t [ACS Full Text ],
[CAS], [Google Scholar]

1275. Raker, J.; Glass, T. E. Cooperative ratiometric chemosensors: Pinwheel receptors with an integrated
uorescence system J. Org. Chem. 2001, 66, 6505– 6512 DOI: 10.1021/jo001775t [ACS Full Text ],
[CAS], [Google Scholar]

1276. Barboiu, M.; Prodi, L.; Montalti, M.; Zaccheroni, N.; Kyritsakas, N.; Lehn, J.-M. Dynamic chemical devices:
Modulation of photophysical properties by reversible, ion-triggered, and proton-fueled nanomechanical shape-
ipping molecular motions Chem. - Eur. J. 2004, 10, 2953– 2959 DOI: 10.1002/chem.200306045 [Crossref],
[PubMed], [CAS], [Google Scholar]

1277. Zhong, Z.; Zhao, Y. Cholate-glutamic acid hybrid foldamer and its uorescent detection of Zn2+ Org. Lett.
2007, 9, 2891– 2894 DOI: 10.1021/ol071130g [ACS Full Text ], [CAS], [Google Scholar]

1278. Kim, U. I.; Suk, J. M.; Naidu, V. R.; Jeong, K. S. Folding and anion-binding properties of uorescent
oligoindole foldamers Chem. - Eur. J. 2008, 14, 11406– 11414 DOI: 10.1002/chem.200801713 [Crossref],
[PubMed], [CAS], [Google Scholar]

1279. Yuasa, H.; Miyagawa, N.; Izumi, T.; Nakatani, M.; Izumi, M.; Hashimoto, H. Hinge sugar as a movable
component of an excimer uorescence sensor Org. Lett. 2004, 6, 1489– 1492 DOI: 10.1021/ol049628v [ACS Full
Text ], [CAS], [Google Scholar]

1280. Monahan, C.; Bien, J. T.; Smith, B. D. Fluorescence sensing due to allosteric switching of pyrene
functionalized cis-cyclohexane-1,3-dicarboxylate Chem. Commun. 1998, 431– 432 DOI: 10.1039/a705445i
[Crossref], [CAS], [Google Scholar]

1281. Krauss, R.; Weinig, H. G.; Seydack, M.; Bendig, J.; Koert, U. Molecular signal transduction through
conformational transmission of a perhydroanthracene transducer Angew. Chem., Int. Ed. 2000, 39, 1835– 1837
 DOI: 10.1002/(SICI)1521-3773(20000515)39:10<1835::AID-ANIE1835>3.0.CO;2-S [Crossref], [PubMed],
[CAS], [Google Scholar]   

1282. Koert, U.; Krauss, R.; Weinig, H. G.; Heumann, C.; Ziemer, B.; Mugge, C.; Seydack, M.; Bendig, J. 2,3,6,7-
tetrasubstituted decalins: Biconformational transducers for molecular signal transduction Eur. J. Org. Chem.
2001, 2001, 575– 586 DOI: 10.1002/1099-0690(200102)2001:3<575::AID-EJOC575>3.0.CO;2-7
[Crossref], [Google Scholar]

1283. Berna, J.; Franco-Pujante, C.; Alajarin, M. Competitive binding for triggering a uorescence response in a
hydrazodicarboxamide-based [2]rotaxane Org. Biomol. Chem. 2014, 12, 474– 478 DOI: 10.1039/C3OB41807C
[Crossref], [PubMed], [CAS], [Google Scholar]

1284. Zhou, W.; Wu, Y.; Zhai, B. Q.; Wang, Q. C.; Qu, D. H. An anthracene-containing bistable [2]rotaxane featuring
color and uorescence changes RSC Adv. 2014, 4, 5148– 5151 DOI: 10.1039/c3ra46517a [Crossref],
[CAS], [Google Scholar]

1285. Perez, E. M.; Dryden, D. T. F.; Leigh, D. A.; Teobaldi, G.; Zerbetto, F. A generic basis for some simple light-
operated mechanical molecular machines J. Am. Chem. Soc. 2004, 126, 12210– 12211 DOI: 10.1021/ja0484193
[ACS Full Text ], [CAS], [Google Scholar]

1286. Buschel, M.; Helldobler, M.; Daub, J. Electronic coupling in 6,6″-donor-substituted terpyridines: tuning of
the mixed valence state by proton and metal ion complexation Chem. Commun. 2002, 1338– 1339
 DOI: 10.1039/b110289c [Crossref], [Google Scholar]

1287. Glass, T. E. Cooperative chemical sensing with bis-tritylacetylenes: Pinwheel receptors with metal ion
recognition properties J. Am. Chem. Soc. 2000, 122, 4522– 4523 DOI: 10.1021/ja994398e [ACS Full Text ],
[CAS], [Google Scholar]

1288. Raker, J.; Glass, T. E. Selectivity via cooperative interactions: Detection of dicarboxylates in water by a
pinwheel chemosensor J. Org. Chem. 2002, 67, 6113– 6116 DOI: 10.1021/jo025903k [ACS Full Text ],
[CAS], [Google Scholar]

1289. Karle, M.; Bockelmann, D.; Schumann, D.; Griesinger, C.; Koert, U. Conformative coupling of two
conformational molecular switches Angew. Chem., Int. Ed. 2003, 42, 4546– 4549 DOI: 10.1002/anie.200352130
[Crossref], [PubMed], [CAS], [Google Scholar]
1290. Zhao, Y.; Zhong, Z. Detection of Hg2+ in aqueous solutions with a foldamer-based uorescent sensor
Text],
modulated by surfactant micelles Org. Lett. 2006, 8, 4715– 4717 DOI: 10.1021/ol061735x [ACS Full 
[CAS], [Google Scholar]

1291. Qu, D. H.; Wang, Q. C.; Tian, H. A half adder based on a photochemically driven [2]rotaxane Angew. Chem.,
Int. Ed. 2005, 44, 5296– 5299 DOI: 10.1002/anie.200501215 [Crossref], [PubMed], [CAS], [Google Scholar]

1292. Li, H.; Zhang, J. N.; Zhou, W.; Zhang, H.; Zhang, Q.; Qu, D. H.; Tian, H. Dual-Mode Operation of a Bistable
[1]Rotaxane with a Fluorescence Signal Org. Lett. 2013, 15, 3070– 3073 DOI: 10.1021/ol401251u [ACS Full Text
], [CAS], [Google Scholar]

1293. Qu, D. H.; Wang, Q. C.; Tian, H. Photodriven and thermal-driven shuttling of alpha-cyclodextrin on the
molecular rotaxane containing azobenzene Mol. Cryst. Liq. Cryst. 2005, 430, 59– 65
 DOI: 10.1080/15421400590946172 [Crossref], [CAS], [Google Scholar]

1294. Qu, D. H.; Wang, Q. C.; Ma, X.; Tian, H. A [3]Rotaxane with three stable states that responds to multiple-
inputs and displays dual uorescence addresses Chem. - Eur. J. 2005, 11, 5929– 5937
 DOI: 10.1002/chem.200401313 [Crossref], [PubMed], [CAS], [Google Scholar]

1295. Lin, Y. C.; Chen, C. T. Alkaline Earth Metal Ion Induced CoilHelixCoil Transition of
LysineCoumarinAzacrown Hybrid Foldamers with OFFOFFON Fluorescence Switching Chem. - Eur. J. 2013, 19,
2531– 2538 DOI: 10.1002/chem.201202998 [Crossref], [PubMed], [CAS], [Google Scholar]

1296. Coskun, A.; Friedman, D. C.; Li, H.; Patel, K.; Khatib, H. A.; Stoddart, J. F. A Light-Gated STOP–GO
Molecular Shuttle J. Am. Chem. Soc. 2009, 131, 2493– 2495 DOI: 10.1021/ja809225e [ACS Full Text ],
[CAS], [Google Scholar]

1297. Cao, J. J.; Ma, X.; Min, M. R.; Cao, T. T.; Wu, S. F.; Tian, H. INHIBIT logic operations based on light-driven
beta-cyclodextrin pseudo[1]rotaxane with room temperature phosphorescence addresses Chem. Commun.
2014, 50, 3224– 3226 DOI: 10.1039/c3cc49820d [Crossref], [PubMed], [CAS], [Google Scholar]

1298. Brown, R. A.; Diemer, V.; Webb, S. J.; Clayden, J. End-to-end conformational communication through a
synthetic purinergic receptor by ligand-induced helicity switching Nat. Chem. 2013, 5, 853– 860
 DOI: 10.1038/nchem.1747 [Crossref], [PubMed], [CAS], [Google Scholar]

1299. Muraoka, T.; Kinbara, K.; Aida, T. A self-locking molecule operative with a photoresponsive key J. Am.
Chem. Soc. 2006, 128, 11600– 11605 DOI: 10.1021/ja0632308 [ACS Full Text ], [CAS], [Google Scholar]
 
 Chem.
1300. Clayden, J.; Vassiliou, N. Stereochemical relays: communication via conformation Org. Biomol.
2006, 4, 2667– 2678 DOI: 10.1039/b604548k [Crossref], [PubMed], [CAS], [Google Scholar]

1301. Jiang, X.; Lim, Y. K.; Zhang, B. J.; Opsitnick, E. A.; Baik, M. H.; Lee, D. Dendritic molecular switch: chiral
folding and helicity inversion J. Am. Chem. Soc. 2008, 130, 16812– 16822 DOI: 10.1021/ja806723e [ACS Full
Text ], [CAS], [Google Scholar]

1302. Clayden, J.; Castellanos, A.; Sola, J.; Morris, G. A. Quantifying End-to-End Conformational Communication
of Chirality through an Achiral Peptide Chain Angew. Chem., Int. Ed. 2009, 48, 5962– 5965
 DOI: 10.1002/anie.200901892 [Crossref], [PubMed], [CAS], [Google Scholar]

1303. Mathews, M.; Tamaoki, N. Reversibly tunable helicity induction and inversion in liquid crystal self-
assembly by a planar chiroptic trigger molecule Chem. Commun. 2009, 3609– 3611 DOI: 10.1039/b905305k
[Crossref], [PubMed], [CAS], [Google Scholar]

1304. Akine, S.; Hotate, S.; Nabeshima, T. A molecular leverage for helicity control and helix inversion J. Am.
Chem. Soc. 2011, 133, 13868– 13871 DOI: 10.1021/ja205570z [ACS Full Text ], [CAS], [Google Scholar]

1305. Wei, K.; Ni, J.; Min, Y.; Chen, S.; Liu, Y. Unexpected helicity control and helix inversion: homochiral helical
nanotubes consisting of an achiral ligand Chem. Commun. 2013, 49, 8220– 8222 DOI: 10.1039/c3cc43898h
[Crossref], [PubMed], [CAS], [Google Scholar]

1306. Sargsyan, G.; Schatz, A. A.; Kubelka, J.; Balaz, M. Formation and helicity control of ssDNA templated
porphyrin nanoassemblies Chem. Commun. 2013, 49, 1020– 1022 DOI: 10.1039/C2CC38150H [Crossref],
[PubMed], [CAS], [Google Scholar]

1307. De Poli, M.; Clayden, J. Thionoglycine as a multifunctional spectroscopic reporter of screw-sense


preference in helical foldamers Org. Biomol. Chem. 2014, 12, 836– 843 DOI: 10.1039/C3OB42167H [Crossref],
[PubMed], [CAS], [Google Scholar]

1308. Duan, P.; Cao, H.; Zhang, L.; Liu, M. Gelation induced supramolecular chirality: chirality transfer,
ampli cation and application Soft Matter 2014, 10, 5428– 5448 DOI: 10.1039/C4SM00507D [Crossref],
[PubMed], [CAS], [Google Scholar]
1309. Nguyen, B. T.; Anslyn, E. V. Indicator-displacement assays Coord. Chem. Rev. 2006, 250, 3118– 3127
 DOI: 10.1016/j.ccr.2006.04.009 [Crossref], [CAS], [Google Scholar]   

1310. Gassensmith, J. J.; Matthys, S.; Lee, J. J.; Wojcik, A.; Kamat, P. V.; Smith, B. D. Squaraine Rotaxane as a
Reversible Optical Chloride Sensor Chem. - Eur. J. 2010, 16, 2916– 2921 DOI: 10.1002/chem.200902547
[Crossref], [PubMed], [CAS], [Google Scholar]

1311. Langton, M. J.; Beer, P. D. Rotaxane and catenane host structures for sensing charged guest species Acc.
Chem. Res. 2014, 47, 1935– 1949 DOI: 10.1021/ar500012a [ACS Full Text ], [CAS], [Google Scholar]

1312. Hsueh, S. Y.; Lai, C. C.; Chiu, S. H. Squaraine-Based [2]Rotaxanes that Function as Visibly Active Molecular
Switches Chem. - Eur. J. 2010, 16, 2997– 3000 DOI: 10.1002/chem.200903304 [Crossref], [PubMed],
[CAS], [Google Scholar]

1313. Klajn, R.; Fang, L.; Coskun, A.; Olson, M. A.; Wesson, P. J.; Stoddart, J. F.; Grzybowski, B. A. Metal
Nanoparticles Functionalized with Molecular and Supramolecular Switches J. Am. Chem. Soc. 2009, 131, 4233–
4235 DOI: 10.1021/ja9001585 [ACS Full Text ], [CAS], [Google Scholar]

1314. Cheng, H. B.; Zhang, H. Y.; Liu, Y. Dual-Stimulus Luminescent Lanthanide Molecular Switch Based on an
Unsymmetrical Diarylper uorocyclopentene J. Am. Chem. Soc. 2013, 135, 10190– 10193
 DOI: 10.1021/ja4018804 [ACS Full Text ], [CAS], [Google Scholar]

1315. Allenmark, S. Induced circular dichroism by chiral molecular interaction Chirality 2003, 15, 409– 422
 DOI: 10.1002/chir.10220 [Crossref], [PubMed], [CAS], [Google Scholar]

1316. Prins, L. J.; Huskens, J.; de Jong, F.; Timmerman, P.; Reinhoudt, D. N. Complete asymmetric induction of
supramolecular chirality in a hydrogen-bonded assembly Nature 1999, 398, 498– 502 DOI: 10.1038/19053
[Crossref], [CAS], [Google Scholar]

1317. Mineo, P.; Villari, V.; Scamporrino, E.; Micali, N. Supramolecular chirality induced by a weak thermal force
Soft Matter 2014, 10, 44– 47 DOI: 10.1039/C3SM52322E [Crossref], [PubMed], [CAS], [Google Scholar]

1318. Tachibana, Y.; Kihara, N.; Takata, T. Asymmetric benzoin condensation catalyzed by chiral rotaxanes
tethering a thiazolium salt moiety via the cooperation of the component: Can rotaxane be an effective reaction
eld? J. Am. Chem. Soc. 2004, 126, 13560– 13560 DOI: 10.1021/ja0408202 [ACS Full Text ], [CAS], [Google
Scholar]
1319. Havinga, E. Spontaneous Formation of Optically Active Substances Biochim. Biophys. Acta 1954, 13,
171– 174 DOI: 10.1016/0006-3002(54)90300-5 [Crossref], [PubMed], [CAS], [Google Scholar]   

1320. Suarez, M.; Branda, N.; Lehn, J.-M.; Decian, A.; Fischer, J. Supramolecular chirality: Chiral hydrogen-
bonded supermolecules from achiral molecular components Helv. Chim. Acta 1998, 81, 1– 13
 DOI: 10.1002/hlca.19980810102 [Crossref], [CAS], [Google Scholar]

1321. Evan-Salem, T.; Baruch, I.; Avram, L.; Cohen, Y.; Palmer, L. C.; Rebek, J. Resorcinarenes are hexameric
capsules in solution Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 12296– 12300 DOI: 10.1073/pnas.0604757103
[Crossref], [PubMed], [CAS], [Google Scholar]

1322. Rivera, J. M.; Martin, T.; Rebek, J. Chiral spaces: Dissymmetric capsules through self-assembly Science
1998, 279, 1021– 1023 DOI: 10.1126/science.279.5353.1021 [Crossref], [PubMed], [CAS], [Google Scholar]

1323. Woods, C. R.; Benaglia, M.; Cozzi, F.; Siegel, J. S. Enantioselective synthesis of copper(I) bipyridine based
helicates by chiral templating of secondary structure: Transmission of stereochemistry on the nanometer scale
Angew. Chem., Int. Ed. Engl. 1996, 35, 1830– 1833 DOI: 10.1002/anie.199618301 [Crossref], [CAS], [Google
Scholar]

1324. Kaminker, R.; de Hatten, X.; Lahav, M.; Lupo, F.; Gulino, A.; Evmenenko, G.; Dutta, P.; Browne, C.; Nitschke,
J. R.; van der Boom, M. E. Assembly of Surface-Con ned Homochiral Helicates: Chiral Discrimination of DOPA
and Unidirectional Charge Transfer J. Am. Chem. Soc. 2013, 135, 17052– 17059 DOI: 10.1021/ja4077205 [ACS
Full Text ], [CAS], [Google Scholar]

1325. Miyake, H.; Tsukube, H. Coordination chemistry strategies for dynamic helicates: time-programmable
chirality switching with labile and inert metal helicates Chem. Soc. Rev. 2012, 41, 6977– 6991
 DOI: 10.1039/c2cs35192g [Crossref], [PubMed], [CAS], [Google Scholar]

1326. Bottari, G.; Leigh, D. A.; Perez, E. M. Chiroptical switching in a bistable molecular shuttle J. Am. Chem.
Soc. 2003, 125, 13360– 13361 DOI: 10.1021/ja036665t [ACS Full Text ], [CAS], [Google Scholar]

1327. Zhou, W. D.; Xu, J. L.; Zheng, H. Y.; Yin, X. D.; Zuo, Z. C.; Liu, H. B.; Li, Y. L. Distinct Nanostructures from a
Molecular Shuttle: Effects of Shuttling Movement on Nanostructural Morphologies Adv. Funct. Mater. 2009, 19,
141– 149 DOI: 10.1002/adfm.200801149 [Crossref], [CAS], [Google Scholar]

1328. Ji, F. Y.; Zhu, L. L.; Zhang, D.; Chen, Z. F.; Tian, H. Coordination-driven self-organization of switchable
[2]rotaxane Tetrahedron 2009, 65, 9081– 9085 DOI: 10.1016/j.tet.2009.09.051 [Crossref], [CAS], [Google Scholar]
1329. Hsueh, S. Y.; Kuo, C. T.; Lu, T. W.; Lai, C. C.; Liu, Y. H.; Hsu, H. F.; Peng, S. M.; Chen, C. H.; Chiu,
S. H.
 
Acid/Base- and Anion-Controllable Organogels Formed From a Urea-Based Molecular Switch Angew. Chem., Int.
Ed. 2010, 49, 9170– 9173 DOI: 10.1002/anie.201004090 [Crossref], [PubMed], [CAS], [Google Scholar]

1330. Gong, H.-Y.; Rambo, B. M.; Karnas, E.; Lynch, V. M.; Sessler, J. L. A ‘Texas-sized’ molecular box that forms
an anion-induced supramolecular necklace Nat. Chem. 2010, 2, 406– 409 DOI: 10.1038/nchem.597 [Crossref],
[PubMed], [CAS], [Google Scholar]

1331. Fasano, V.; Baroncini, M.; Moffa, M.; Iandolo, D.; Camposeo, A.; Credi, A.; Pisignano, D. Organic Nano bers
Embedding Stimuli-Responsive Threaded Molecular Components J. Am. Chem. Soc. 2014, 136, 14245– 14254
 DOI: 10.1021/ja5080322 [ACS Full Text ], [CAS], [Google Scholar]

1332. Jimenez, M. C.; Dietrich-Buchecker, C.; Sauvage, J.-P. Towards synthetic molecular muscles: Contraction
and stretching of a linear rotaxane dimer Angew. Chem., Int. Ed. 2000, 39, 3284– 3287
 DOI: 10.1002/1521-3773(20000915)39:18<3284::AID-ANIE3284>3.0.CO;2-7 [Crossref], [PubMed], [CAS], [Google
Scholar]

1333. Gao, L.; Zhang, Z.; Zheng, B.; Huang, F. Construction of muscle-like metallo-supramolecular polymers
from a pillar[5]arene-based [c2]daisy chain Polym. Chem. 2014, 5, 5734– 5739 DOI: 10.1039/C4PY00733F
[Crossref], [CAS], [Google Scholar]

1334. Yan, X.; Zheng, B.; Huang, F. Integrated motion of molecular machines in supramolecular polymeric
scaffolds Polym. Chem. 2013, 4, 2395– 2399 DOI: 10.1039/c3py00060e [Crossref], [CAS], [Google Scholar]

1335. Bruns, C. J.; Frasconi, M.; Iehl, J.; Hartlieb, K. J.; Schneebeli, S. T.; Cheng, C.; Stupp, S. I.; Stoddart, J. F.
Redox Switchable Daisy Chain Rotaxanes Driven by Radical–Radical Interactions J. Am. Chem. Soc. 2014, 136,
4714– 4723 DOI: 10.1021/ja500675y [ACS Full Text ], [CAS], [Google Scholar]

1336. Silvi, S.; Venturi, M.; Credi, A. Arti cial molecular shuttles: from concepts to devices J. Mater. Chem. 2009,
19, 2279– 2294 DOI: 10.1039/b818609j [Crossref], [CAS], [Google Scholar]

1337. Balzani, V.; Credi, A.; Silvi, S.; Venturi, M. Arti cial nanomachines based on interlocked molecular species:
recent advances Chem. Soc. Rev. 2006, 35, 1135– 1149 DOI: 10.1039/b517102b [Crossref], [PubMed],
[CAS], [Google Scholar]
1338. Romuald, C.; Arda, A.; Clavel, C.; Jimenez-Barbero, J.; Coutrot, F. Tightening or loosening a pH-sensitive
double-lasso molecular machine readily synthesized from an ends-activated [c2]daisy chain Chem.  3,
Sci. 2012,
1851– 1857 DOI: 10.1039/c2sc20072d [Crossref], [CAS], [Google Scholar]

1339. Fang, L.; Hmadeh, M.; Wu, J.; Olson, M. A.; Spruell, J. M.; Trabolsi, A.; Yang, Y.-W.; Elhabiri, M.; Albrecht-
Gary, A.-M.; Stoddart, J. F. Acid–Base Actuation of [c2]Daisy Chains J. Am. Chem. Soc. 2009, 131, 7126– 7134
 DOI: 10.1021/ja900859d [ACS Full Text ], [CAS], [Google Scholar]

1340. Dawson, R. E.; Lincoln, S. F.; Easton, C. J. The foundation of a light driven molecular muscle based on
stilbene and alpha-cyclodextrin Chem. Commun. 2008, 3980– 3982 DOI: 10.1039/b809014a [Crossref],
[PubMed], [CAS], [Google Scholar]

1341. Huang, T. J.; Brough, B.; Ho, C. M.; Liu, Y.; Flood, A. H.; Bonvallet, P. A.; Tseng, H. R.; Stoddart, J. F.; Baller,
M.; Magonov, S. A nanomechanical device based on linear molecular motors Appl. Phys. Lett. 2004, 85, 5391–
5393 DOI: 10.1063/1.1826222 [Crossref], [CAS], [Google Scholar]

1342. Ooya, T.; Yui, N. Synthesis of theophylline–polyrotaxane conjugates and their drug release via
supramolecular dissociation J. Controlled Release 1999, 58, 251– 269 DOI: 10.1016/S0168-3659(98)00163-1
[Crossref], [PubMed], [CAS], [Google Scholar]

1343. Tooru, O.; Atsushi, Y.; Motoichi, K.; Yuko, S.; Atsushi, M.; Nobuhiko, Y. Effects of polyrotaxane structure on
polyion complexation with DNA Sci. Technol. Adv. Mater. 2004, 5, 363– 369 DOI: 10.1016/j.stam.2003.12.014
[Crossref], [Google Scholar]

1344. Ooya, T.; Choi, H. S.; Yamashita, A.; Yui, N.; Sugaya, Y.; Kano, A.; Maruyama, A.; Akita, H.; Ito, R.; Kogure, K.;
Harashima, H. Biocleavable Polyrotaxane–Plasmid DNA Polyplex for Enhanced Gene Delivery J. Am. Chem. Soc.
2006, 128, 3852– 3853 DOI: 10.1021/ja055868+ [ACS Full Text ], [CAS], [Google Scholar]

1345. Moon, C.; Kwon, Y. M.; Lee, W. K.; Park, Y. J.; Yang, V. C. In vitro assessment of a novel polyrotaxane-based
drug delivery system integrated with a cell-penetrating peptide J. Controlled Release 2007, 124, 43– 50
 DOI: 10.1016/j.jconrel.2007.08.029 [Crossref], [PubMed], [CAS], [Google Scholar]

1346. Zhang, X.; Zhu, X.; Ke, F.; Ye, L.; Chen, E.-q.; Zhang, A.-y.; Feng, Z.-g. Preparation and self-assembly of
amphiphilic triblock copolymers with polyrotaxane as a middle block and their application as carrier for the
controlled release of Amphotericin B Polymer 2009, 50, 4343– 4351 DOI: 10.1016/j.polymer.2009.07.006
[Crossref], [CAS], [Google Scholar]
1347. Harada, A. H. A.; Yamaguchi, H.; Takashima, Y. Polymeric Rotaxanes Chem. Rev. 2009, 109, 5974– 6023
 DOI: 10.1021/cr9000622 [ACS Full Text ], [CAS], [Google Scholar]   

1348. Li, J. J.; Zhao, F.; Li, J. Polyrotaxanes for applications in life science and biotechnology Appl. Microbiol.
Biotechnol. 2011, 90, 427– 443 DOI: 10.1007/s00253-010-3037-x [Crossref], [PubMed], [CAS], [Google Scholar]

1349. Hashidzume, A.; Yamaguchi, H.; Harada, A. Cyclodextrin-Based Molecular Machines Topics Curr. Chem.
2014, 354, 71– 110 DOI: 10.1007/128_2014_547 [Crossref], [PubMed], [CAS], [Google Scholar]

1350. Fernandes, A.; Viterisi, A.; Coutrot, F.; Potok, S.; Leigh, D. A.; Aucagne, V.; Papot, S. Rotaxane-Based
Propeptides: Protection and Enzymatic Release of a Bioactive Pentapeptide Angew. Chem., Int. Ed. 2009, 48,
6443– 6447 DOI: 10.1002/anie.200903215 [Crossref], [PubMed], [CAS], [Google Scholar]

1351. Fernandes, A.; Viterisi, A.; Aucagne, V.; Leigh, D. A.; Papot, S. Second generation speci c-enzyme-activated
rotaxane propeptides Chem. Commun. 2012, 48, 2083– 2085 DOI: 10.1039/c2cc17458h [Crossref], [PubMed],
[CAS], [Google Scholar]

1352. Barat, R.; Legigan, T.; Tranoy-Opalinski, I.; Renoux, B.; Peraudeau, E.; Clarhaut, J.; Poinot, P.; Fernandes, A.
E.; Aucagne, V.; Leigh, D. A.; Papot, S. A mechanically interlocked molecular system programmed for the delivery
of an anticancer drug Chem. Sci. 2015, 6, 2608– 2613 DOI: 10.1039/C5SC00648A [Crossref], [PubMed],
[CAS], [Google Scholar]

1353. Singleton, M. R.; Dillingham, M. S.; Wigley, D. B. Structure and mechanism of helicases and nucleic acid
translocases Annu. Rev. Biochem. 2007, 76, 23– 50 DOI: 10.1146/annurev.biochem.76.052305.115300
[Crossref], [PubMed], [CAS], [Google Scholar]

1354. Virag, L.; Szabo, C. The therapeutic potential of poly(ADP-ribose) polymerase inhibitors Pharmacol. Rev.
2002, 54, 375– 429 DOI: 10.1124/pr.54.3.375 [Crossref], [PubMed], [CAS], [Google Scholar]

1355. Gore, J.; Bryant, Z.; Stone, M. D.; Nollmann, M. N.; Cozzarelli, N. R.; Bustamante, C. Mechanochemical
analysis of DNA gyrase using rotor bead tracking Nature 2006, 439, 100– 104 DOI: 10.1038/nature04319
[Crossref], [PubMed], [CAS], [Google Scholar]

1356. Reece, R. J.; Maxwell, A. DNA Gyrase - Structure and Function Crit. Rev. Biochem. Mol. Biol. 1991, 26,
335– 375 DOI: 10.3109/10409239109114072 [Crossref], [PubMed], [CAS], [Google Scholar]
1357. Chen, S. H.; Chan, N. L.; Hsieh, T. S. New Mechanistic and Functional Insights into DNA Topoisomerases
Annu. Rev. Biochem. 2013, 82, 139– 170 DOI: 10.1146/annurev-biochem-061809-100002 [Crossref],  
[PubMed],
[CAS], [Google Scholar]

1358. Wang, J. C. Cellular roles of DNA topoisomerases: A molecular perspective Nat. Rev. Mol. Cell Biol. 2002,
3, 430– 440 DOI: 10.1038/nrm831 [Crossref], [PubMed], [CAS], [Google Scholar]

1359. von Ballmoos, C.; Cook, G. M.; Dimroth, P. Unique rotary ATP synthase and its biological diversity Annu.
Rev. Biophys. 2008, 37, 43– 64 DOI: 10.1146/annurev.biophys.37.032807.130018 [Crossref], [PubMed],
[CAS], [Google Scholar]

1360. Gadsby, D. C. Ion channels versus ion pumps: the principal difference, in principle Nat. Rev. Mol. Cell Biol.
2009, 10, 344– 352 DOI: 10.1038/nrm2668 [Crossref], [PubMed], [CAS], [Google Scholar]

1361. Yoneda, Y. How proteins are transported from cytoplasm to the nucleus J. Biochem. 1997, 121, 811–
817 DOI: 10.1093/oxfordjournals.jbchem.a021657 [Crossref], [PubMed], [CAS], [Google Scholar]

1362. Sweeney, H. L.; Houdusse, A. Structural and Functional Insights into the Myosin Motor Mechanism Annu.
Rev. Biophys. 2010, 39, 539– 557 DOI: 10.1146/annurev.biophys.050708.133751 [Crossref], [PubMed],
[CAS], [Google Scholar]

1363. Koonce, M. P.; Samso, M. Of rings and levers: the dynein motor comes of age Trends Cell Biol. 2004, 14,
612– 619 DOI: 10.1016/j.tcb.2004.09.013 [Crossref], [PubMed], [CAS], [Google Scholar]

1364. Vale, R. D. The molecular motor toolbox for intracellular transport Cell 2003, 112, 467– 480
 DOI: 10.1016/S0092-8674(03)00111-9 [Crossref], [PubMed], [CAS], [Google Scholar]

1365. Schliwa, M.; Woehlke, G. Molecular motors Nature 2003, 422, 759– 765 DOI: 10.1038/nature01601
[Crossref], [PubMed], [CAS], [Google Scholar]

1366. Ptacin, J. L.; Lee, S. F.; Garner, E. C.; Toro, E.; Eckart, M.; Comolli, L. R.; Moerner, W.; Shapiro, L. A spindle-
like apparatus guides bacterial chromosome segregation Nat. Cell Biol. 2010, 12, 791– 798
 DOI: 10.1038/ncb2083 [Crossref], [PubMed], [CAS], [Google Scholar]

1367. Saffarian, S.; Collier, I. E.; Marmer, B. L.; Elson, E. L.; Goldberg, G. Interstitial collagenase is a Brownian
ratchet driven by proteolysis of collagen Science 2004, 306, 108– 111 DOI: 10.1126/science.1099179 [Crossref],
[PubMed], [CAS], [Google Scholar]

  
1368. Xie, P. Molecular motors that digest their track to rectify Brownian motion: processive movement of
exonuclease enzymes J. Phys.: Condens. Matter 2009, 21, 375108 DOI: 10.1088/0953-8984/21/37/375108
[Crossref], [PubMed], [CAS], [Google Scholar]

1369. von Delius, M.; Leigh, D. A. Walking molecules Chem. Soc. Rev. 2011, 40, 3656– 3676
 DOI: 10.1039/c1cs15005g [Crossref], [PubMed], [CAS], [Google Scholar]

1370. Vale, R. D.; Reese, T. S.; Sheetz, M. P. Identi cation of a Novel Force-Generating Protein, Kinesin, Involved
in Microtubule-Based Motility Cell 1985, 42, 39– 50 DOI: 10.1016/S0092-8674(85)80099-4 [Crossref], [PubMed],
[CAS], [Google Scholar]

1371. Block, S. M. Kinesin motor mechanics: Binding, stepping, tracking, gating, and limping Biophys. J. 2007,
92, 2986– 2995 DOI: 10.1529/biophysj.106.100677 [Crossref], [PubMed], [CAS], [Google Scholar]

1372. Asbury, C. L.; Fehr, A. N.; Block, S. M. Kinesin moves by an asymmetric hand-over-hand mechanism
Science 2003, 302, 2130– 2134 DOI: 10.1126/science.1092985 [Crossref], [PubMed], [CAS], [Google Scholar]

1373. Yildiz, A.; Tomishige, M.; Vale, R. D.; Selvin, P. R. Kinesin walks hand-over-hand Science 2004, 303, 676–
678 DOI: 10.1126/science.1093753 [Crossref], [PubMed], [CAS], [Google Scholar]

1374. Howard, J.; Hudspeth, A. J.; Vale, R. D. Movement of Microtubules by Single Kinesin Molecules Nature
1989, 342, 154– 158 DOI: 10.1038/342154a0 [Crossref], [PubMed], [CAS], [Google Scholar]

1375. Block, S. M.; Goldstein, L. S. B.; Schnapp, B. J. Bead Movement by Single Kinesin Molecules Studied with
Optical Tweezers Nature 1990, 348, 348– 352 DOI: 10.1038/348348a0 [Crossref], [PubMed], [CAS], [Google
Scholar]

1376. Hackney, D. D. Highly Processive Microtubule-Stimulated ATP Hydrolysis by Dimeric Kinesin Head
Domains Nature 1995, 377, 448– 450 DOI: 10.1038/377448a0 [Crossref], [PubMed], [CAS], [Google Scholar]

1377. Vale, R. D.; Funatsu, T.; Pierce, D. W.; Romberg, L.; Harada, Y.; Yanagida, T. Direct observation of single
kinesin molecules moving along microtubules Nature 1996, 380, 451– 453 DOI: 10.1038/380451a0 [Crossref],
[PubMed], [CAS], [Google Scholar]
1378. Hirokawa, N.; Nitta, R.; Okada, Y. The mechanisms of kinesin motor motility: lessons from the monomeric
motor KIF1A Nat. Rev. Mol. Cell Biol. 2009, 10, 877– 884 DOI: 10.1038/nrm2807 [Crossref], [PubMed],
  
[CAS], [Google Scholar]

1379. Yin, P.; Yan, H.; Daniell, X. G.; Turber eld, A. J.; Reif, J. H. A unidirectional DNA walker that moves
autonomously along a track Angew. Chem., Int. Ed. 2004, 43, 4906– 4911 DOI: 10.1002/anie.200460522
[Crossref], [PubMed], [CAS], [Google Scholar]

1380. Tian, Y.; He, Y.; Chen, Y.; Yin, P.; Mao, C. D. Molecular devices - A DNAzyme that walks processively and
autonomously along a one-dimensional track Angew. Chem., Int. Ed. 2005, 44, 4355– 4358
 DOI: 10.1002/anie.200500703 [Crossref], [PubMed], [CAS], [Google Scholar]

1381. Muscat, R. A.; Bath, J.; Turber eld, A. J. Small Molecule Signals that Direct the Route of a Molecular
Cargo Small 2012, 8, 3593– 3597 DOI: 10.1002/smll.201201055 [Crossref], [PubMed], [CAS], [Google Scholar]

1382. You, M. X.; Chen, Y.; Zhang, X. B.; Liu, H. P.; Wang, R. W.; Wang, K. L.; Williams, K. R.; Tan, W. H. An
Autonomous and Controllable Light-Driven DNA Walking Device Angew. Chem., Int. Ed. 2012, 51, 2457– 2460
 DOI: 10.1002/anie.201107733 [Crossref], [PubMed], [CAS], [Google Scholar]

1383. You, M. X.; Huang, F. J.; Chen, Z.; Wang, R. W.; Tan, W. H. Building a Nanostructure with Reversible
Motions Using Photonic Energy ACS Nano 2012, 6, 7935– 7941 DOI: 10.1021/nn302388e [ACS Full Text ],
[CAS], [Google Scholar]

1384. Perl, A.; Gomez-Casado, A.; Thompson, D.; Dam, H. H.; Jonkheijm, P.; Reinhoudt, D. N.; Huskens, J.
Gradient-driven motion of multivalent ligand molecules along a surface functionalized with multiple receptors
Nat. Chem. 2011, 3, 317– 322 DOI: 10.1038/nchem.1005 [Crossref], [PubMed], [CAS], [Google Scholar]

1385. Weimann, D. P.; Winkler, H. D. F.; Falenski, J. A.; Koksch, B.; Schalley, C. A. Highly dynamic motion of crown
ethers along oligolysine peptide chains Nat. Chem. 2009, 1, 668– 668 DOI: 10.1038/nchem.431 [Crossref],
[CAS], [Google Scholar]

1386. Strawser, D.; Karton, A.; Zenkina, E. V.; Iron, M. A.; Shimon, L. J. W.; Martin, J. M. L.; van der Boom, M. E.
Platinum stilbazoles: Ring-walking coupled with aryl-halide bond activation J. Am. Chem. Soc. 2005, 127, 9322–
9323 DOI: 10.1021/ja050613h [ACS Full Text ], [CAS], [Google Scholar]

1387. Tkachov, R.; Senkovskyy, V.; Komber, H.; Sommer, J. U.; Kiriy, A. Random Catalyst Walking along
Polymerized Poly(3-hexylthlophene) Chains in Kumada Catalyst-Transfer Polycondensation J. Am. Chem. Soc.
2010, 132, 7803– 7810 DOI: 10.1021/ja102210r [ACS Full Text ], [CAS], [Google Scholar]

  
1388. Claisen, L.; Tietze, E.; Über den Mechanismus der Umlagerung der Phenol-allyläther Ber. Dtsch. Chem.
Ges. B 1925, 58, 275– 281 DOI: 10.1002/cber.19250580207 [Crossref], [CAS], [Google Scholar]

1389. Claisen, L. The rearrangement of phenol-allyl-ather in C-allyl-phenole Ber. Dtsch. Chem. Ges. 1912, 45,
3157– 3166 DOI: 10.1002/cber.19120450348 [Crossref], [CAS], [Google Scholar]

1390. Cope, A. C.; Hardy, E. M. The introduction of substituted vinyl groups. VI. The regeneration of substituted
vinyl malonic esters from their sodium enolates J. Am. Chem. Soc. 1940, 62, 3319– 3323
 DOI: 10.1021/ja01869a013 [ACS Full Text ], [CAS], [Google Scholar]

1391. Cope, A. C.; Hardy, E. M. The introduction of substituted vinyl groups V A rearrangement involving the
migration of an allyl group in a three-carbon system J. Am. Chem. Soc. 1940, 62, 441– 444
 DOI: 10.1021/ja01859a055 [ACS Full Text ], [CAS], [Google Scholar]

1392. Cope, A. C.; Hartung, W. H.; Hancock, E. M.; Crossley, F. S. Substituted vinyl barbituric acids IV Derivatives
containing a primary 1-alkenyl group J. Am. Chem. Soc. 1940, 62, 1199– 1201 DOI: 10.1021/ja01862a060 [ACS
Full Text ], [CAS], [Google Scholar]

1393. Cope, A. C.; Hartung, W. H.; Hancock, E. M.; Crossley, F. S. The introduction of substituted vinyl groups IV
(Primary 1-alkenyl) alkyl malonic esters J. Am. Chem. Soc. 1940, 62, 314– 316 DOI: 10.1021/ja01859a020 [ACS
Full Text ], [CAS], [Google Scholar]

1394. Mitra, S.; Lawton, R. G. Reagents for the Cross-Linking of Proteins by Equilibrium Transfer Alkylation J.
Am. Chem. Soc. 1979, 101, 3097– 3110 DOI: 10.1021/ja00505a043 [ACS Full Text ], [CAS], [Google Scholar]

1395. Kovaricek, P.; Lehn, J.-M. Merging Constitutional and Motional Covalent Dynamics in Reversible Imine
Formation and Exchange Processes J. Am. Chem. Soc. 2012, 134, 9446– 9455 DOI: 10.1021/ja302793c [ACS
Full Text ], [CAS], [Google Scholar]

1396. Kovaricek, P.; Lehn, J.-M. Directional Dynamic Covalent Motion of a Carbonyl Walker on a Polyamine
Track Chem. - Eur. J. 2015, 21, 9380– 9384 DOI: 10.1002/chem.201500987 [Crossref], [PubMed], [CAS], [Google
Scholar]
1397. Campana, A. G.; Carlone, A.; Chen, K.; Dryden, D. T. F.; Leigh, D. A.; Lewandowska, U.; Mullen, K. M. A Small
2012,
Int. Ed.
Molecule that Walks Non-Directionally Along a Track Without External Intervention Angew. Chem., 
51, 5480– 5483 DOI: 10.1002/anie.201200822 [Crossref], [PubMed], [CAS], [Google Scholar]

1398. Campana, A. G.; Leigh, D. A.; Lewandowska, U. One-Dimensional Random Walk of a Synthetic Small
Molecule Toward a Thermodynamic Sink J. Am. Chem. Soc. 2013, 135, 8639– 8645 DOI: 10.1021/ja402382n
[ACS Full Text ], [CAS], [Google Scholar]

1399. von Delius, M.; Geertsema, E. M.; Leigh, D. A. A synthetic small molecule that can walk down a track Nat.
Chem. 2010, 2, 96– 101 DOI: 10.1038/nchem.481 [Crossref], [PubMed], [CAS], [Google Scholar]

1400. von Delius, M.; Geertsema, E. M.; Leigh, D. A.; Tang, D. T. D. Design, Synthesis, and Operation of Small
Molecules That Walk along Tracks J. Am. Chem. Soc. 2010, 132, 16134– 16145 DOI: 10.1021/ja106486b [ACS
Full Text ], [CAS], [Google Scholar]

1401. Pulcu, G. S.; Mikhailova, E.; Choi, L. S.; Bayley, H. Continuous observation of the stochastic motion of an
individual small-molecule walker Nat. Nanotechnol. 2014, 10, 76– 83 DOI: 10.1038/nnano.2014.264 [Crossref],
[PubMed], [Google Scholar]

1402. Haq, S.; Wit, B.; Sang, H.; Floris, A.; Wang, Y.; Wang, J.; Pérez-García, L.; Kantorovitch, L.; Amabilino, D. B.;
Raval, R. A Small Molecule Walks Along a Surface Between Porphyrin Fences That Are Assembled In Situ Angew.
Chem., Int. Ed. 2015, 54, 7101– 7105 DOI: 10.1002/anie.201502153 [Crossref], [PubMed], [CAS], [Google Scholar]

1403. Beves, J. E.; Blanco, V.; Blight, B. A.; Carrillo, R.; D’Souza, D. M.; Howgego, D.; Leigh, D. A.; Slawin, A. M. Z.;
Symes, M. D. Toward Metal Complexes That Can Directionally Walk Along Tracks: Controlled Stepping of a
Molecular Biped with a Palladium(II) Foot J. Am. Chem. Soc. 2014, 136, 2094– 2100 DOI: 10.1021/ja4123973
[ACS Full Text ], [CAS], [Google Scholar]

1404. Barrell, M. J.; Campana, A. G.; von Delius, M.; Geertsema, E. M.; Leigh, D. A. Light-Driven Transport of a
Molecular Walker in Either Direction along a Molecular Track Angew. Chem., Int. Ed. 2011, 50, 285– 290
 DOI: 10.1002/anie.201004779 [Crossref], [PubMed], [CAS], [Google Scholar]

1405. Luning, U. Switchable catalysis Angew. Chem., Int. Ed. 2012, 51, 8163– 8165
 DOI: 10.1002/anie.201204567 [Crossref], [PubMed], [CAS], [Google Scholar]

1406. Kumagai, N.; Shibasaki, M. Catalytic chemical transformations with conformationally dynamic catalytic
systems Catal. Sci. Technol. 2013, 3, 41– 57 DOI: 10.1039/C2CY20257C [Crossref], [CAS], [Google Scholar]
 
1407. Neilson, B. M.; Bielawski, C. W. Illuminating Photoswitchable Catalysis ACS Catal. 2013, 3, 1874– 1885

 DOI: 10.1021/cs4003673 [ACS Full Text ], [CAS], [Google Scholar]

1408. Leigh, D. A.; Marcos, V.; Wilson, M. R. Rotaxane Catalysts ACS Catal. 2014, 4, 4490– 4497
 DOI: 10.1021/cs5013415 [ACS Full Text ], [CAS], [Google Scholar]

1409. Blanco, V.; Leigh, D. A.; Marcos, V. Arti cial switchable catalysts Chem. Soc. Rev. 2015, 44, 5341– 5370
 DOI: 10.1039/C5CS00096C [Crossref], [PubMed], [CAS], [Google Scholar]

1410. Wiester, M. J.; Ulmann, P. A.; Mirkin, C. A. Enzyme Mimics Based Upon Supramolecular Coordination
Chemistry Angew. Chem., Int. Ed. 2011, 50, 114– 137 DOI: 10.1002/anie.201000380 [Crossref], [PubMed],
[CAS], [Google Scholar]

1411. Yoon, H. J.; Mirkin, C. A. PCR-like cascade reactions in the context of an allosteric enzyme mimic J. Am.
Chem. Soc. 2008, 130, 11590– 11591 DOI: 10.1021/ja804076q [ACS Full Text ], [CAS], [Google Scholar]

1412. Yoon, H. J.; Heo, J.; Mirkin, C. A. Allosteric regulation of phosphate diester transesteri cation based upon
a dinuclear zinc catalyst assembled via the weak-link approach J. Am. Chem. Soc. 2007, 129, 14182– 14183
 DOI: 10.1021/ja077467v [ACS Full Text ], [CAS], [Google Scholar]

1413. Vlatkovic, M.; Bernardi, L.; Otten, E.; Feringa, B. L. Dual stereocontrol over the Henry reaction using a light-
and heat-triggered organocatalyst Chem. Commun. 2014, 50, 7773– 7775 DOI: 10.1039/c4cc00794h [Crossref],
[PubMed], [CAS], [Google Scholar]

1414. Stoll, R. S.; Hecht, S. Arti cial Light-Gated Catalyst Systems Angew. Chem., Int. Ed. 2010, 49, 5054– 5075
 DOI: 10.1002/anie.201000146 [Crossref], [PubMed], [CAS], [Google Scholar]

1415. Imahori, T.; Kurihara, S. Stimuli-responsive Cooperative Catalysts Based on Dynamic Conformational
Changes toward Spatiotemporal Control of Chemical Reactions Chem. Lett. 2014, 43, 1524– 1531
 DOI: 10.1246/cl.140680 [Crossref], [CAS], [Google Scholar]

1416. Frank Wuerthner; Julius Rebek, J. Light-Switchable Catalysis in Synthetic Receptors Angew. Chem., Int.
Ed. Engl. 1995, 34, 446– 448 DOI: 10.1002/anie.199504461 [Crossref], [Google Scholar]
1417. Berryman, O. B.; Sather, A. C.; Lledo, A.; Rebek, J., Jr. Switchable catalysis with a light-responsive cavitand
  
Angew. Chem., Int. Ed. 2011, 50, 9400– 9403 DOI: 10.1002/anie.201105374 [Crossref], [PubMed], [CAS], [Google
Scholar]

1418. Peters, M. V.; Stoll, R. S.; Kuhn, A.; Hecht, S. Photoswitching of basicity Angew. Chem., Int. Ed. 2008, 47,
5968– 5972 DOI: 10.1002/anie.200802050 [Crossref], [PubMed], [CAS], [Google Scholar]

1419. Osorio-Planes, L.; Rodriguez-Escrich, C.; Pericas, M. A. Photoswitchable thioureas for the external
manipulation of catalytic activity Org. Lett. 2014, 16, 1704– 1707 DOI: 10.1021/ol500381c [ACS Full Text ],
[CAS], [Google Scholar]

1420. Viehmann, P.; Hecht, S. Design and synthesis of a photoswitchable guanidine catalyst Beilstein J. Org.
Chem. 2012, 8, 1825– 1830 DOI: 10.3762/bjoc.8.209 [Crossref], [PubMed], [CAS], [Google Scholar]

1421. Neilson, B. M.; Bielawski, C. W. Photoswitchable Organocatalysis: Using Light To Modulate the Catalytic
Activities of N-Heterocyclic Carbenes J. Am. Chem. Soc. 2012, 134, 12693– 12699 DOI: 10.1021/ja304067k
[ACS Full Text ], [CAS], [Google Scholar]

1422. Neilson, B. M.; Bielawski, C. W. Photoswitchable Metal-Mediated Catalysis: Remotely Tuned Alkene and
Alkyne Hydroborations Organometallics 2013, 32, 3121– 3128 DOI: 10.1021/om400348h [ACS Full Text ],
[CAS], [Google Scholar]

1423. Wilson, D.; Branda, N. R. Turning ″on″ and ″off″ a pyridoxal 5′-phosphate mimic using light Angew. Chem.,
Int. Ed. 2012, 51, 5431– 5434 DOI: 10.1002/anie.201201447 [Crossref], [PubMed], [CAS], [Google Scholar]

1424. Wang, J.; Feringa, B. L. Dynamic control of chiral space in a catalytic asymmetric reaction using a
molecular motor Science 2011, 331, 1429– 1432 DOI: 10.1126/science.1199844 [Crossref], [PubMed],
[CAS], [Google Scholar]

1425. Zhao, D.; Neubauer, T. M.; Feringa, B. L. Dynamic control of chirality in phosphine ligands for
enantioselective catalysis Nat. Commun. 2015, 6, 6652 DOI: 10.1038/ncomms7652 [Crossref], [PubMed],
[CAS], [Google Scholar]

1426. Li, Y.; Feng, Y.; He, Y. M.; Chen, F.; Pan, J.; Fan, Q. H. Supramolecular chiral phosphorous ligands based on
a [2]pseudorotaxane complex for asymmetric hydrogenation Tetrahedron Lett. 2008, 49, 2878– 2881
 DOI: 10.1016/j.tetlet.2008.03.039 [Crossref], [CAS], [Google Scholar]
1427. Tachibana, Y.; Kihara, N.; Nakazono, K.; Takata, T. Thiazolium-Tethering Rotaxane-Catalyzed Asymmetric
 
Benzoin Condensation: Unique Asymmetric Field Constructed by the Cooperation of Rotaxane Components 
Phosphorus, Sulfur Silicon Relat. Elem. 2010, 185, 1182– 1205 DOI: 10.1080/10426501003773589 [Crossref],
[CAS], [Google Scholar]

1428. Suzaki, Y.; Shimada, K.; Chihara, E.; Saito, T.; Tsuchido, Y.; Osakada, K. [3]Rotaxane-Based Dinuclear
Palladium Catalysts for Ring-closure Mizoroki-Heck Reaction Org. Lett. 2011, 13, 3774– 3777
 DOI: 10.1021/ol201357b [ACS Full Text ], [CAS], [Google Scholar]

1429. Caputo, C. B.; Zhu, K. L.; Vukotic, V. N.; Loeb, S. J.; Stephan, D. W. Heterolytic Activation of H2 Using a
Mechanically Interlocked Molecule as a Frustrated Lewis Base Angew. Chem., Int. Ed. 2013, 52, 960– 963
 DOI: 10.1002/anie.201207783 [Crossref], [PubMed], [CAS], [Google Scholar]

1430. Miyagawa, N.; Watanabe, M.; Matsuyama, T.; Koyama, Y.; Moriuchi, T.; Hirao, T.; Furusho, Y.; Takata, T.
Successive catalytic reactions speci c to Pd-based rotaxane complexes as a result of wheel translation along
the axle Chem. Commun. 2010, 46, 1920– 1922 DOI: 10.1039/b917053g [Crossref], [PubMed], [CAS], [Google
Scholar]

1431. Blanco, V.; Carlone, A.; Hanni, K. D.; Leigh, D. A.; Lewandowski, B. A rotaxane-based switchable
organocatalyst Angew. Chem., Int. Ed. 2012, 51, 5166– 5169 DOI: 10.1002/anie.201201364 [Crossref], [PubMed],
[CAS], [Google Scholar]

1432. Blanco, V.; Leigh, D. A.; Lewandowska, U.; Lewandowsld, B.; Marcos, V. Exploring the Activation Modes of
a Rotaxane-Based Switchable Organocatalyst J. Am. Chem. Soc. 2014, 136, 15775– 15780
 DOI: 10.1021/ja509236u [ACS Full Text ], [CAS], [Google Scholar]

1433. Blanco, V.; Leigh, D. A.; Marcos, V.; Morales-Serna, J. A.; Nussbaumer, A. L. A Switchable [2]Rotaxane
Asymmetric Organocatalyst That Utilizes an Acyclic Chiral Secondary Amine J. Am. Chem. Soc. 2014, 136,
4905– 4908 DOI: 10.1021/ja501561c [ACS Full Text ], [CAS], [Google Scholar]

1434. Beswick, J.; Blanco, V.; De Bo, G.; Leigh, D. A.; Lewandowska, U.; Lewandowski, B.; Mishiro, K. Selecting
reactions and reactants using a switchable rotaxane organocatalyst with two different active sites Chem. Sci.
2015, 6, 140– 143 DOI: 10.1039/C4SC03279A [Crossref], [PubMed], [CAS], [Google Scholar]

1435. Yoon, H. J.; Kuwabara, J.; Kim, J. H.; Mirkin, C. A. Allosteric supramolecular triple-layer catalysts Science
2010, 330, 66– 69 DOI: 10.1126/science.1193928 [Crossref], [PubMed], [CAS], [Google Scholar]
1436. Wiester, M. J.; Braunschweig, A. B.; Yoo, H.; Mirkin, C. A. Solvent and Temperature Induced Switching
Between Structural Isomers of Rh-I Phosphinoalkyl Thioether (PS) Complexes Inorg. Chem. 2010,  
49, 7188– 
7196 DOI: 10.1021/ic101021t [ACS Full Text ], [CAS], [Google Scholar]

1437. McGuirk, C. M.; Mendez-Arroyo, J.; Lifschitz, A. M.; Mirkin, C. A. Allosteric Regulation of Supramolecular
Oligomerization and Catalytic Activity via Coordination-Based Control of Competitive Hydrogen-Bonding Events
J. Am. Chem. Soc. 2014, 136, 16594– 16601 DOI: 10.1021/ja508804n [ACS Full Text ], [CAS], [Google Scholar]

1438. Lifschitz, A. M.; Young, R. M.; Mendez-Arroyo, J.; Stern, C. L.; McGuirk, C. M.; Wasielewski, M. R.; Mirkin, C.
A. An allosteric photoredox catalyst inspired by photosynthetic machinery Nat. Commun. 2015, 6, 6541
 DOI: 10.1038/ncomms7541 [Crossref], [PubMed], [CAS], [Google Scholar]

1439. Schmittel, M.; De, S.; Pramanik, S. Reversible ON/OFF nanoswitch for organocatalysis: mimicking the
locking and unlocking operation of CaMKII Angew. Chem., Int. Ed. 2012, 51, 3832– 3836
 DOI: 10.1002/anie.201108089 [Crossref], [PubMed], [CAS], [Google Scholar]

1440. Schmittel, M.; Pramanik, S.; De, S. A reversible nanoswitch as an ON-OFF photocatalyst Chem. Commun.
2012, 48, 11730– 11732 DOI: 10.1039/c2cc36408e [Crossref], [PubMed], [CAS], [Google Scholar]

1441. Samanta, S. K.; Schmittel, M. Four-component supramolecular nanorotors J. Am. Chem. Soc. 2013, 135,
18794– 18797 DOI: 10.1021/ja411011a [ACS Full Text ], [CAS], [Google Scholar]

1442. Zahn, S. Electron-Induced Inversion of Helical Chirality in Copper Complexes of N,N-Dialkylmethionines


Science 2000, 288, 1404– 1407 DOI: 10.1126/science.288.5470.1404 [Crossref], [PubMed], [CAS], [Google
Scholar]

1443. Canary, J. W.; Mortezaei, S.; Liang, J. Redox-recon gurable tripodal coordination complexes:
stereodynamic molecular switches Chem. Commun. 2010, 46, 5850– 5860 DOI: 10.1039/c0cc00469c
[Crossref], [PubMed], [CAS], [Google Scholar]

1444. Berova, N.; Nakanishi, K. o.; Woody, R. Circular Dichroism: Principles and Applications, 2nd ed.; Wiley-
VCH: New York, 2000. [Google Scholar]

1445. Mortezaei, S.; Catarineu, N. R.; Canary, J. W. A redox-recon gurable, ambidextrous asymmetric catalyst J.
Am. Chem. Soc. 2012, 134, 8054– 8057 DOI: 10.1021/ja302283s [ACS Full Text ], [CAS], [Google Scholar]
1446. Gartner, Z. J.; Kanan, M. W.; Liu, D. R. Multistep Small-Molecule Synthesis Programmed by DNA
Templates J. Am. Chem. Soc. 2002, 124, 10304– 10306 DOI: 10.1021/ja027307d [ACS Full Text ],
  
[CAS], [Google Scholar]

1447. He, Y.; Liu, D. R. Autonomous multistep organic synthesis in a single isothermal solution mediated by a
DNA walker Nat. Nanotechnol. 2010, 5, 778– 782 DOI: 10.1038/nnano.2010.190 [Crossref], [PubMed],
[CAS], [Google Scholar]

1448. He, Y.; Liu, D. R. A Sequential Strand-Displacement Strategy Enables E cient Six-Step DNA-Templated
Synthesis J. Am. Chem. Soc. 2011, 133, 9972– 9975 DOI: 10.1021/ja201361t [ACS Full Text ], [CAS], [Google
Scholar]

1449. McKee, M. L.; Milnes, P. J.; Bath, J.; Stulz, E.; O’Reilly, R. K.; Turber eld, A. J. Programmable One-Pot
Multistep Organic Synthesis Using DNA Junctions J. Am. Chem. Soc. 2012, 134, 1446– 1449
 DOI: 10.1021/ja2101196 [ACS Full Text ], [CAS], [Google Scholar]

1450. McKee, M. L.; Milnes, P. J.; Bath, J.; Stulz, E.; Turber eld, A. J.; O’Reilly, R. K. Multistep DNA-Templated
Reactions for the Synthesis of Functional Sequence Controlled Oligomers Angew. Chem., Int. Ed. 2010, 49,
7948– 7951 DOI: 10.1002/anie.201002721 [Crossref], [PubMed], [CAS], [Google Scholar]

1451. Gartner, Z. J.; Kanan, M. W.; Liu, D. R. Expanding the Reaction Scope of DNA-Templated Synthesis Angew.
Chem., Int. Ed. 2002, 41, 1796– 1800 DOI: 10.1002/1521-3773(20020517)41:10<1796::AID-ANIE1796>3.0.CO;2-Z
[Crossref], [PubMed], [CAS], [Google Scholar]

1452. Gartner, Z. J.; Liu, D. R. The Generality of DNA-Templated Synthesis as a Basis for Evolving Non-Natural
Small Molecules J. Am. Chem. Soc. 2001, 123, 6961– 6963 DOI: 10.1021/ja015873n [ACS Full Text ],
[CAS], [Google Scholar]

1453. Li, X.; Gartner, Z. J.; Tse, B. N.; Liu, D. R. Translation of DNA into Synthetic N-Acyloxazolidines J. Am.
Chem. Soc. 2004, 126, 5090– 5092 DOI: 10.1021/ja049666+ [ACS Full Text ], [CAS], [Google Scholar]

1454. Gartner, Z. J.; Grubina, R.; Calderone, C. T.; Liu, D. R. Two Enabling Architectures for DNA-Templated
Organic Synthesis Angew. Chem., Int. Ed. 2003, 42, 1370– 1375 DOI: 10.1002/anie.200390351 [Crossref],
[PubMed], [CAS], [Google Scholar]

1455. Gu, H.; Chao, J.; Xiao, S. J.; Seeman, N. C. A proximity-based programmable DNA nanoscale assembly
line Nature 2010, 465, 202– 205 DOI: 10.1038/nature09026 [Crossref], [PubMed], [CAS], [Google Scholar]
by a  
1456. Thordarson, P.; Bijsterveld, E. J. A.; Rowan, A. E.; Nolte, R. J. M. Epoxidation of polybutadiene
topologically linked catalyst Nature 2003, 424, 915– 918 DOI: 10.1038/nature01925 [Crossref], [PubMed],
[CAS], [Google Scholar]

1457. Elemans, J. A. A. W.; Bijsterveld, E. J. A.; Rowan, A. E.; Nolte, R. J. M. A host-guest epoxidation catalyst
with enhanced activity and stability Chem. Commun. 2000, 2443– 2444 DOI: 10.1039/b008567g [Crossref],
[CAS], [Google Scholar]

1458. Monnereau, C.; Ramos, P. H.; Deutman, A. B. C.; Elemans, J. A. A. W.; Nolte, R. J. M.; Rowan, A. E.
Porphyrin Macrocyclic Catalysts for the Processive Oxidation of Polymer Substrates J. Am. Chem. Soc. 2010,
132, 1529– 1531 DOI: 10.1021/ja908524x [ACS Full Text ], [CAS], [Google Scholar]

1459. Deutman, A. B. C.; Monnereau, C.; Elemans, J. A. A. W.; Ercolani, G.; Nolte, R. J. M.; Rowan, A. E.
Mechanism of Threading a Polymer Through a Macrocyclic Ring Science 2008, 322, 1668– 1671
 DOI: 10.1126/science.1164647 [Crossref], [PubMed], [CAS], [Google Scholar]

1460. Takashima, Y.; Osaki, M.; Ishimaru, Y.; Yamaguchi, H.; Harada, A. Arti cial Molecular Clamp: A Novel
Device for Synthetic Polymerases Angew. Chem., Int. Ed. 2011, 50, 7524– 7528 DOI: 10.1002/anie.201102834
[Crossref], [PubMed], [CAS], [Google Scholar]

1461. Dawson, P.; Muir, T.; Clark-Lewis, I.; Kent, S. Synthesis of proteins by native chemical ligation Science
1994, 266, 776– 779 DOI: 10.1126/science.7973629 [Crossref], [PubMed], [CAS], [Google Scholar]

1462. Hughs, M.; Jimenez, M.; Khan, S.; Garcia-Garibay, M. A. Synthesis, rotational dynamics, and photophysical
characterization of a crystalline linearly conjugated phenyleneethynylene molecular dirotor J. Org. Chem. 2013,
78, 5293– 5302 DOI: 10.1021/jo4004053 [ACS Full Text ], [CAS], [Google Scholar]

1463. Garcia-Garibay, M. A. Crystalline molecular machines: encoding supramolecular dynamics into molecular
structure Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 10771– 10776 DOI: 10.1073/pnas.0502816102 [Crossref],
[PubMed], [CAS], [Google Scholar]

1464. Rodriguez-Molina, B.; Pozos, A.; Cruz, R.; Romero, M.; Flores, B.; Farfan, N.; Santillan, R.; Garcia-Garibay, M.
A. Synthesis and solid state characterization of molecular rotors with steroidal stators: ethisterone and
norethisterone Org. Biomol. Chem. 2010, 8, 2993– 3000 DOI: 10.1039/c003778h [Crossref], [PubMed],
[CAS], [Google Scholar]
1465. Vogelsberg, C. S.; Garcia-Garibay, M. A. Crystalline molecular machines: function, phase order,
dimensionality, and composition Chem. Soc. Rev. 2012, 41, 1892– 1910 DOI: 10.1039/C1CS15197E  
[Crossref],
[PubMed], [CAS], [Google Scholar]

1466. Rodriguez-Molina, B.; Perez-Estrada, S.; Garcia-Garibay, M. A. Amphidynamic crystals of a steroidal


bicyclo[2.2.2]octane rotor: a high symmetry group that rotates faster than smaller methyl and methoxy groups J.
Am. Chem. Soc. 2013, 135, 10388– 10395 DOI: 10.1021/ja4024463 [ACS Full Text ], [CAS], [Google Scholar]

1467. Staehle, I. O.; Rodríguez-Molina, B.; Khan, S. I.; Garcia-Garibay, M. A. Engineered Photochromism in
Crystalline Salicylidene Anilines by Facilitating Rotation to Reach the Colored trans-Keto Form Cryst. Growth Des.
2014, 14, 3667– 3673 DOI: 10.1021/cg500762a [ACS Full Text ], [CAS], [Google Scholar]

1468. Jiang, X.; Rodriguez-Molina, B.; Nazarian, N.; Garcia-Garibay, M. A. Rotation of a bulky triptycene in the
solid state: toward engineered nanoscale arti cial molecular machines J. Am. Chem. Soc. 2014, 136, 8871–
8874 DOI: 10.1021/ja503467e [ACS Full Text ], [CAS], [Google Scholar]

1469. Commins, P.; Garcia-Garibay, M. A. Photochromic molecular gyroscope with solid state rotational states
determined by an azobenzene bridge J. Org. Chem. 2014, 79, 1611– 1619 DOI: 10.1021/jo402516n [ACS Full
Text ], [CAS], [Google Scholar]

1470. Karlen, S. D.; Ortiz, R.; Chapman, O. L.; Garcia-Garibay, M. A. Effects of Rotational Symmetry Order on the
Solid State Dynamics Phenylene and Diamantane Rotators J. Am. Chem. Soc. 2005, 127, 6554– 6555
 DOI: 10.1021/ja042512+ [ACS Full Text ], [CAS], [Google Scholar]

1471. Comotti, A.; Bracco, S.; Valsesia, P.; Beretta, M.; Sozzani, P. Fast molecular rotor dynamics modulated by
guest inclusion in a highly organized nanoporous organosilica Angew. Chem., Int. Ed. 2010, 49, 1760– 1764
 DOI: 10.1002/anie.200906255 [Crossref], [PubMed], [CAS], [Google Scholar]

1472. Vogelsberg, C. S.; Bracco, S.; Beretta, M.; Comotti, A.; Sozzani, P.; Garcia-Garibay, M. A. Dynamics of
molecular rotors con ned in two dimensions: transition from a 2D rotational glass to a 2D rotational uid in a
periodic mesoporous organosilica J. Phys. Chem. B 2012, 116, 1623– 1632 DOI: 10.1021/jp2119263 [ACS Full
Text ], [CAS], [Google Scholar]

1473. Comotti, A.; Bracco, S.; Yamamoto, A.; Beretta, M.; Hirukawa, T.; Tohnai, N.; Miyata, M.; Sozzani, P.
Engineering switchable rotors in molecular crystals with open porosity J. Am. Chem. Soc. 2014, 136, 618– 621
 DOI: 10.1021/ja411233p [ACS Full Text ], [CAS], [Google Scholar]
1474. Comotti, A.; Bracco, S.; Ben, T.; Qiu, S.; Sozzani, P. Molecular rotors in porous organic frameworks Angew.
 
Chem., Int. Ed. 2014, 53, 1043– 1047 DOI: 10.1002/anie.201309362 [Crossref], [PubMed], [CAS], [Google 
Scholar]

1475. Vukotic, V. N.; Kelong Zhu, K. J. H.; Schurko, R. W.; Loeb, S. J. Metal–organic frameworks with dynamic
interlocked components Nat. Chem. 2012, 4, 456– 460 DOI: 10.1038/nchem.1354 [Crossref], [PubMed],
[CAS], [Google Scholar]

1476. Zhu, K.; O’Keefe, C. A.; Vukotic, V. N.; Schurko, R. W.; Loeb, S. J. A molecular shuttle that operates inside a
metal-organic framework Nat. Chem. 2015, 7, 514– 519 DOI: 10.1038/nchem.2258 [Crossref], [PubMed],
[CAS], [Google Scholar]

1477. Ahuja, R. C.; Caruso, P. L.; Mobius, D.; Wildburg, G.; Ringsdorf, H.; Philp, D.; Preece, J. A.; Stoddart, J. F.
Molecular-Organization Via Ionic Interactions at Interfaces 0.1. Monolayers and Lb Films of Cyclic
Bisbipyridinium Tetracations and Dimyristoylphosphatidic Acid Langmuir 1993, 9, 1534– 1544
 DOI: 10.1021/la00030a019 [ACS Full Text ], [CAS], [Google Scholar]

1478. Collier, C. P.; Mattersteig, G.; Wong, E. W.; Luo, Y.; Beverly, K.; Sampaio, J.; Raymo, F. M.; Stoddart, J. F.;
Heath, J. R. A [2]catenane-based solid state electronically recon gurable switch Science 2000, 289, 1172– 1175
 DOI: 10.1126/science.289.5482.1172 [Crossref], [PubMed], [CAS], [Google Scholar]

1479. Pease, A. R.; Jeppesen, J. O.; Stoddart, J. F.; Luo, Y.; Collier, C. P.; Heath, J. R. Switching devices based on
interlocked molecules Acc. Chem. Res. 2001, 34, 433– 444 DOI: 10.1021/ar000178q [ACS Full Text ],
[CAS], [Google Scholar]

1480. Heath, J. R.; Ratner, M. A. Molecular electronics Phys. Today 2003, 56, 43– 49 DOI: 10.1063/1.1583533
[Crossref], [CAS], [Google Scholar]

1481. Heath, J. R.; Stoddart, J. F.; Williams, R. S. More on molecular electronics Science 2004, 303, 1136– 1137
 DOI: 10.1126/science.303.5661.1136c [Crossref], [PubMed], [CAS], [Google Scholar]

1482. Flood, A. H.; Ramirez, R.; Deng, W.; Muller, R.; Goddard, W. A., III; Stoddart, J. F. Meccano on the nanoscale
- A blueprint for making some of the world’s tiniest machines Aust. J. Chem. 2004, 57, 301– 322
 DOI: 10.1071/CH03307 [Crossref], [CAS], [Google Scholar]

1483. Mendes, P.; Flood, A. H.; Stoddart, J. F. Nanoelectronic devices from self-organized molecular switches
Appl. Phys. A: Mater. Sci. Process. 2005, 80, 1197– 1209 DOI: 10.1007/s00339-004-3172-2 [Crossref],
[CAS], [Google Scholar]
  
1484. Beckman, R.; Beverly, K.; Boukai, A.; Bunimovich, Y.; Choi, J. W.; DeIonno, E.; Green, J.; Johnston-Halperin,
E.; Luo, Y.; Sheriff, B.; Stoddart, J. F.; Heath, J. R. Spiers Memorial Lecture Molecular mechanics and molecular
electronics Faraday Discuss. 2006, 131, 9– 22 DOI: 10.1039/B513148K [Crossref], [PubMed], [CAS], [Google
Scholar]

1485. Coskun, A.; Spruell, J. M.; Barin, G.; Dichtel, W. R.; Flood, A. H.; Botros, Y. Y.; Stoddart, J. F. High hopes: can
molecular electronics realise its potential? Chem. Soc. Rev. 2012, 41, 4827– 4859 DOI: 10.1039/c2cs35053j
[Crossref], [PubMed], [CAS], [Google Scholar]

1486. Steuerman, D. W.; Tseng, H. R.; Peters, A. J.; Flood, A. H.; Jeppesen, J. O.; Nielsen, K. A.; Stoddart, J. F.;
Heath, J. R. Molecular-mechanical switch-based solid-state electrochromic devices Angew. Chem., Int. Ed. 2004,
43, 6486– 6491 DOI: 10.1002/anie.200461723 [Crossref], [PubMed], [CAS], [Google Scholar]

1487. Deng, W. Q.; Muller, R. P.; Goddard, W. A., III Mechanism of the Stoddart-Heath bistable rotaxane
molecular switch J. Am. Chem. Soc. 2004, 126, 13562– 13563 DOI: 10.1021/ja036498x [ACS Full Text ],
[CAS], [Google Scholar]

1488. Norgaard, K.; Laursen, B.; Nygaard, S.; Kjaer, K.; Tseng, H.; Flood, A. H.; Stoddart, J. F.; Bjornholm, T.
Structural evidence of mechanical shuttling in condensed monolayers of bistable rotaxane molecules Angew.
Chem., Int. Ed. 2005, 44, 7035– 7039 DOI: 10.1002/anie.200501538 [Crossref], [PubMed], [CAS], [Google Scholar]

1489. Dichtel, W.; Heath, J. R.; Stoddart, J. F. Designing bistable [2] rotaxanes for molecular electronic devices
Philos. Trans. R. Soc., A 2007, 365, 1607– 1625 DOI: 10.1098/rsta.2007.2034 [Crossref], [PubMed],
[CAS], [Google Scholar]

1490. Nygaard, S.; Leung, K. C. F.; Aprahamian, I.; Ikeda, T.; Saha, S.; Laursen, B. W.; Kim, S. Y.; Hansen, S. W.;
Stein, P. C.; Flood, A. H.; Stoddart, J. F.; Jeppesen, J. O. Functionally rigid bistable [2]rotaxanes J. Am. Chem. Soc.
2007, 129, 960– 970 DOI: 10.1021/ja0663529 [ACS Full Text ], [CAS], [Google Scholar]

1491. Coskun, A.; Saha, S.; Aprahamian, I.; Stoddart, J. F. A reverse donor-acceptor bistable [2]Catenane Org.
Lett. 2008, 10, 3187– 3190 DOI: 10.1021/ol800931z [ACS Full Text ], [CAS], [Google Scholar]

1492. Dey, S. K.; Coskun, A.; Fahrenbach, A. C.; Barin, G.; Basuray, A. N.; Trabolsi, A.; Botros, Y. Y.; Stoddart, J. F. A
redox-active reverse donor-acceptor bistable [2]rotaxane Chem. Sci. 2011, 2, 1046– 1053
 DOI: 10.1039/C0SC00586J [Crossref], [CAS], [Google Scholar]
1493. Wang, C.; Cao, D.; Fahrenbach, A. C.; Grunder, S.; Dey, S. K.; Sarjeant, A. A.; Stoddart, J. F. The effects of
conformation on the noncovalent bonding interactions in a bistable donor-acceptor [3]catenane Chem.  
 Commun.
2012, 48, 9245– 9247 DOI: 10.1039/c2cc34190e [Crossref], [PubMed], [CAS], [Google Scholar]

1494. Fahrenbach, A. C.; Bruns, C. J.; Li, H.; Trabolsi, A.; Coskun, A.; Stoddart, J. F. Ground-State Kinetics of
Bistable Redox-Active Donor-Acceptor Mechanically Interlocked Molecules Acc. Chem. Res. 2014, 47, 482– 493
 DOI: 10.1021/ar400161z [ACS Full Text ], [CAS], [Google Scholar]

1495. Olson, M. A.; Braunschweig, A. B.; Fang, L.; Ikeda, T.; Klajn, R.; Trabolsi, A.; Wesson, P. J.; Benitez, D.; Mirkin,
C. A.; Grzybowski, B. A., et al. A Bistable Poly[2]catenane Forms Nanosuperstructures Angew. Chem., Int. Ed.
2009, 48, 1792– 1797 DOI: 10.1002/anie.200804558 [Crossref], [PubMed], [CAS], [Google Scholar]

1496. Asakawa, M.; Higuchi, M.; Mattersteig, G.; Nakamura, T.; Pease, A. R.; Raymo, F. M.; Shimizu, T.; Stoddart,
J. F. Current/Voltage Characteristics of Monolayers of Redox-Switchable [2]Catenanes on Gold Adv. Mater. 2000,
12, 1099– 1102 DOI: 10.1002/1521-4095(200008)12:15<1099::AID-ADMA1099>3.0.CO;2-2 [Crossref],
[CAS], [Google Scholar]

1497. Deng, W.-Q.; Flood, A. H.; Stoddart, J. F.; Goddard, W. A., III An Electrochemical Color-Switchable RGB Dye:
Tristable [2]Catenane J. Am. Chem. Soc. 2005, 127, 15994– 15995 DOI: 10.1021/ja0431298 [ACS Full Text ],
[CAS], [Google Scholar]

1498. Kim, Y.-H.; Jang, S.; Jang, Y.; Goddard, W. A., III First-Principles Study of the Switching Mechanism of
[2]Catenane Molecular Electronic Devices Phys. Rev. Lett. 2005, 94, 156801
 DOI: 10.1103/PhysRevLett.94.156801 [Crossref], [PubMed], [CAS], [Google Scholar]

1499. Ahuja, R. C.; Caruso, P.-L.; Moebius, D.; Philp, D.; Preece, J. A.; Ringsdorf, H.; Stoddart, J. F.; Wildburg, G.
Thin Solid Films 1996, 285, 671– 677 DOI: 10.1016/S0040-6090(95)08418-5 [Crossref], [CAS], [Google Scholar]

1500. Brown, C. L.; Jonas, U.; Preece, J. A.; Ringsdorf, H.; Seitz, M.; Stoddart, J. F. Introduction of [2]Catenanes
into Langmuir Films and Langmuir–Blodgett Multilayers. A Possible Strategy for Molecular Information Storage
Materials Langmuir 2000, 16, 1924– 1930 DOI: 10.1021/la990791m [ACS Full Text ], [CAS], [Google Scholar]

1501. Norgaard, K.; Jeppesen, J. O.; Laursen, B. W.; Simonsen, J. B.; Weygand, M. J.; Kjaer, K.; Stoddart, J. F.;
Bjornholm, T. J. Phys. Chem. B 2005, 109, 1063– 1066 DOI: 10.1021/jp0448494 [ACS Full Text ], [CAS], [Google
Scholar]
1502. Heinrich, T.; Traulsen, C. H. H.; Holzweber, M.; Richter, S.; Kunz, V.; Kastner, S. K.; Krabbenborg, S. O.;
and
Huskens, J.; Unger, W. E. S.; Schalley, C. A. Coupled Molecular Switching Processes in Ordered Mono- 
Multilayers of Stimulus-Responsive Rotaxanes on Gold Surfaces J. Am. Chem. Soc. 2015, 137, 4382– 4390
 DOI: 10.1021/ja512654d [ACS Full Text ], [CAS], [Google Scholar]

1503. Pace, G.; Petitjean, A.; Lalloz-Vogel, M.-N.; Harrow eld, J.; Lehn, J.-M.; Samorì, P. Subnanometer-Resolved
Patterning of Bicomponent Self-Assembled Monolayers on Au(111) Angew. Chem., Int. Ed. 2008, 47, 2484– 2488
 DOI: 10.1002/anie.200704731 [Crossref], [PubMed], [CAS], [Google Scholar]

1504. Jang, S. S.; Jang, Y. H.; Kim, Y.-H.; Goddard, W. A., III; Flood, A. H.; Laursen, B. W.; Tseng, H.-R.; Stoddart, J.
F.; Jeppesen, J. O.; Choik, J. W., et al. Structures and Properties of Self-Assembled Monolayers of Bistable
[2]Rotaxanes on Au (111) Surfaces from Molecular Dynamics Simulations Validated with Experiment J. Am.
Chem. Soc. 2005, 127, 1563– 1575 DOI: 10.1021/ja044530x [ACS Full Text ], [CAS], [Google Scholar]

1505. Jang, Y. H.; Jang, S. S.; Goddard, W. A., III Molecular Dynamics Simulation Study on a Monolayer of Half
[2]Rotaxane Self-Assembled on Au(111) J. Am. Chem. Soc. 2005, 127, 4959– 4964 DOI: 10.1021/ja044762w
[ACS Full Text ], [CAS], [Google Scholar]

1506. Richter, S.; Poppenberg, J.; Traulsen, C. H. H.; Darlatt, E.; Sokolowski, A.; Sattler, D.; Unger, W. E. S.;
Schalley, C. A. Deposition of Ordered Layers of Tetralactam Macrocycles and Ether Rotaxanes on Pyridine-
Terminated Self-Assembled Monolayers on Gold J. Am. Chem. Soc. 2012, 134, 16289– 16297
 DOI: 10.1021/ja306212m [ACS Full Text ], [CAS], [Google Scholar]

1507. Wong, E. W.; Collier, C. P.; Behloradsky, M.; Raymo, F. M.; Stoddart, J. F.; Heath, J. R. Fabrication and
transport properties of single-molecule-thick electrochemical junctions J. Am. Chem. Soc. 2000, 122, 5831–
5840 DOI: 10.1021/ja993890v [ACS Full Text ], [CAS], [Google Scholar]

1508. Collier, C. P.; Jeppesen, J. O.; Luo, Y.; Perkins, J.; Wong, E. W.; Heath, J. R.; Stoddart, J. F. Molecular-based
electronically switchable tunnel junction devices J. Am. Chem. Soc. 2001, 123, 12632– 12641
 DOI: 10.1021/ja0114456 [ACS Full Text ], [CAS], [Google Scholar]

1509. Lee, I. C.; Frank, C. W.; Yamamoto, T.; Tseng, H. R.; Flood, A. H.; Stoddart, J. F.; Jeppesen, J. O. Langmuir
and Langmuir-Blodgett lms of amphiphilic bistable rotaxanes Langmuir 2004, 20, 5809– 5828
 DOI: 10.1021/la0361518 [ACS Full Text ], [CAS], [Google Scholar]

1510. Jang, S. S.; Jang, Y. H.; Kim, Y. H.; Goddard, W. A., III; Choi, J. W.; Heath, J. R.; Laursen, B. W.; Flood, A. H.;
Stoddart, J. F.; Norgaard, K., et al. Molecular dynamics simulation of amphiphilic bistable [2]rotaxane Langmuir
monolayers at the air/water interface J. Am. Chem. Soc. 2005, 127, 14804– 14816 DOI: 10.1021/ja0531531
[ACS Full Text ], [CAS], [Google Scholar]   

1511. Mendes, P. M.; Lu, W.; Tseng, H. R.; Shinder, S.; Iijima, T.; Miyaji, M.; Knobler, C. M.; Stoddart, J. F. A soliton
phenomenon in langmuir monolayers of amphiphilic bistable rotaxanes J. Phys. Chem. B 2006, 110, 3845– 3848
 DOI: 10.1021/jp058287f [ACS Full Text ], [CAS], [Google Scholar]

1512. Ferri, V.; Elbing, M.; Pace, G.; Dickey, M. D.; Zharnikov, M.; Samori, P.; Mayor, M.; Rampi, M. A. Light-
powered electrical switch based on cargo-lifting azobenzene monolayers Angew. Chem., Int. Ed. 2008, 47, 3407–
3409 DOI: 10.1002/anie.200705339 [Crossref], [PubMed], [CAS], [Google Scholar]

1513. Green, J. E.; Choi, J. W.; Boukai, A.; Bunimovich, Y.; Johnston-Halperin, E.; DeIonno, E.; Luo, Y.; Sheriff, B. A.;
Xu, K.; Shin, Y. S.; Tseng, H. R.; Stoddart, J. F.; Heath, J. R. A 160-kilobit molecular electronic memory patterned at
10(11) bits per square centimetre Nature 2007, 445, 414– 417 DOI: 10.1038/nature05462 [Crossref], [PubMed],
[CAS], [Google Scholar]

1514. Yu, H.; Luo, Y.; Beverly, K.; Stoddart, J. F.; Tseng, H. R.; Heath, J. R. The molecule-electrode interface in
single-molecule transistors Angew. Chem., Int. Ed. 2003, 42, 5706– 5711 DOI: 10.1002/anie.200352352
[Crossref], [PubMed], [CAS], [Google Scholar]

1515. Lau, C. N.; Stewart, D. R.; Williams, R. S.; Bockrath, M. Direct Observation of Nanoscale Switching Centers
in Metal/Molecule/Metal Structures Nano Lett. 2004, 4, 569– 572 DOI: 10.1021/nl035117a [ACS Full Text ],
[CAS], [Google Scholar]

1516. Avellini, T.; Li, H.; Coskun, A.; Barin, G.; Trabolsi, A.; Basuray, A. N.; Dey, S. K.; Credi, A.; Silvi, S.; Stoddart, J.
F.; Venturi, M. Photoinduced Memory Effect in a Redox Controllable Bistable Mechanical Molecular Switch
Angew. Chem., Int. Ed. 2012, 51, 1611– 1615 DOI: 10.1002/anie.201107618 [Crossref], [PubMed], [CAS], [Google
Scholar]

1517. Ichimura, K. Photoalignment of Liquid-Crystal Systems Chem. Rev. 2000, 100, 1847– 1873
 DOI: 10.1021/cr980079e [ACS Full Text ], [CAS], [Google Scholar]

1518. Hoogboom, J.; Rasing, T.; Rowan, A. E.; Nolte, R. J. M. LCD alignment layers. Controlling nematic domain
properties J. Mater. Chem. 2006, 16, 1305– 1314 DOI: 10.1039/B510579J [Crossref], [CAS], [Google Scholar]

1519. Yu, Y.; Ikeda, T. Soft Actuators Based on Liquid-Crystalline Elastomers Angew. Chem., Int. Ed. 2006, 45,
5416– 5418 DOI: 10.1002/anie.200601760 [Crossref], [PubMed], [CAS], [Google Scholar]
1520. Geelhaar, T.; Griesar, K.; Reckmann, B. 125Years of Liquid CrystalsA Scienti c Revolution in  
the Home
Angew. Chem., Int. Ed. 2013, 52, 8798– 8809 DOI: 10.1002/anie.201301457 [Crossref], [PubMed], [CAS], [Google
Scholar]

1521. Feringa, B. L.; Huck, N. P. M.; van Doren, H. A. Chiroptical Switching between Liquid Crystalline Phases J.
Am. Chem. Soc. 1995, 117, 9929– 9930 DOI: 10.1021/ja00144a027 [ACS Full Text ], [CAS], [Google Scholar]

1522. Huck, N. P. M.; Jager, W. F.; de Lange, B.; Feringa, B. L. Dynamic Control and Ampli cation of Molecular
Chirality by Circular Polarized Light Science 1996, 273, 1686– 1688 DOI: 10.1126/science.273.5282.1686
[Crossref], [CAS], [Google Scholar]

1523. Sagisaka, T.; Yokoyama, Y. Reversible Control of the Pitch of Cholesteric Liquid Crystals by
Photochromism of Chiral Fulgide Derivatives Bull. Chem. Soc. Jpn. 2000, 73, 191– 196 DOI: 10.1246/bcsj.73.191
[Crossref], [CAS], [Google Scholar]

1524. Palffy-Muhoray, P.; Kosa, T.; Weinan, E. Dynamics of a Light Driven Molecular Motor Mol. Cryst. Liq. Cryst.
2002, 375, 577– 591 DOI: 10.1080/10587250210584 [Crossref], [CAS], [Google Scholar]

1525. van Delden, R. A.; Mecca, T.; Rosini, C.; Feringa, B. L. A chiroptical molecular switch with distinct chiral
and photochromic entities and its application in optical switching of a cholesteric liquid crystal Chem. - Eur. J.
2004, 10, 61– 70 DOI: 10.1002/chem.200305276 [Crossref], [PubMed], [CAS], [Google Scholar]

1526. Kosa, T.; Palffy-Muhoray, P. Brownian motors in the photoalignment of liquid crystals Int. J. Eng. Sci.
2000, 38, 1077– 1084 DOI: 10.1016/S0020-7225(99)00107-X [Crossref], [CAS], [Google Scholar]

1527. Palffy-Muhoray, P.; Kosa, T.; E, W. Brownian motors in the photoalignment of liquid crystals Appl. Phys. A:
Mater. Sci. Process. 2002, 75, 293– 300 DOI: 10.1007/s003390201321 [Crossref], [CAS], [Google Scholar]

1528. van Delden, R. A.; Koumura, N.; Harada, N.; Feringa, B. L. Unidirectional rotary motion in a liquid crystalline
environment: color tuning by a molecular motor Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 4945– 4949
 DOI: 10.1073/pnas.062660699 [Crossref], [PubMed], [CAS], [Google Scholar]

1529. van Delden, R. A.; van Gelder, M. B.; Huck, N. P. M.; Feringa, B. L. Controlling the Color of Cholesteric
Liquid-Crystalline Films by Photoirradiation of a Chiroptical Molecular Switch Used as Dopant Adv. Funct. Mater.
2003, 13, 319– 324 DOI: 10.1002/adfm.200304313 [Crossref], [CAS], [Google Scholar]
1530. Setaka, W.; Yamaguchi, K. Thermal modulation of birefringence observed in a crystalline molecular
  
gyrotop Proc. Natl. Acad. Sci. U. S. A. 2012, 109, 9271– 9275 DOI: 10.1073/pnas.1114733109 [Crossref],
[PubMed], [CAS], [Google Scholar]

1531. Setaka, W.; Yamaguchi, K. A Molecular Balloon: Expansion of a Molecular Gyrotop Cage Due to Rotation
of the Phenylene Rotor J. Am. Chem. Soc. 2012, 134, 12458– 12461 DOI: 10.1021/ja305822e [ACS Full Text ],
[CAS], [Google Scholar]

1532. Setaka, W.; Higa, S.; Yamaguchi, K. Ring-closing metathesis for the synthesis of a molecular gyrotop Org.
Biomol. Chem. 2014, 12, 3354– 3357 DOI: 10.1039/c4ob00470a [Crossref], [PubMed], [CAS], [Google Scholar]

1533. Suzaki, Y.; Taira, T.; Osakada, K.; Horie, M. Rotaxanes and pseudorotaxanes with Fe-, Pd- and Pt-
containing axles. Molecular motion in the solid state and aggregation in solution Dalton Trans. 2008, 4823–
4833 DOI: 10.1039/b804125c [Crossref], [PubMed], [CAS], [Google Scholar]

1534. Horie, M.; Suzaki, Y.; Hashizume, D.; Abe, T.; Wu, T.; Sassa, T.; Hosokai, T.; Osakada, K. Thermally-induced
phase transition of pseudorotaxane crystals: changes in conformation and interaction of the molecules and
optical properties of the crystals J. Am. Chem. Soc. 2012, 134, 17932– 17944 DOI: 10.1021/ja304406c [ACS Full
Text ], [CAS], [Google Scholar]

1535. Horie, M.; Sassa, T.; Hashizume, D.; Suzaki, Y.; Osakada, K.; Wada, T. A Crystalline Supramolecular Switch:
Controlling the Optical Anisotropy through the Collective Dynamic Motion of Molecules Angew. Chem., Int. Ed.
2007, 46, 4983– 4986 DOI: 10.1002/anie.200700708 [Crossref], [PubMed], [CAS], [Google Scholar]

1536. Leigh, D. A.; Morales, M. A.; Perez, E. M.; Wong, J. K.; Saiz, C. G.; Slawin, A. M.; Carmichael, A. J.;
Haddleton, D. M.; Brouwer, A. M.; Buma, W. J.; Wurpel, G. W.; Leon, S.; Zerbetto, F. Patterning through controlled
submolecular motion: rotaxane-based switches and logic gates that function in solution and polymer lms
Angew. Chem., Int. Ed. 2005, 44, 3062– 3067 DOI: 10.1002/anie.200500101 [Crossref], [PubMed], [CAS], [Google
Scholar]

1537. Yasushi Okumura, K. I. The Polyrotaxane Gel: A Topological Gel by Figure-of-Eight Cross-links Adv. Mater.
2001, 13, 485– 487 DOI: 10.1002/1521-4095(200104)13:7<485::AID-ADMA485>3.0.CO;2-T [Crossref], [Google
Scholar]

1538. Ito, K. Novel Cross-Linking Concept of Polymer Network: Synthesis, Structure, and Properties of Slide-
Ring Gels with Freely Movable Junctions Polym. J. 2007, 39, 489– 499 DOI: 10.1295/polymj.PJ2006239
[Crossref], [CAS], [Google Scholar]

  
1539. Ma, X.; Wang, Q.; Qu, D.; Xu, Y.; Ji, F.; Tian, H. A light-driven pseudo[4]rotaxane encoded by induced circular
dichroism in a hydrogel Adv. Funct. Mater. 2007, 17, 829– 837 DOI: 10.1002/adfm.200600981 [Crossref],
[CAS], [Google Scholar]

1540. Taira, T.; Suzaki, Y.; Osakada, K. Thermosensitive hydrogels composed of cyclodextrin pseudorotaxanes.
Role of [3]pseudorotaxane in the gel formation Chem. Commun. 2009, 7027– 7029 DOI: 10.1039/b911667b
[Crossref], [PubMed], [CAS], [Google Scholar]

1541. Suzaki, Y.; Taira, T.; Osakada, K. Physical gels based on supramolecular gelators, including host-guest
complexes and pseudorotaxanes J. Mater. Chem. 2011, 21, 930– 938 DOI: 10.1039/C0JM02219E [Crossref],
[CAS], [Google Scholar]

1542. Zhu, L.; Ma, X.; Ji, F.; Wang, Q.; Tian, H. Effective enhancement of uorescence signals in rotaxane-doped
reversible hydrosol-gel systems Chem. - Eur. J. 2007, 13, 9216– 9222 DOI: 10.1002/chem.200700860 [Crossref],
[PubMed], [CAS], [Google Scholar]

1543. Madden, J. D. W.; Vandesteeg, N. A.; Anquetil, P. A.; Madden, P. G. A.; Takshi, A.; Pytel, R. Z.; Lafontaine, S.
R.; Wieringa, P. A.; Hunter, I. W. Arti cial muscle technology: physical principles and naval prospects IEEE J.
Oceanic Eng. 2004, 29, 706– 728 DOI: 10.1109/JOE.2004.833135 [Crossref], [Google Scholar]

1544. Baughman, R. H. Playing Nature’s Game with Arti cial Muscles Science 2005, 308, 63– 65
 DOI: 10.1126/science.1099010 [Crossref], [PubMed], [CAS], [Google Scholar]

1545. Frank, S.; Lauterbur, P. C. Voltage-sensitive magnetic gels as magnetic resonance monitoring agents
Nature 1993, 363, 334– 336 DOI: 10.1038/363334a0 [Crossref], [PubMed], [CAS], [Google Scholar]

1546. Osada, Y.; Gong, J. Stimuli-responsive polymer gels and their application to chemomechanical systems
Prog. Polym. Sci. 1993, 18, 187– 226 DOI: 10.1016/0079-6700(93)90025-8 [Crossref], [CAS], [Google Scholar]

1547. Osada, Y.; Gong, J.-P. Soft and Wet Materials: Polymer Gels Adv. Mater. 1998, 10, 827– 837
 DOI: 10.1002/(SICI)1521-4095(199808)10:11<827::AID-ADMA827>3.0.CO;2-L [Crossref], [CAS], [Google Scholar]

1548. Berndt, I.; Popescu, C.; Wortmann, F.-J.; Richtering, W. Mechanics versus Thermodynamics: Swelling in
Multiple-Temperature-Sensitive Core–Shell Microgels Angew. Chem., Int. Ed. 2006, 45, 1081– 1085
 DOI: 10.1002/anie.200502893 [Crossref], [PubMed], [CAS], [Google Scholar]

  
1549. Juodkazis, S.; Mukai, N.; Wakaki, R.; Yamaguchi, A.; Matsuo, S.; Misawa, H. Reversible phase transitions in
polymer gels induced by radiation forces Nature 2000, 408, 178– 181 DOI: 10.1038/35041522 [Crossref],
[PubMed], [CAS], [Google Scholar]

1550. Yoshida, R.; Takahashi, T.; Yamaguchi, T.; Ichijo, H. Self-Oscillating Gel J. Am. Chem. Soc. 1996, 118,
5134– 5135 DOI: 10.1021/ja9602511 [ACS Full Text ], [CAS], [Google Scholar]

1551. Berndt, I.; Pedersen, J. S.; Richtering, W. Temperature-Sensitive Core–Shell Microgel Particles with Dense
Shell Angew. Chem., Int. Ed. 2006, 45, 1737– 1741 DOI: 10.1002/anie.200503888 [Crossref], [PubMed],
[CAS], [Google Scholar]

1552. Kiyonaka, S.; Sugiyasu, K.; Shinkai, S.; Hamachi, I. First Thermally Responsive Supramolecular Polymer
Based on Glycosylated Amino Acid J. Am. Chem. Soc. 2002, 124, 10954– 10955 DOI: 10.1021/ja027277e [ACS
Full Text ], [CAS], [Google Scholar]

1553. You, J.-O.; Almeda, D.; Ye, G.; Auguste, D. Bioresponsive matrices in drug delivery J. Biol. Eng. 2010, 4, 15
 DOI: 10.1186/1754-1611-4-15 [Crossref], [PubMed], [CAS], [Google Scholar]

1554. Chaterji, S.; Kwon, I. K.; Park, K. Smart polymeric gels: Rede ning the limits of biomedical devices Prog.
Polym. Sci. 2007, 32, 1083– 1122 DOI: 10.1016/j.progpolymsci.2007.05.018 [Crossref], [PubMed], [CAS], [Google
Scholar]

1555. Schild, H. G. Poly(N-isopropylacrylamide): experiment, theory and application Prog. Polym. Sci. 1992, 17,
163– 249 DOI: 10.1016/0079-6700(92)90023-R [Crossref], [CAS], [Google Scholar]

1556. Shahinpoor, M.; Bar-Cohen, Y.; Simpson, J. O.; Smith, J. Ionic polymer-metal composites (IPMCs) as
biomimetic sensors, actuators and arti cial muscles - a review Smart Mater. Struct. 1998, 7, R15– R30
 DOI: 10.1088/0964-1726/7/6/001 [Crossref], [CAS], [Google Scholar]

1557. Urry, D. W. Physical Chemistry of Biological Free Energy Transduction As Demonstrated by Elastic
Protein-Based Polymers J. Phys. Chem. B 1997, 101, 11007– 11028 DOI: 10.1021/jp972167t [ACS Full Text ],
[CAS], [Google Scholar]
1558. Urry, D. W. Five axioms for the functional design of peptide-based polymers as molecular machines and
materials: Principle for macromolecular assemblies Biopolymers 1998, 47, 167– 178   
 DOI: 10.1002/(SICI)1097-0282(1998)47:2<167::AID-BIP6>3.0.CO;2-S [Crossref], [CAS], [Google Scholar]

1559. Steinberg, I. Z.; Oplatka, A.; Katchalsky, A. Mechanochemical Engines Nature 1966, 210, 568– 571
 DOI: 10.1038/210568a0 [Crossref], [CAS], [Google Scholar]

1560. Sussman, M. V.; Katchalsky, A. Mechanochemical Turbine: A New Power Cycle Science 1970, 167, 45–
47 DOI: 10.1126/science.167.3914.45 [Crossref], [PubMed], [CAS], [Google Scholar]

1561. Yoshida, R.; Kokufuta, E.; Yamaguchi, T. Beating polymer gels coupled with a nonlinear chemical reaction
Chaos 1999, 9, 260– 266 DOI: 10.1063/1.166402 [Crossref], [PubMed], [CAS], [Google Scholar]

1562. Yoshida, R.; Otoshi, G.; Yamaguchi, T.; Kokufuta, E. Traveling Chemical Waves for Measuring Solute
Diffusivity in Thermosensitive Poly(N-isopropylacrylamide) Gel J. Phys. Chem. A 2001, 105, 3667– 3672
 DOI: 10.1021/jp004187s [ACS Full Text ], [CAS], [Google Scholar]

1563. Yoshida, R.; Takei, K.; Yamaguchi, T. Self-Beating Motion of Gels and Modulation of Oscillation Rhythm
Synchronized with Organic Acid Macromolecules 2003, 36, 1759– 1761 DOI: 10.1021/ma0259618 [ACS Full
Text ], [CAS], [Google Scholar]

1564. Crook, C. J.; Smith, A.; Jones, R. A. L.; Ryan, A. J. Chemically induced oscillations in a pH-responsive
hydrogel Phys. Chem. Chem. Phys. 2002, 4, 1367– 1369 DOI: 10.1039/b109977a [Crossref], [CAS], [Google
Scholar]

1565. Howse, J. R.; Topham, P.; Crook, C. J.; Gleeson, A. J.; Bras, W.; Jones, R. A. L.; Ryan, A. J. Reciprocating
Power Generation in a Chemically Driven Synthetic Muscle Nano Lett. 2006, 6, 73– 77 DOI: 10.1021/nl0520617
[ACS Full Text ], [CAS], [Google Scholar]

1566. Jones, C. D.; Lyon, L. A. Synthesis and Characterization of Multiresponsive Core–Shell Microgels
Macromolecules 2000, 33, 8301– 8306 DOI: 10.1021/ma001398m [ACS Full Text ], [CAS], [Google Scholar]

1567. Berndt, I.; Richtering, W. Doubly Temperature Sensitive Core–Shell Microgels Macromolecules 2003, 36,
8780– 8785 DOI: 10.1021/ma034771+ [ACS Full Text ], [CAS], [Google Scholar]
1568. Beebe, D. J.; Moore, J. S.; Bauer, J. M.; Yu, Q.; Liu, R. H.; Devadoss, C.; Jo, B.-H. Functional hydrogel
structures for autonomous ow control inside micro uidic channels Nature 2000, 404, 588– 590   
 DOI: 10.1038/35007047 [Crossref], [PubMed], [CAS], [Google Scholar]

1569. Yu, Q.; Bauer, J. M.; Moore, J. S.; Beebe, D. J. Responsive biomimetic hydrogel valve for micro uidics
Appl. Phys. Lett. 2001, 78, 2589– 2591 DOI: 10.1063/1.1367010 [Crossref], [CAS], [Google Scholar]

1570. Liu, R. H.; Yu, Q.; Beebe, D. J. Fabrication and characterization of hydrogel-based microvalves J.
Microelectromech. Syst. 2002, 11, 45– 53 DOI: 10.1109/84.982862 [Crossref], [CAS], [Google Scholar]

1571. Osada, Y.; Okuzaki, H.; Hori, H. A polymer gel with electrically driven motility Nature 1992, 355, 242– 244
 DOI: 10.1038/355242a0 [Crossref], [CAS], [Google Scholar]

1572. Galaev, I. Y.; Mattiasson, B. ‘Smart’ polymers and what they could do in biotechnology and medicine
Trends Biotechnol. 1999, 17, 335– 340 DOI: 10.1016/S0167-7799(99)01345-1 [Crossref], [PubMed],
[CAS], [Google Scholar]

1573. Peppas, N. A.; Huang, Y.; Torres-Lugo, M.; Ward, J. H.; Zhang, J. Physicochemical Foundations and
Structural Design of Hydrogels in Medicine and Biology Annu. Rev. Biomed. Eng. 2000, 2, 9– 29
 DOI: 10.1146/annurev.bioeng.2.1.9 [Crossref], [PubMed], [CAS], [Google Scholar]

1574. Qiu, Y.; Park, K. Environment-sensitive hydrogels for drug delivery Adv. Drug Delivery Rev. 2001, 53, 321–
339 DOI: 10.1016/S0169-409X(01)00203-4 [Crossref], [PubMed], [CAS], [Google Scholar]

1575. Jeong, B.; Gutowska, A. Lessons from nature: stimuli-responsive polymers and their biomedical
applications Trends Biotechnol. 2002, 20, 305– 311 DOI: 10.1016/S0167-7799(02)01962-5 [Crossref], [PubMed],
[CAS], [Google Scholar]

1576. Kopecek, J. Polymer chemistry: Swell gels Nature 2002, 417, 388– 391 DOI: 10.1038/417388a
[Crossref], [PubMed], [CAS], [Google Scholar]

1577. Langer, R.; Tirrell, D. A. Designing materials for biology and medicine Nature 2004, 428, 487– 492
 DOI: 10.1038/nature02388 [Crossref], [PubMed], [CAS], [Google Scholar]

1578. Lin, K.-J.; Fu, S.-J.; Cheng, C.-Y.; Chen, W.-H.; Kao, H.-M. Towards Electrochemical Arti cial Muscles: A
Supramolecular Machine Based on a One-Dimensional Copper-Containing Organophosphonate System Angew.
Chem., Int. Ed. 2004, 43, 4186– 4189 DOI: 10.1002/anie.200454159 [Crossref], [PubMed], [CAS], [Google Scholar]

  
1579. Li, C.; Madsen, J.; Armes, S. P.; Lewis, A. L. A New Class of Biochemically Degradable, Stimulus-
Responsive Triblock Copolymer Gelators Angew. Chem., Int. Ed. 2006, 45, 3510– 3513
 DOI: 10.1002/anie.200600324 [Crossref], [PubMed], [CAS], [Google Scholar]

1580. Samoei, G. K.; Wang, W.; Escobedo, J. O.; Xu, X.; Schneider, H.-J.; Cook, R. L.; Strongin, R. M. A
Chemomechanical Polymer that Functions in Blood Plasma with High Glucose Selectivity Angew. Chem., Int. Ed.
2006, 45, 5319– 5322 DOI: 10.1002/anie.200601398 [Crossref], [PubMed], [CAS], [Google Scholar]

1581. Dong, L.; Agarwal, A. K.; Beebe, D. J.; Jiang, H. Adaptive liquid microlenses activated by stimuli-responsive
hydrogels Nature 2006, 442, 551– 554 DOI: 10.1038/nature05024 [Crossref], [PubMed], [CAS], [Google Scholar]

1582. Oya, T.; Enoki, T.; Grosberg, A. Y.; Masamune, S.; Sakiyama, T.; Takeoka, Y.; Tanaka, K.; Wang, G.; Yilmaz, Y.;
Feld, M. S.; Dasari, R.; Tanaka, T. Reversible Molecular Adsorption Based on Multiple-Point Interaction by
Shrinkable Gels Science 1999, 286, 1543– 1545 DOI: 10.1126/science.286.5444.1543 [Crossref], [PubMed],
[CAS], [Google Scholar]

1583. Pennadam, S. S.; Lavigne, M. D.; Dutta, C. F.; Firman, K.; Mernagh, D.; Górecki, D. C.; Alexander, C. Control
of A Multisubunit DNA Motor by a Thermoresponsive Polymer Switch J. Am. Chem. Soc. 2004, 126, 13208–
13209 DOI: 10.1021/ja045275j [ACS Full Text ], [CAS], [Google Scholar]

1584. Kokufata, E.; Zhang, Y.-Q.; Tanaka, T. Saccharide-sensitive phase transition of a lectin-loaded gel Nature
1991, 351, 302– 304 DOI: 10.1038/351302a0 [Crossref], [CAS], [Google Scholar]

1585. Irie, M.; Misumi, Y.; Tanaka, T. Stimuli-responsive polymers: chemical induced reversible phase separation
of an aqueous solution of poly(N-isopropylacrylamide) with pendent crown ether groups Polymer 1993, 34,
4531– 4535 DOI: 10.1016/0032-3861(93)90160-C [Crossref], [CAS], [Google Scholar]

1586. Lee, K.; Asher, S. A. Photonic Crystal Chemical Sensors: pH and Ionic Strength J. Am. Chem. Soc. 2000,
122, 9534– 9537 DOI: 10.1021/ja002017n [ACS Full Text ], [CAS], [Google Scholar]

1587. Miyata, T.; Asami, N.; Uragami, T. A reversibly antigen-responsive hydrogel Nature 1999, 399, 766– 769
 DOI: 10.1038/21619 [Crossref], [PubMed], [CAS], [Google Scholar]
1588. Schneider, H.-J.; Tianjun, L.; Lomadze, N. Molecular Recognition in a Supramolecular Polymer System
  
Translated into Mechanical Motion Angew. Chem., Int. Ed. 2003, 42, 3544– 3546 DOI: 10.1002/anie.200219965
[Crossref], [PubMed], [CAS], [Google Scholar]

1589. Sharma, A. C.; Jana, T.; Kesavamoorthy, R.; Shi, L.; Virji, M. A.; Finegold, D. N.; Asher, S. A. A General
Photonic Crystal Sensing Motif: Creatinine in Bodily Fluids J. Am. Chem. Soc. 2004, 126, 2971– 2977
 DOI: 10.1021/ja038187s [ACS Full Text ], [CAS], [Google Scholar]

1590. Goponenko, A. V.; Asher, S. A. Modeling of Stimulated Hydrogel Volume Changes in Photonic Crystal Pb2+
Sensing Materials J. Am. Chem. Soc. 2005, 127, 10753– 10759 DOI: 10.1021/ja051456p [ACS Full Text ],
[CAS], [Google Scholar]

1591. Holtz, J. H.; Asher, S. A. Polymerized colloidal crystal hydrogel lms as intelligent chemical sensing
materials Nature 1997, 389, 829– 832 DOI: 10.1038/39834 [Crossref], [PubMed], [CAS], [Google Scholar]

1592. Holtz, J. H.; Holtz, J. S. W.; Munro, C. H.; Asher, S. A. Intelligent Polymerized Crystalline Colloidal Arrays:
Novel Chemical Sensor Materials Anal. Chem. 1998, 70, 780– 791 DOI: 10.1021/ac970853i [ACS Full Text ],
[CAS], [Google Scholar]

1593. Asher, S. A.; Alexeev, V. L.; Goponenko, A. V.; Sharma, A. C.; Lednev, I. K.; Wilcox, C. S.; Finegold, D. N.
Photonic Crystal Carbohydrate Sensors: Low Ionic Strength Sugar Sensing J. Am. Chem. Soc. 2003, 125, 3322–
3329 DOI: 10.1021/ja021037h [ACS Full Text ], [CAS], [Google Scholar]

1594. Alexeev, V. L.; Sharma, A. C.; Goponenko, A. V.; Das, S.; Lednev, I. K.; Wilcox, C. S.; Finegold, D. N.; Asher, S.
A. High Ionic Strength Glucose-Sensing Photonic Crystal Anal. Chem. 2003, 75, 2316– 2323
 DOI: 10.1021/ac030021m [ACS Full Text ], [CAS], [Google Scholar]

1595. Asher, S. A.; Sharma, A. C.; Goponenko, A. V.; Ward, M. M. Photonic Crystal Aqueous Metal Cation Sensing
Materials Anal. Chem. 2003, 75, 1676– 1683 DOI: 10.1021/ac026328n [ACS Full Text ], [CAS], [Google Scholar]

1596. Schneider, H.-J.; Liu, T. Large macroscopic size changes in chemomechanical polymers with binding
sites for metal ions Chem. Commun. 2004, 100– 101 DOI: 10.1039/b310939a [Crossref], [PubMed],
[CAS], [Google Scholar]

1597. Schneider, H.-J.; Tianjun, L.; Lomadze, N. Sensitivity increase in molecular recognition by decrease of the
sensing particle size and by increase of the receptor binding site - a case with chemomechanical polymers
Chem. Commun. 2004, 2436– 2437 DOI: 10.1039/b409331c [Crossref], [PubMed], [CAS], [Google Scholar]
 
1598. Schneider, H. J.; Tianjun, L.; Lomadze, N.; Palm, B. Cooperativity in a Chemomechanical Polymer: A

Chemically Induced Macroscopic Logic Gate Adv. Mater. 2004, 16, 613– 615 DOI: 10.1002/adma.200306249
[Crossref], [CAS], [Google Scholar]

1599. Wei, Z. G.; Sandstroröm, R.; Miyazaki, S. Shape-memory materials and hybrid composites for smart
systems: Part I Shape-memory materials J. Mater. Sci. 1998, 33, 3743– 3762 DOI: 10.1023/A:1004692329247
[Crossref], [CAS], [Google Scholar]

1600. Lendlein, A.; Kelch, S. Shape-Memory Polymers Angew. Chem., Int. Ed. 2002, 41, 2034– 2057
 DOI: 10.1002/1521-3773(20020617)41:12<2034::AID-ANIE2034>3.0.CO;2-M [Crossref], [CAS], [Google Scholar]

1601. Scott, T. F.; Schneider, A. D.; Cook, W. D.; Bowman, C. N. Photoinduced Plasticity in Cross-Linked Polymers
Science 2005, 308, 1615– 1617 DOI: 10.1126/science.1110505 [Crossref], [PubMed], [CAS], [Google Scholar]

1602. Lendlein, A.; Jiang, H.; Junger, O.; Langer, R. Light-induced shape-memory polymers Nature 2005, 434,
879– 882 DOI: 10.1038/nature03496 [Crossref], [PubMed], [CAS], [Google Scholar]

1603. Baughman, R. H.; Cui, C.; Zakhidov, A. A.; Iqbal, Z.; Barisci, J. N.; Spinks, G. M.; Wallace, G. G.; Mazzoldi, A.;
De Rossi, D.; Rinzler, A. G.; Jaschinski, O.; Roth, S.; Kertesz, M. Carbon Nanotube Actuators Science 1999, 284,
1340– 1344 DOI: 10.1126/science.284.5418.1340 [Crossref], [PubMed], [CAS], [Google Scholar]

1604. Raguse, B.; Müller, K. H.; Wieczorek, L. Nanoparticle Actuators Adv. Mater. 2003, 15, 922– 926
 DOI: 10.1002/adma.200304698 [Crossref], [CAS], [Google Scholar]

1605. Herrmann, L. O.; Valev, V. K.; Tserkezis, C.; Barnard, J. S.; Kasera, S.; Scherman, O. A.; Aizpurua, J.;
Baumberg, J. J. Threading plasmonic nanoparticle strings with light Nat. Commun. 2014, 5, 4568
 DOI: 10.1038/ncomms5568 [Crossref], [PubMed], [CAS], [Google Scholar]

1606. Jager, E. W. H.; Smela, E.; Inganäs, O. Microfabricating Conjugated Polymer Actuators Science 2000, 290,
1540– 1545 DOI: 10.1126/science.290.5496.1540 [Crossref], [PubMed], [CAS], [Google Scholar]

1607. Smela, E. Conjugated Polymer Actuators for Biomedical Applications Adv. Mater. 2003, 15, 481– 494
 DOI: 10.1002/adma.200390113 [Crossref], [CAS], [Google Scholar]
1608. Baughman, R. H. Conducting polymer arti cial muscles Synth. Met. 1996, 78, 339– 353
 DOI: 10.1016/0379-6779(96)80158-5 [Crossref], [CAS], [Google Scholar]   

1609. Otero, T. F.; Sansieña, J. M. Soft and Wet Conducting Polymers for Arti cial Muscles Adv. Mater. 1998,
10, 491– 494 DOI: 10.1002/(SICI)1521-4095(199804)10:6<491::AID-ADMA491>3.0.CO;2-Q [Crossref], [PubMed],
[CAS], [Google Scholar]

1610. Sansinena, J.; Olazabal, V.; Otero, T.; Sansinena, J.; Polo da Fonseca, C.; De Paoli, M.-A. A solid state
arti cial muscle based on polypyrrole and a solid polymeric electrolyte working in air Chem. Commun. 1997,
2217– 2218 DOI: 10.1039/a705341j [Crossref], [CAS], [Google Scholar]

1611. Madden, J. D.; Cush, R. A.; Kanigan, T. S.; Brenan, C. J.; Hunter, I. W. Encapsulated polypyrrole actuators
Synth. Met. 1999, 105, 61– 64 DOI: 10.1016/S0379-6779(99)00034-X [Crossref], [CAS], [Google Scholar]

1612. Lu, W.; Fadeev, A. G.; Qi, B.; Smela, E.; Mattes, B. R.; Ding, J.; Spinks, G. M.; Mazurkiewicz, J.; Zhou, D.;
Wallace, G. G.; MacFarlane, D. R.; Forsyth, S. A.; Forsyth, M. Use of Ionic Liquids for π-Conjugated Polymer
Electrochemical Devices Science 2002, 297, 983– 987 DOI: 10.1126/science.1072651 [Crossref], [PubMed],
[CAS], [Google Scholar]

1613. Yu, H.-h.; Xu, B.; Swager, T. M. A Proton-Doped Calix[4]arene-Based Conducting Polymer J. Am. Chem.
Soc. 2003, 125, 1142– 1143 DOI: 10.1021/ja028545b [ACS Full Text ], [CAS], [Google Scholar]

1614. Scherlis, D. A.; Marzari, N. π-Stacking in Thiophene Oligomers as the Driving Force for Electroactive
Materials and Devices J. Am. Chem. Soc. 2005, 127, 3207– 3212 DOI: 10.1021/ja043557d [ACS Full Text ],
[CAS], [Google Scholar]

1615. Hugel, T.; Holland, N. B.; Cattani, A.; Moroder, L.; Seitz, M.; Gaub, H. E. Single-Molecule Optomechanical
Cycle Science 2002, 296, 1103– 1106 DOI: 10.1126/science.1069856 [Crossref], [PubMed], [CAS], [Google
Scholar]

1616. Kumar, G. S.; Neckers, D. C. Photochemistry of azobenzene-containing polymers Chem. Rev. 1989, 89,
1915– 1925 DOI: 10.1021/cr00098a012 [ACS Full Text ], [CAS], [Google Scholar]

1617. Natansohn, A.; Rochon, P. Photoinduced Motions in Azo-Containing Polymers Chem. Rev. 2002, 102,
4139– 4176 DOI: 10.1021/cr970155y [ACS Full Text ], [CAS], [Google Scholar]
1618. Maxein, G.; Zentel, R. Photochemical Inversion of the Helical Twist Sense in Chiral Polyisocyanates
Macromolecules 1995, 28, 8438– 8440 DOI: 10.1021/ma00128a068 [ACS Full Text ], [CAS], [Google  
 Scholar]

1619. Okamoto, Y.; Nakano, T. Asymmetric Polymerization Chem. Rev. 1994, 94, 349– 372
 DOI: 10.1021/cr00026a004 [ACS Full Text ], [CAS], [Google Scholar]

1620. Nakano, T.; Okamoto, Y. Synthetic Helical Polymers: Conformation and Function Chem. Rev. 2001, 101,
4013– 4038 DOI: 10.1021/cr0000978 [ACS Full Text ], [CAS], [Google Scholar]

1621. Dumont, M.; El Osman, A. On spontaneous and photoinduced orientational mobility of dye molecules in
polymers Chem. Phys. 1999, 245, 437– 462 DOI: 10.1016/S0301-0104(99)00096-8 [Crossref], [CAS], [Google
Scholar]

1622. Delaire, J. A.; Nakatani, K. Linear and Nonlinear Optical Properties of Photochromic Molecules and
Materials Chem. Rev. 2000, 100, 1817– 1846 DOI: 10.1021/cr980078m [ACS Full Text ], [CAS], [Google
Scholar]

1623. Kim, D. Y.; Tripathy, S. K.; Li, L.; Kumar, J. Laser-induced holographic surface relief gratings on nonlinear
optical polymer lms Appl. Phys. Lett. 1995, 66, 1166– 1168 DOI: 10.1063/1.113845 [Crossref], [CAS], [Google
Scholar]

1624. Viswanathan, N.; Yu Kim, D.; Bian, S.; Williams, J.; Liu, W.; Li, L.; Samuelson, L.; Kumar, J.; K. Tripathy, S.
Surface relief structures on azo polymer lms J. Mater. Chem. 1999, 9, 1941– 1955 DOI: 10.1039/a902424g
[Crossref], [CAS], [Google Scholar]

1625. Yager, K. G.; Barrett, C. J. All-optical patterning of azo polymer lms Curr. Opin. Solid State Mater. Sci.
2001, 5, 487– 494 DOI: 10.1016/S1359-0286(02)00020-7 [Crossref], [CAS], [Google Scholar]

1626. Tanchak, O. M.; Barrett, C. J. Light-Induced Reversible Volume Changes in Thin Films of Azo Polymers:
The Photomechanical Effect Macromolecules 2005, 38, 10566– 10570 DOI: 10.1021/ma051564w [ACS Full
Text ], [CAS], [Google Scholar]

1627. Labarthet, F. L.; Bruneel, J.-L.; Buffeteau, T.; Sourisseau, C.; Huber, M. R.; Zilker, S. J.; Bieringer, T.
Photoinduced orientations of azobenzene chromophores in two distinct holographic diffraction gratings as
studied by polarized Raman confocal microspectrometry Phys. Chem. Chem. Phys. 2000, 2, 5154– 5167
 DOI: 10.1039/b005632o [Crossref], [CAS], [Google Scholar]
1628. Schulz, B. M.; Huber, M. R.; Bieringer, T.; Krausch, G.; Zilker, S. J. Length-scale dependence of surface
relief gratings in azobenzene side-chain polymers Synth. Met. 2001, 124, 155– 157   
 DOI: 10.1016/S0379-6779(01)00457-X [Crossref], [CAS], [Google Scholar]

1629. Xie, P.; Zhang, R. Liquid crystal elastomers, networks and gels: advanced smart materials J. Mater. Chem.
2005, 15, 2529– 2550 DOI: 10.1039/b413835j [Crossref], [CAS], [Google Scholar]

1630. Li, M. H.; Keller, P.; Li, B.; Wang, X.; Brunet, M. Light-Driven Side-On Nematic Elastomer Actuators Adv.
Mater. 2003, 15, 569– 572 DOI: 10.1002/adma.200304552 [Crossref], [CAS], [Google Scholar]

1631. Buguin, A.; Li, M.-H.; Silberzan, P.; Ladoux, B.; Keller, P. Micro-Actuators: When Arti cial Muscles Made of
Nematic Liquid Crystal Elastomers Meet Soft Lithography J. Am. Chem. Soc. 2006, 128, 1088– 1089
 DOI: 10.1021/ja0575070 [ACS Full Text ], [CAS], [Google Scholar]

1632. Finkelmann, H.; Nishikawa, E.; Pereira, G. G.; Warner, M. A New Opto-Mechanical Effect in Solids Phys.
Rev. Lett. 2001, 87, 015501 DOI: 10.1103/PhysRevLett.87.015501 [Crossref], [CAS], [Google Scholar]

1633. Hogan, P. M.; Tajbakhsh, A. R.; Terentjev, E. M. UV manipulation of order and macroscopic shape in
nematic elastomers Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 2002, 65, 041720
 DOI: 10.1103/PhysRevE.65.041720 [Crossref], [CAS], [Google Scholar]

1634. Ikeda, T.; Nakano, M.; Yu, Y.; Tsutsumi, O.; Kanazawa, A. Anisotropic Bending and Unbending Behavior of
Azobenzene Liquid-Crystalline Gels by Light Exposure Adv. Mater. 2003, 15, 201– 205
 DOI: 10.1002/adma.200390045 [Crossref], [CAS], [Google Scholar]

1635. Camacho-Lopez, M.; Finkelmann, H.; Palffy-Muhoray, P.; Shelley, M. Fast liquid-crystal elastomer swims
into the dark Nat. Mater. 2004, 3, 307– 310 DOI: 10.1038/nmat1118 [Crossref], [PubMed], [CAS], [Google Scholar]

1636. Iamsaard, S.; Aßhoff, S. J.; Matt, B.; Kudernac, T.; Cornelissen Jeroen, J. L. M.; Fletcher, S. P.; Katsonis, N.
Conversion of light into macroscopic helical motion Nat. Chem. 2014, 6, 229– 235 DOI: 10.1038/nchem.1859
[Crossref], [PubMed], [CAS], [Google Scholar]

1637. Eelkema, R.; Pollard, M. M.; Vicario, J.; Katsonis, N.; Ramon, B. S.; Bastiaansen, C. W. M.; Broer, D. J.;
Feringa, B. L. Molecular machines: Nanomotor rotates microscale objects Nature 2006, 440, 163– 163
 DOI: 10.1038/440163a [Crossref], [PubMed], [CAS], [Google Scholar]
1638. Chen, J.; Kistemaker, J. C.; Robertus, J.; Feringa, B. L. Molecular stirrers in action J. Am. Chem. Soc.
2014, 136, 14924– 14932 DOI: 10.1021/ja507711h [ACS Full Text ], [CAS], [Google Scholar]   

1639. Fritz, J.; Baller, M. K.; Lang, H. P.; Rothuizen, H.; Vettiger, P.; Meyer, E.-J.; Güntherodt, H.; Gerber, C.;
Gimzewski, J. K. Translating Biomolecular Recognition into Nanomechanics Science 2000, 288, 316– 318
 DOI: 10.1126/science.288.5464.316 [Crossref], [PubMed], [CAS], [Google Scholar]

1640. Wu, G.; Ji, H.; Hansen, K.; Thundat, T.; Datar, R.; Cote, R.; Hagan, M. F.; Chakraborty, A. K.; Majumdar, A.
Origin of nanomechanical cantilever motion generated from biomolecular interactions Proc. Natl. Acad. Sci. U. S.
A. 2001, 98, 1560– 1564 DOI: 10.1073/pnas.98.4.1560 [Crossref], [PubMed], [CAS], [Google Scholar]

1641. Dhayal, B.; Henne, W. A.; Doorneweerd, D. D.; Reifenberger, R. G.; Low, P. S. Detection of Bacillus subtilis
Spores Using Peptide-Functionalized Cantilever Arrays J. Am. Chem. Soc. 2006, 128, 3716– 3721
 DOI: 10.1021/ja0570887 [ACS Full Text ], [CAS], [Google Scholar]

1642. Wu, G.; Datar, R. H.; Hansen, K. M.; Thundat, T.; Cote, R. J.; Majumdar, A. Bioassay of prostate-speci c
antigen (PSA) using microcantilevers Nat. Biotechnol. 2001, 19, 856– 860 DOI: 10.1038/nbt0901-856 [Crossref],
[PubMed], [CAS], [Google Scholar]

1643. McKendry, R.; Zhang, J.; Arntz, Y.; Strunz, T.; Hegner, M.; Lang, H. P.; Baller, M. K.; Certa, U.; Meyer, E.;
Güntherodt, H.-J., et al. Multiple label-free biodetection and quantitative DNA-binding assays on a
nanomechanical cantilever array Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 9783– 9788
 DOI: 10.1073/pnas.152330199 [Crossref], [PubMed], [CAS], [Google Scholar]

1644. Shu, W.; Liu, D.; Watari, M.; Riener, C. K.; Strunz, T.; Welland, M. E.; Balasubramanian, S.; McKendry, R. A.
DNA Molecular Motor Driven Micromechanical Cantilever Arrays J. Am. Chem. Soc. 2005, 127, 17054– 17060
 DOI: 10.1021/ja0554514 [ACS Full Text ], [CAS], [Google Scholar]

1645. Su, M.; Dravid, V. P. Surface Combustion Microengines Based on Photocatalytic Oxidations of
Hydrocarbons at Room Temperature Nano Lett. 2005, 5, 2023– 2028 DOI: 10.1021/nl0515605 [ACS Full Text ],
[CAS], [Google Scholar]

1646. Ji, H.-F.; Feng, Y.; Xu, X.; Purushotham, V.; Thundat, T.; Brown, G. M. Photon-driven nanomechanical cyclic
motion Chem. Commun. 2004, 2532– 2533 DOI: 10.1039/b408997a [Crossref], [PubMed], [CAS], [Google
Scholar]
1647. Li, Q.; Fuks, G.; Moulin, E.; Maaloum, M.; Rawiso, M.; Kulic, I.; Foy, J. T.; Giuseppone, N. Macroscopic
 
contraction of a gel induced by the integrated motion of light-driven molecular motors Nat. Nanotechnol. 2015,
10, 161– 165 DOI: 10.1038/nnano.2014.315 [Crossref], [PubMed], [CAS], [Google Scholar]

1648. Coleman, A. C.; Beierle, J. M.; Stuart, M. C.; Macia, B.; Caroli, G.; Mika, J. T.; van Dijken, D. J.; Chen, J.;
Browne, W. R.; Feringa, B. L. Light-induced disassembly of self-assembled vesicle-capped nanotubes observed in
real time Nat. Nanotechnol. 2011, 6, 547– 552 DOI: 10.1038/nnano.2011.120 [Crossref], [PubMed],
[CAS], [Google Scholar]

1649. Milner, S. T. Polymer Brushes Science 1991, 251, 905– 914 DOI: 10.1126/science.251.4996.905
[Crossref], [PubMed], [CAS], [Google Scholar]

1650. Tsukruk, V. V. Assembly of supramolecular polymers in ultrathin lms Prog. Polym. Sci. 1997, 22, 247–
311 DOI: 10.1016/S0079-6700(96)00005-6 [Crossref], [CAS], [Google Scholar]

1651. Nath, N.; Chilkoti, A. Creating “Smart” Surfaces Using Stimuli Responsive Polymers Adv. Mater. 2002, 14,
1243– 1247 DOI: 10.1002/1521-4095(20020903)14:17<1243::AID-ADMA1243>3.0.CO;2-M [Crossref],
[CAS], [Google Scholar]

1652. Cole, M. A.; Voelcker, N. H.; Thissen, H.; Griesser, H. J. Stimuli-responsive interfaces and systems for the
control of protein–surface and cell–surface interactions Biomaterials 2009, 30, 1827– 1850
 DOI: 10.1016/j.biomaterials.2008.12.026 [Crossref], [PubMed], [CAS], [Google Scholar]

1653. Xia, F.; Zhu, Y.; Feng, L.; Jiang, L. Smart responsive surfaces switching reversibly between super-
hydrophobicity and super-hydrophilicity Soft Matter 2009, 5, 275– 281 DOI: 10.1039/B803951H [Crossref],
[CAS], [Google Scholar]

1654. Luzinov, I.; Minko, S.; Tsukruk, V. V. Adaptive and responsive surfaces through controlled reorganization of
interfacial polymer layers Prog. Polym. Sci. 2004, 29, 635– 698 DOI: 10.1016/j.progpolymsci.2004.03.001
[Crossref], [CAS], [Google Scholar]

1655. Huber, D. L.; Manginell, R. P.; Samara, M. A.; Kim, B.-I.; Bunker, B. C. Programmed Adsorption and Release
of Proteins in a Micro uidic Device Science 2003, 301, 352– 354 DOI: 10.1126/science.1080759 [Crossref],
[PubMed], [CAS], [Google Scholar]

1656. Rama Rao, G. V.; López, G. P. Encapsulation of Poly(N-Isopropyl Acrylamide) in Silica: A Stimuli-
Responsive Porous Hybrid Material That Incorporates Molecular Nano-Valves Adv. Mater. 2000, 12, 1692– 1695
 DOI: 10.1002/1521-4095(200011)12:22<1692::AID-ADMA1692>3.0.CO;2-C [Crossref], [Google Scholar]

  
1657. Rao, G. V. R.; Balamurugan, S.; Meyer, D. E.; Chilkoti, A.; López, G. P. Hybrid Bioinorganic Smart
Membranes That Incorporate Protein-Based Molecular Switches Langmuir 2002, 18, 1819– 1824
 DOI: 10.1021/la011188i [ACS Full Text ], [CAS], [Google Scholar]

1658. Rama Rao, G. V.; Krug, M. E.; Balamurugan, S.; Xu, H.; Xu, Q.; López, G. P. Synthesis and Characterization
of Silica–Poly(N-isopropylacrylamide) Hybrid Membranes: Switchable Molecular Filters Chem. Mater. 2002, 14,
5075– 5080 DOI: 10.1021/cm020627b [ACS Full Text ], [Google Scholar]

1659. Liu; Dunphy, D. R.; Atanassov, P.; Bunge, S. D.; Chen, Z.; López, G. P.; Boyle, T. J.; Brinker, C. J.
Photoregulation of Mass Transport through a Photoresponsive Azobenzene-Modi ed Nanoporous Membrane
Nano Lett. 2004, 4, 551– 554 DOI: 10.1021/nl0350783 [ACS Full Text ], [CAS], [Google Scholar]

1660. Casasús, R.; Marcos, M. D.; Martínez-Máñez, R.; Ros-Lis, J. V.; Soto, J.; Villaescusa, L. A.; Amorós, P.;
Beltrán, D.; Guillem, C.; Latorre, J. Toward the Development of Ionically Controlled Nanoscopic Molecular Gates J.
Am. Chem. Soc. 2004, 126, 8612– 8613 DOI: 10.1021/ja048095i [ACS Full Text ], [CAS], [Google Scholar]

1661. Pardo-Yissar, V.; Gabai, R.; Shipway, A. N.; Bourenko, T.; Willner, I. Gold Nanoparticle/Hydrogel Composites
with Solvent-Switchable Electronic Properties Adv. Mater. 2001, 13, 1320– 1323
 DOI: 10.1002/1521-4095(200109)13:17<1320::AID-ADMA1320>3.0.CO;2-8 [Crossref], [CAS], [Google Scholar]

1662. Huang, F. H.; Scherman, O. A. Supramolecular polymers Chem. Soc. Rev. 2012, 41, 5879– 5880
 DOI: 10.1039/c2cs90071h [Crossref], [PubMed], [CAS], [Google Scholar]

1663. Appel, E. A.; del Barrio, J.; Loh, X. J.; Scherman, O. A. Supramolecular polymeric hydrogels Chem. Soc.
Rev. 2012, 41, 6195– 6214 DOI: 10.1039/c2cs35264h [Crossref], [PubMed], [CAS], [Google Scholar]

1664. Nguyen, T. D.; Tseng, H.-R.; Celestre, P. C.; Flood, A. H.; Liu, Y.; Stoddart, J. F.; Zink, J. I. A reversible
molecular valve Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 10029– 10034 DOI: 10.1073/pnas.0504109102
[Crossref], [PubMed], [CAS], [Google Scholar]

1665. Klajn, R.; Olson, M. A.; Wesson, P. J.; Fang, L.; Coskun, A.; Trabolsi, A.; Soh, S.; Stoddart, J. F.; Grzybowski,
B. A. Dynamic hook-and-eye nanoparticle sponges Nat. Chem. 2009, 1, 733– 738 DOI: 10.1038/nchem.432
[Crossref], [PubMed], [CAS], [Google Scholar]
1666. Shipway, A. N.; Willner, I. Electronically Transduced Molecular Mechanical and Information Functions on
Surfaces Acc. Chem. Res. 2001, 34, 421– 432 DOI: 10.1021/ar000180h [ACS Full Text  
], [CAS], [Google 
Scholar]

1667. Pardo-Yissar, V.; Katz, E.; Willner, I.; Kotlyar, A. B.; Sanders, C.; Lill, H. Biomaterial engineered electrodes for
bioelectronics Faraday Discuss. 2000, 116, 119– 134 DOI: 10.1039/b001508n [Crossref], [PubMed],
[CAS], [Google Scholar]

1668. Willner, I.; Willner, B. Layered molecular optoelectronic assemblies J. Mater. Chem. 1998, 8, 2543– 2556
 DOI: 10.1039/a804135k [Crossref], [CAS], [Google Scholar]

1669. Willner, I.; Pardo-Yissar, V.; Katz, E.; Ranjit, K. T. A photoactivated ‘molecular train’ for optoelectronic
applications: light-stimulated translocation of a β-cyclodextrin receptor within a stoppered azobenzene-alkyl
chain supramolecular monolayer assembly on a Au-electrode J. ElectroAnal. Chem. 2001, 497, 172– 177
 DOI: 10.1016/S0022-0728(00)00455-1 [Crossref], [CAS], [Google Scholar]

1670. Katz, E.; Sheeney-Haj-Ichia, L.; Willner, I. Electrical Contacting of Glucose Oxidase in a Redox-Active
Rotaxane Con guration Angew. Chem., Int. Ed. 2004, 43, 3292– 3300 DOI: 10.1002/anie.200353455 [Crossref],
[PubMed], [CAS], [Google Scholar]

1671. Sheeney-Haj-Ichia, L.; Willner, I. Enhanced Photoelectrochemistry in Supramolecular CdS-Nanoparticle-


Stoppered Pseudorotaxane Monolayers Assembled on Electrodes J. Phys. Chem. B 2002, 106, 13094– 13097
 DOI: 10.1021/jp022102c [ACS Full Text ], [CAS], [Google Scholar]

1672. Katz, E.; Lioubashevsky, O.; Willner, I. Electromechanics of a Redox-Active Rotaxane in a Monolayer
Assembly on an Electrode J. Am. Chem. Soc. 2004, 126, 15520– 15532 DOI: 10.1021/ja045465u [ACS Full Text
], [CAS], [Google Scholar]

1673. Blossey, R. Self-cleaning surfaces - virtual realities Nat. Mater. 2003, 2, 301– 306 DOI: 10.1038/nmat856
[Crossref], [PubMed], [CAS], [Google Scholar]

1674. Xin, B.; Hao, J. Reversibly switchable wettability Chem. Soc. Rev. 2010, 39, 769– 782
 DOI: 10.1039/B913622C [Crossref], [PubMed], [CAS], [Google Scholar]

1675. Hua, Z.; Yang, J.; Wang, T.; Liu, G.; Zhang, G. Transparent Surface with Reversibly Switchable Wettability
between Superhydrophobicity and Superhydrophilicity Langmuir 2013, 29, 10307– 10312
 DOI: 10.1021/la402584v [ACS Full Text ], [CAS], [Google Scholar]
 
Langmuir
1676. Brochard, F. Motions of droplets on solid surfaces induced by chemical or thermal gradients
1989, 5, 432– 438 DOI: 10.1021/la00086a025 [ACS Full Text ], [CAS], [Google Scholar]

1677. Chaudhury, M. K.; Whitesides, G. M. How to Make Water Run Uphill Science 1992, 256, 1539– 1541
 DOI: 10.1126/science.256.5063.1539 [Crossref], [PubMed], [CAS], [Google Scholar]

1678. Daniel, S.; Chaudhury, M. K.; Chen, J. C. Fast Drop Movements Resulting from the Phase Change on a
Gradient Surface Science 2001, 291, 633– 636 DOI: 10.1126/science.291.5504.633 [Crossref], [PubMed],
[CAS], [Google Scholar]

1679. Daniel, S.; Sircar, S.; Gliem, J.; Chaudhury, M. K. Ratcheting Motion of Liquid Drops on Gradient Surfaces†
Langmuir 2004, 20, 4085– 4092 DOI: 10.1021/la036221a [ACS Full Text ], [CAS], [Google Scholar]

1680. Liu, Y.; Mu, L.; Liu, B.; Kong, J. Controlled Switchable Surface Chem. - Eur. J. 2005, 11, 2622– 2631
 DOI: 10.1002/chem.200400931 [Crossref], [PubMed], [CAS], [Google Scholar]

1681. Lahann, J.; Mitragotri, S.; Tran, T.-N.; Kaido, H.; Sundaram, J.; Choi, I. S.; Hoffer, S.; Somorjai, G. A.; Langer,
R. A Reversibly Switching Surface Science 2003, 299, 371– 374 DOI: 10.1126/science.1078933 [Crossref],
[PubMed], [CAS], [Google Scholar]

1682. Wang, X.; Kharitonov, A. B.; Katz, E.; Willner, I. Potential-controlled molecular machinery of bipyridinium
monolayer-functionalized surfaces: an electrochemical and contact angle analysis Chem. Commun. 2003,
1542– 1543 DOI: 10.1039/b303845a [Crossref], [CAS], [Google Scholar]

1683. Wan, P.; Jiang, Y.; Wang, Y.; Wang, Z.; Zhang, X. Tuning surface wettability through photocontrolled
reversible molecular shuttle Chem. Commun. 2008, 5710– 5712 DOI: 10.1039/b811729b [Crossref], [PubMed],
[CAS], [Google Scholar]

1684. Zhang, X.; Zhao, H.; Tian, D.; Deng, H.; Li, H. A Photoresponsive Wettability Switch Based on a
Dimethylamino Calix[4]arene Chem. - Eur. J. 2014, 20, 9367– 9371 DOI: 10.1002/chem.201402476 [Crossref],
[PubMed], [CAS], [Google Scholar]

1685. Chen, K.-Y.; Ivashenko, O.; Carroll, G. T.; Robertus, J.; Kistemaker, J. C. M.; London, G.; Browne, W. R.;
Rudolf, P.; Feringa, B. L. Control of Surface Wettability Using Tripodal Light-Activated Molecular Motors J. Am.
Chem. Soc. 2014, 136, 3219– 3224 DOI: 10.1021/ja412110t [ACS Full Text ], [CAS], [Google Scholar]
1686. Gau, H.; Herminghaus, S.; Lenz, P.; Lipowsky, R. Liquid Morphologies on Structured Surfaces:
From 
Microchannels to Microchips Science 1999, 283, 46– 49 DOI: 10.1126/science.283.5398.46 [Crossref],
[PubMed], [CAS], [Google Scholar]

1687. Gallardo, B. S.; Gupta, V. K.; Eagerton, F. D.; Jong, L. I.; Craig, V. S.; Shah, R. R.; Abbott, N. L.
Electrochemical Principles for Active Control of Liquids on Submillimeter Scales Science 1999, 283, 57– 60
 DOI: 10.1126/science.283.5398.57 [Crossref], [PubMed], [CAS], [Google Scholar]

1688. Daniel, S.; Chaudhury, M. K. Recti ed Motion of Liquid Drops on Gradient Surfaces Induced by Vibration
Langmuir 2002, 18, 3404– 3407 DOI: 10.1021/la025505c [ACS Full Text ], [CAS], [Google Scholar]

1689. Suda, H.; Yamada, S. Force Measurements for the Movement of a Water Drop on a Surface with a
Surface Tension Gradient Langmuir 2003, 19, 529– 531 DOI: 10.1021/la0264163 [ACS Full Text ],
[CAS], [Google Scholar]

1690. Choi, S.-H.; Zhang Newby, B.-m. Micrometer-Scaled Gradient Surfaces Generated Using Contact Printing
of Octadecyltrichlorosilane Langmuir 2003, 19, 7427– 7435 DOI: 10.1021/la035027l [ACS Full Text ],
[CAS], [Google Scholar]

1691. Grunze, M. Driven Liquids Science 1999, 283, 41– 42 DOI: 10.1126/science.283.5398.41 [Crossref],
[CAS], [Google Scholar]

1692. Reyes, D. R.; Iossi dis, D.; Auroux, P.-A.; Manz, A. Micro Total Analysis Systems. 1. Introduction, Theory,
and Technology Anal. Chem. 2002, 74, 2623– 2636 DOI: 10.1021/ac0202435 [ACS Full Text ], [CAS], [Google
Scholar]

1693. Ichimura, K.; Oh, S.-K.; Nakagawa, M. Light-Driven Motion of Liquids on a Photoresponsive Surface
Science 2000, 288, 1624– 1626 DOI: 10.1126/science.288.5471.1624 [Crossref], [PubMed], [CAS], [Google
Scholar]

1694. Minko, S.; Müller, M.; Motornov, M.; Nitschke, M.; Grundke, K.; Stamm, M. Two-Level Structured Self-
Adaptive Surfaces with Reversibly Tunable Properties J. Am. Chem. Soc. 2003, 125, 3896– 3900
 DOI: 10.1021/ja0279693 [ACS Full Text ], [CAS], [Google Scholar]

1695. Sun, T.; Wang, G.; Feng, L.; Liu, B.; Ma, Y.; Jiang, L.; Zhu, D. Reversible Switching between
Superhydrophilicity and Superhydrophobicity Angew. Chem., Int. Ed. 2004, 43, 357– 360
 DOI: 10.1002/anie.200352565 [Crossref], [PubMed], [CAS], [Google Scholar]

  
1696. Fu, Q.; Rama Rao, G. V.; Basame, S. B.; Keller, D. J.; Artyushkova, K.; Fulghum, J. E.; López, G. P. Reversible
Control of Free Energy and Topography of Nanostructured Surfaces J. Am. Chem. Soc. 2004, 126, 8904– 8905
 DOI: 10.1021/ja047895q [ACS Full Text ], [CAS], [Google Scholar]

1697. Rosario, R.; Gust, D.; Garcia, A. A.; Hayes, M.; Taraci, J. L.; Clement, T.; Dailey, J. W.; Picraux, S. T. Lotus
Effect Ampli es Light-Induced Contact Angle Switching J. Phys. Chem. B 2004, 108, 12640– 12642
 DOI: 10.1021/jp0473568 [ACS Full Text ], [CAS], [Google Scholar]

1698. Ariga, K.; Ji, Q.; Mori, T.; Naito, M.; Yamauchi, Y.; Abe, H.; Hill, J. P. Enzyme nanoarchitectonics:
organization and device application Chem. Soc. Rev. 2013, 42, 6322– 6345 DOI: 10.1039/c2cs35475f [Crossref],
[PubMed], [CAS], [Google Scholar]

1699. Teller, C.; Willner, I. Functional nucleic acid nanostructures and DNA machines Curr. Opin. Biotechnol.
2010, 21, 376– 391 DOI: 10.1016/j.copbio.2010.06.001 [Crossref], [PubMed], [CAS], [Google Scholar]

1700. Hirokawa, N.; Noda, Y.; Tanaka, Y.; Niwa, S. Kinesin superfamily motor proteins and intracellular transport
Nat. Rev. Mol. Cell Biol. 2009, 10, 682– 696 DOI: 10.1038/nrm2774 [Crossref], [PubMed], [CAS], [Google Scholar]

1701. Montemagno, C.; Bachand, G. Constructing nanomechanical devices powered by biomolecular motors
Nanotechnology 1999, 10, 225– 231 DOI: 10.1088/0957-4484/10/3/301 [Crossref], [CAS], [Google Scholar]

1702. Hess, H.; Vogel, V. Molecular shuttles based on motor proteins: active transport in synthetic
environments Rev. Mol. Biotechnol. 2001, 82, 67– 85 DOI: 10.1016/S1389-0352(01)00029-0 [Crossref],
[CAS], [Google Scholar]

1703. Adachi, K.; Yasuda, R.; Noji, H.; Itoh, H.; Harada, Y.; Yoshida, M.; Kinosita, K. Stepping rotation of F-1-
ATPase visualized through angle-resolved single- uorophore imaging Proc. Natl. Acad. Sci. U. S. A. 2000, 97,
7243– 7247 DOI: 10.1073/pnas.120174297 [Crossref], [PubMed], [CAS], [Google Scholar]

1704. Yasuda, R.; Noji, H.; Yoshida, M.; Kinosita, K.; Itoh, H. Resolution of distinct rotational substeps by
submillisecond kinetic analysis of F-1-ATPase Nature 2001, 410, 898– 904 DOI: 10.1038/35073513 [Crossref],
[PubMed], [CAS], [Google Scholar]
1705. Hirono-Hara, Y.; Noji, H.; Nishiura, M.; Muneyuki, E.; Hara, K. Y.; Yasuda, R.; Kinosita, K., Jr.; Yoshida, M.
13654
Pause and rotation of F(1)-ATPase during catalysis Proc. Natl. Acad. Sci. U. S. A. 2001, 98, 13649– 
 DOI: 10.1073/pnas.241365698 [Crossref], [PubMed], [CAS], [Google Scholar]

1706. Noji, H.; Yasuda, R.; Yoshida, M.; Kinosita, K., Jr. Direct observation of the rotation of F1-ATPase Nature
1997, 386, 299– 302 DOI: 10.1038/386299a0 [Crossref], [PubMed], [CAS], [Google Scholar]

1707. Yasuda, R.; Noji, H.; Kinosita, K.; Yoshida, M. F1-ATPase is a highly e cient molecular motor that rotates
with discrete 120 degree steps Cell 1998, 93, 1117– 1124 DOI: 10.1016/S0092-8674(00)81456-7 [Crossref],
[PubMed], [CAS], [Google Scholar]

1708. Diez, M.; Zimmermann, B.; Borsch, M.; Konig, M.; Schweinberger, E.; Steigmiller, S.; Reuter, R.; Felekyan, S.;
Kudryavtsev, V.; Seidel, C. A. M.; Graber, P. Proton-powered subunit rotation in single membrane-bound F0F1-ATP
synthase Nat. Struct. Mol. Biol. 2004, 11, 135– 141 DOI: 10.1038/nsmb718 [Crossref], [PubMed], [CAS], [Google
Scholar]

1709. Itoh, H.; Takahashi, A.; Adachi, K.; Noji, H.; Yasuda, R.; Yoshida, M.; Kinosita, K. Mechanically driven ATP
synthesis by F1-ATPase Nature 2004, 427, 465– 468 DOI: 10.1038/nature02212 [Crossref], [PubMed],
[CAS], [Google Scholar]

1710. Soong, R. K.; Bachand, G. D.; Neves, H. P.; Olkhovets, A. G.; Craighead, H. G.; Montemagno, C. D. Powering
an inorganic nanodevice with a biomolecular motor Science 2000, 290, 1555– 1558
 DOI: 10.1126/science.290.5496.1555 [Crossref], [PubMed], [CAS], [Google Scholar]

1711. Duan, L.; He, Q.; Wang, K. W.; Yan, X. H.; Cui, Y.; Moewald, H.; Li, J. B. Adenosine triphosphate biosynthesis
catalyzed by F0F1 ATP synthase assembled in polymer microcapsules Angew. Chem., Int. Ed. 2007, 46, 6996–
7000 DOI: 10.1002/anie.200700331 [Crossref], [PubMed], [CAS], [Google Scholar]

1712. Oiwa, K.; Chaen, S.; Kamitsubo, E.; Shimmen, T.; Sugi, H. Steady-state force-velocity relation in the ATP-
dependent sliding movement of myosin-coated beads on actin cables in vitro studied with a centrifuge
microscope Proc. Natl. Acad. Sci. U. S. A. 1990, 87, 7893– 7897 DOI: 10.1073/pnas.87.20.7893 [Crossref],
[PubMed], [CAS], [Google Scholar]

1713. Reuther, C.; Hajdo, L.; Tucker, R.; Kasprzak, A. A.; Diez, S. Biotemplated nanopatterning of planar surfaces
with molecular motors Nano Lett. 2006, 6, 2177– 2183 DOI: 10.1021/nl060922l [ACS Full Text ], [CAS], [Google
Scholar]
1714. Ramachandran, S.; Ernst, K. H.; Bachand, G. D.; Vogel, V.; Hess, H. Selective loading of kinesin-powered
molecular shuttles with protein cargo and its application to biosensing Small 2006, 2, 330– 334   
 DOI: 10.1002/smll.200500265 [Crossref], [PubMed], [CAS], [Google Scholar]

1715. Diez, S.; Reuther, C.; Dinu, C.; Seidel, R.; Mertig, M.; Pompe, W.; Howard, J. Stretching and transporting
DNA molecules using motor proteins Nano Lett. 2003, 3, 1251– 1254 DOI: 10.1021/nl034504h [ACS Full Text ],
[CAS], [Google Scholar]

1716. Dumont, E. L. P.; Do, C.; Hess, H. Molecular wear of microtubules propeled by surface-adhered kinesins
Nat. Nanotechnol. 2015, 10, 166– 169 DOI: 10.1038/nnano.2014.334 [Crossref], [PubMed], [CAS], [Google
Scholar]

1717. Yu, T. Y.; Wang, Q.; Johnson, D. S.; Wang, M. D.; Ober, C. K. Functional hydrogel surfaces: Binding kinesin-
based molecular motor proteins to selected patterned sites Adv. Funct. Mater. 2005, 15, 1303– 1309
 DOI: 10.1002/adfm.200400117 [Crossref], [CAS], [Google Scholar]

1718. Yildiz, A.; Selvin, P. R. Kinesin: walking, crawling or sliding along? Trends Cell Biol. 2005, 15, 112– 120
 DOI: 10.1016/j.tcb.2004.12.007 [Crossref], [PubMed], [CAS], [Google Scholar]

1719. Beeg, J.; Klumpp, S.; Dimova, R.; Gracia, R. S.; Unger, E.; Lipowsky, R. Transport of beads by several
kinesin motors Biophys. J. 2008, 94, 532– 541 DOI: 10.1529/biophysj.106.097881 [Crossref], [PubMed],
[CAS], [Google Scholar]

1720. Yamasaki, H.; Nakayama, H. Fluctuation analysis of myosin-coated bead movement along actin bundles
of Nitella Biochem. Biophys. Res. Commun. 1996, 221, 831– 836 DOI: 10.1006/bbrc.1996.0682 [Crossref],
[PubMed], [CAS], [Google Scholar]

1721. Spudich, J. A.; Kron, S. J.; Sheetz, M. P. Movement of Myosin-Coated Beads on Oriented Filaments
Reconstituted from Puri ed Actin Nature 1985, 315, 584– 586 DOI: 10.1038/315584a0 [Crossref], [PubMed],
[CAS], [Google Scholar]

1722. Sheetz, M. P.; Spudich, J. A. Movement of myosin-coated uorescent beads on actin cables in vitro
Nature 1983, 303, 31– 35 DOI: 10.1038/303031a0 [Crossref], [PubMed], [CAS], [Google Scholar]

1723. Yokokawa, R.; Takeuchi, S.; Kon, T.; Nishiura, M.; Sutoh, K.; Fujita, H. Unidirectional transport of kinesin-
coated beads on microtubules oriented in a micro uidic device Nano Lett. 2004, 4, 2265– 2270
 DOI: 10.1021/nl048851i [ACS Full Text ], [CAS], [Google Scholar]
1724. Bormuth, V.; Varga, V.; Howard, J.; Schaffer, E. Protein Friction Limits Diffusive and Directed  
Movements
of Kinesin Motors on Microtubules Science 2009, 325, 870– 873 DOI: 10.1126/science.1174923 [Crossref],
[PubMed], [CAS], [Google Scholar]

1725. Agarwal, A.; Katira, P.; Hess, H. Millisecond Curing Time of a Molecular Adhesive Causes Velocity-
Dependent Cargo-Loading of Molecular Shuttles Nano Lett. 2009, 9, 1170– 1175 DOI: 10.1021/nl803831y [ACS
Full Text ], [CAS], [Google Scholar]

1726. Bohm, K. J.; Stracke, R.; Muhlig, P.; Unger, E. Motor protein-driven unidirectional transport of micrometer-
sized cargoes across isopolar microtubule arrays Nanotechnology 2001, 12, 238– 244
 DOI: 10.1088/0957-4484/12/3/307 [Crossref], [CAS], [Google Scholar]

1727. Bottier, C.; Fattaccioli, J.; Tarhan, M. C.; Yokokawa, R.; Morin, F. O.; Kim, B.; Collard, D.; Fujita, H. Active
transport of oil droplets along oriented microtubules by kinesin molecular motors Lab Chip 2009, 9, 1694– 1700
 DOI: 10.1039/b822519b [Crossref], [PubMed], [CAS], [Google Scholar]

1728. Nitta, T.; Tanahashi, A.; Obara, Y.; Hirano, M.; Razumova, M.; Regnier, M.; Hess, H. Comparing guiding track
requirements for myosin- and kinesin-powered molecular shuttles Nano Lett. 2008, 8, 2305– 2309
 DOI: 10.1021/nl8010885 [ACS Full Text ], [CAS], [Google Scholar]

1729. Vikhorev, P. G.; Vikhoreva, N. N.; Mansson, A. Bending exibility of actin laments during motor-induced
sliding Biophys. J. 2008, 95, 5809– 5819 DOI: 10.1529/biophysj.108.140335 [Crossref], [PubMed], [CAS], [Google
Scholar]

1730. Vikhorev, P. G.; Vikhoreva, N. N.; Sundberg, M.; Balaz, M.; Albet-Torres, N.; Bunk, R.; Kvennefors, A.;
Liljesson, K.; Nicholls, I. A.; Nilsson, L.; Omling, P.; Tagerud, S.; Montelius, L.; Mansson, A. Diffusion dynamics of
motor-driven transport: gradient production and self-organization of surfaces Langmuir 2008, 24, 13509– 13517
 DOI: 10.1021/la8016112 [ACS Full Text ], [CAS], [Google Scholar]

1731. Campbell, J.; Paul, D.; Kurabayashi, K.; Meyhofer, E. A Kinesin Driven Enzyme Linked Immunosorbant
Assay (ELISA) for Ultra Low Protein Detection Applications Biophys. J. 2014, 106, 622a– 622a
 DOI: 10.1016/j.bpj.2013.11.3441 [Crossref], [Google Scholar]

1732. Kakugo, A.; Sugimoto, S.; Gong, J. P.; Osada, Y. Gel machines constructed from chemically cross-linked
actins and myosins Adv. Mater. 2002, 14, 1124– 1126
 DOI: 10.1002/1521-4095(20020816)14:16<1124::AID-ADMA1124>3.0.CO;2-M [Crossref], [CAS], [Google Scholar]
  
1733. Sanchez, T.; Chen, D. T. N.; DeCamp, S. J.; Heymann, M.; Dogic, Z. Spontaneous motion in hierarchically
assembled active matter Nature 2012, 491, 431– 434 DOI: 10.1038/nature11591 [Crossref], [PubMed],
[CAS], [Google Scholar]

1734. Marchetti, M. C. Active Matter Spontaneous ows and self-propeled drops Nature 2012, 491, 340– 341
 DOI: 10.1038/nature11750 [Crossref], [PubMed], [CAS], [Google Scholar]

1735. Kinosita, K.; Yasuda, R.; Noji, H.; Adachi, K. A rotary molecular motor that can work at near 100%
e ciency Philos. Trans. R. Soc., B 2000, 355, 473– 489 DOI: 10.1098/rstb.2000.0589 [Crossref], [PubMed],
[CAS], [Google Scholar]

1736. Boyer, P. D. The ATP synthase - A splendid molecular machine Annu. Rev. Biochem. 1997, 66, 717– 749
 DOI: 10.1146/annurev.biochem.66.1.717 [Crossref], [PubMed], [CAS], [Google Scholar]

1737. Yoshida, M.; Muneyuki, E.; Hisabori, T. ATP synthase - A marvellous rotary engine of the cell Nat. Rev. Mol.
Cell Biol. 2001, 2, 669– 677 DOI: 10.1038/35089509 [Crossref], [PubMed], [CAS], [Google Scholar]

1738. Howard, J. The movement of kinesin along microtubules Annu. Rev. Physiol. 1996, 58, 703– 729
 DOI: 10.1146/annurev.ph.58.030196.003415 [Crossref], [PubMed], [CAS], [Google Scholar]

1739. Oesterhelt, D.; Brauchle, C.; Hampp, N. Bacteriorhodopsin - a Biological-Material for Information-
Processing Q. Rev. Biophys. 1991, 24, 425– 478 DOI: 10.1017/S0033583500003863 [Crossref], [PubMed],
[CAS], [Google Scholar]

1740. Schick, G. A.; Lawrence, A. F.; Birge, R. R. Biotechnology and Molecular Computing Trends Biotechnol.
1988, 6, 159– 163 DOI: 10.1016/0167-7799(88)90086-8 [Crossref], [CAS], [Google Scholar]

1741. Birge, R. R. Photophysics and Molecular Electronic Applications of the Rhodopsins Annu. Rev. Phys.
Chem. 1990, 41, 683– 733 DOI: 10.1146/annurev.pc.41.100190.003343 [Crossref], [PubMed], [CAS], [Google
Scholar]

1742. Chen, Z. P.; Birge, R. R. Protein-Based Arti cial Retinas Trends Biotechnol. 1993, 11, 292– 300
 DOI: 10.1016/0167-7799(93)90017-4 [Crossref], [PubMed], [CAS], [Google Scholar]
1743. Hampp, N.; Silber, A. Functional dyes from nature: Potentials for technical applications Pure Appl. Chem.
1996, 68, 1361– 1366 DOI: 10.1351/pac199668071361 [Crossref], [CAS], [Google Scholar]   

1744. Birge, R. R.; Gillespie, N. B.; Izaguirre, E. W.; Kusnetzow, A.; Lawrence, A. F.; Singh, D.; Song, Q. W.; Schmidt,
E.; Stuart, J. A.; Seetharaman, S.; Wise, K. J. Biomolecular electronics: Protein-based associative processors and
volumetric memories J. Phys. Chem. B 1999, 103, 10746– 10766 DOI: 10.1021/jp991883n [ACS Full Text ],
[CAS], [Google Scholar]

1745. Hampp, N. Bacteriorhodopsin as a photochromic retinal protein for optical memories Chem. Rev. 2000,
100, 1755– 1776 DOI: 10.1021/cr980072x [ACS Full Text ], [CAS], [Google Scholar]

1746. Masthay, M. B.; Sammeth, D. M.; Helvenston, M. C.; Buckman, C. B.; Li, W. Y.; Cde-Baca, M. J.; Kofron, J. T.
The laser-induced blue state of bacteriorhodopsin: Mechanistic and color regulatory roles of protein-protein
interactions, protein-lipid interactions, and metal ions J. Am. Chem. Soc. 2002, 124, 3418– 3430
 DOI: 10.1021/ja010116a [ACS Full Text ], [CAS], [Google Scholar]

1747. Einati, H.; Mishra, D.; Friedman, N.; Sheves, M.; Naaman, R. Light-Controlled Spin Filtering in
Bacteriorhodopsin Nano Lett. 2015, 15, 1052– 1056 DOI: 10.1021/nl503961p [ACS Full Text ], [CAS], [Google
Scholar]

1748. Rakovich, A.; Nabiev, I.; Sukhanova, A.; Lesnyak, V.; Gaponik, N.; Rakovich, Y. P.; Donegan, J. F. Large
Enhancement of Nonlinear Optical Response in a Hybrid Nanobiomaterial Consisting of Bacteriorhodopsin and
Cadmium Telluride Quantum Dots ACS Nano 2013, 7, 2154– 2160 DOI: 10.1021/nn3049939 [ACS Full Text ],
[CAS], [Google Scholar]

1749. Jin, Y. D.; Friedman, N.; Sheves, M.; Cahen, D. Bacteriorhodopsin-monolayer-based planar metal-insulator-
metal junctions via biomimetic vesicle fusion: Preparation, characterization, and bio-optoelectronic
characteristics Adv. Funct. Mater. 2007, 17, 1417– 1428 DOI: 10.1002/adfm.200600545 [Crossref],
[CAS], [Google Scholar]

1750. Szacilowski, K. Digital information processing in molecular systems Chem. Rev. 2008, 108, 3481– 3548
 DOI: 10.1021/cr068403q [ACS Full Text ], [CAS], [Google Scholar]

1751. Matile, S.; Som, A.; Sorde, N. Recent synthetic ion channels and pores Tetrahedron 2004, 60, 6405– 6435
 DOI: 10.1016/j.tet.2004.05.052 [Crossref], [CAS], [Google Scholar]
1752. Gokel, G. W.; Negin, S. Synthetic Ion Channels: From Pores to Biological Applications Acc. Chem. Res.
2013, 46, 2824– 2833 DOI: 10.1021/ar400026x [ACS Full Text ], [CAS], [Google Scholar]   

1753. Sakai, N.; Matile, S. Synthetic Ion Channels Langmuir 2013, 29, 9031– 9040 DOI: 10.1021/la400716c
[ACS Full Text ], [CAS], [Google Scholar]

1754. Fyles, T. M. Synthetic ion channels in bilayer membranes Chem. Soc. Rev. 2007, 36, 335– 347
 DOI: 10.1039/B603256G [Crossref], [PubMed], [CAS], [Google Scholar]

1755. Carr, R.; Weinstock, I. A.; Sivaprasadarao, A.; Muller, A.; Aksimentiev, A. Synthetic Ion Channels via Self-
Assembly: A Route for Embedding Porous Polyoxometalate Nanocapsules in Lipid Bilayer Membranes Nano Lett.
2008, 8, 3916– 3921 DOI: 10.1021/nl802366k [ACS Full Text ], [CAS], [Google Scholar]

1756. Branton, D.; Deamer, D. W.; Marziali, A.; Bayley, H.; Benner, S. A.; Butler, T.; Di Ventra, M.; Garaj, S.; Hibbs, A.;
Huang, X.; Jovanovich, S. B.; Krstic, P. S.; Lindsay, S.; Ling, X. S.; Mastrangelo, C. H.; Meller, A.; Oliver, J. S.; Pershin,
Y. V.; Ramsey, J. M.; Riehn, R.; Soni, G. V.; Tabard-Cossa, V.; Wanunu, M.; Wiggin, M.; Schloss, J. A. The potential
and challenges of nanopore sequencing Nat. Biotechnol. 2008, 26, 1146– 1153 DOI: 10.1038/nbt.1495
[Crossref], [PubMed], [CAS], [Google Scholar]

1757. Kang, X.-f.; Gu, L.-Q.; Cheley, S.; Bayley, H. Single Protein Pores Containing Molecular Adapters at High
Temperatures Angew. Chem., Int. Ed. 2005, 44, 1495– 1499 DOI: 10.1002/anie.200461885 [Crossref], [PubMed],
[CAS], [Google Scholar]

1758. Meller, A.; Nivon, L.; Branton, D. Voltage-Driven DNA Translocations through a Nanopore Phys. Rev. Lett.
2001, 86, 3435– 3438 DOI: 10.1103/PhysRevLett.86.3435 [Crossref], [PubMed], [CAS], [Google Scholar]

1759. Gershow, M.; Golovchenko, J. A. Recapturing and trapping single molecules with a solid-state nanopore
Nat. Nanotechnol. 2007, 2, 775– 779 DOI: 10.1038/nnano.2007.381 [Crossref], [PubMed], [CAS], [Google Scholar]

1760. Lubensky, D. K.; Nelson, D. R. Driven Polymer Translocation Through a Narrow Pore Biophys. J. 1999, 77,
1824– 1838 DOI: 10.1016/S0006-3495(99)77027-X [Crossref], [PubMed], [CAS], [Google Scholar]

1761. Holden, M. A.; Bayley, H. Direct Introduction of Single Protein Channels and Pores into Lipid Bilayers J.
Am. Chem. Soc. 2005, 127, 6502– 6503 DOI: 10.1021/ja042470p [ACS Full Text ], [CAS], [Google Scholar]
1762. González, J. E.; Oades, K.; Leychkis, Y.; Harootunian, A.; Negulescu, P. A. Cell-based assays and
instrumentation for screening ion-channel targets Drug Discovery Today 1999, 4, 431– 439   
 DOI: 10.1016/S1359-6446(99)01383-5 [Crossref], [PubMed], [CAS], [Google Scholar]

1763. Jentsch, T. J.; Hubner, C. A.; Fuhrmann, J. C. Ion channels: Function unravelled by dysfunction Nat. Cell
Biol. 2004, 6, 1039– 1047 DOI: 10.1038/ncb1104-1039 [Crossref], [PubMed], [CAS], [Google Scholar]

1764. Majd, S.; Yusko, E. C.; Billeh, Y. N.; Macrae, M. X.; Yang, J.; Mayer, M. Applications of biological pores in
nanomedicine, sensing, and nanoelectronics Curr. Opin. Biotechnol. 2010, 21, 439– 476
 DOI: 10.1016/j.copbio.2010.05.002 [Crossref], [PubMed], [CAS], [Google Scholar]

1765. Koçer, A.; Walko, M.; Bulten, E.; Halza, E.; Feringa, B. L.; Meijberg, W. Rationally Designed Chemical
Modulators Convert a Bacterial Channel Protein into a pH-Sensory Valve Angew. Chem., Int. Ed. 2006, 45, 3126–
3130 DOI: 10.1002/anie.200503403 [Crossref], [PubMed], [CAS], [Google Scholar]

1766. Kiwada, T.; Sonomura, K.; Sugiura, Y.; Asami, K.; Futaki, S. Transmission of Extramembrane
Conformational Change into Current: Construction of Metal-Gated Ion Channel J. Am. Chem. Soc. 2006, 128,
6010– 6011 DOI: 10.1021/ja060515b [ACS Full Text ], [CAS], [Google Scholar]

1767. Sukharev, S.; Anishkin, A. Mechanosensitive channels: what can we learn from ‘simple’ model systems?
Trends Neurosci. 2004, 27, 345– 351 DOI: 10.1016/j.tins.2004.04.006 [Crossref], [PubMed], [CAS], [Google
Scholar]

1768. Louhivuori, M.; Risselada, H. J.; van der Giessen, E.; Marrink, S. J. Release of content through mechano-
sensitive gates in pressurized liposomes Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 19856– 19860
 DOI: 10.1073/pnas.1001316107 [Crossref], [PubMed], [CAS], [Google Scholar]

1769. Steinberg-Yfrach, G.; Rigaud, J.-L.; Durantini, E. N.; Moore, A. L.; Gust, D.; Moore, T. A. Light-driven
production of ATP catalysed by F0F1-ATP synthase in an arti cial photosynthetic membrane Nature 1998, 392,
479– 482 DOI: 10.1038/33116 [Crossref], [PubMed], [CAS], [Google Scholar]

1770. Banghart, M.; Borges, K.; Isacoff, E.; Trauner, D.; Kramer, R. H. Light-activated ion channels for remote
control of neuronal ring Nat. Neurosci. 2004, 7, 1381– 1386 DOI: 10.1038/nn1356 [Crossref], [PubMed],
[CAS], [Google Scholar]

1771. Volgraf, M.; Gorostiza, P.; Numano, R.; Kramer, R. H.; Isacoff, E. Y.; Trauner, D. Allosteric control of an
ionotropic glutamate receptor with an optical switch Nat. Chem. Biol. 2006, 2, 47– 52
 DOI: 10.1038/nchembio756 [Crossref], [PubMed], [CAS], [Google Scholar]

  
1772. Stein, M.; Breit, A.; Fehrentz, T.; Gudermann, T.; Trauner, D. Optical Control of TRPV1 Channels Angew.
Chem., Int. Ed. 2013, 52, 9845– 9848 DOI: 10.1002/anie.201302530 [Crossref], [PubMed], [CAS], [Google Scholar]

1773. Schoenberger, M.; Damijonaitis, A.; Zhang, Z. N.; Nagel, D.; Trauner, D. Development of a New
Photochromic Ion Channel Blocker via Azologization of Fomocaine ACS Chem. Neurosci. 2014, 5, 514– 518
 DOI: 10.1021/cn500070w [ACS Full Text ], [CAS], [Google Scholar]

1774. Schonberger, M.; Althaus, M.; Fronius, M.; Clauss, W.; Trauner, D. Controlling epithelial sodium channels
with light using photoswitchable amilorides Nat. Chem. 2014, 6, 712– 719 DOI: 10.1038/nchem.2004 [Crossref],
[PubMed], [CAS], [Google Scholar]

1775. Kramer, R. H.; Chambers, J. J.; Trauner, D. Photochemical tools for remote control of ion channels in
excitable cells Nat. Chem. Biol. 2005, 1, 360– 365 DOI: 10.1038/nchembio750 [Crossref], [PubMed],
[CAS], [Google Scholar]

1776. Banghart, M. R.; Mourot, A.; Fortin, D. L.; Yao, J. Z.; Kramer, R. H.; Trauner, D. Photochromic Blockers of
Voltage-Gated Potassium Channels Angew. Chem., Int. Ed. 2009, 48, 9097– 9101 DOI: 10.1002/anie.200904504
[Crossref], [PubMed], [CAS], [Google Scholar]

1777. Mourot, A.; Fehrentz, T.; Le Feuvre, Y.; Smith, C. M.; Herold, C.; Dalkara, D.; Nagy, F.; Trauner, D.; Kramer, R.
H. Rapid optical control of nociception with an ion-channel photoswitch Nat. Methods 2012, 9, 396– 402
 DOI: 10.1038/nmeth.1897 [Crossref], [PubMed], [CAS], [Google Scholar]

1778. Lougheed, T.; Borisenko, V.; Hennig, T.; Ruck-Braun, K.; Woolley, G. A. Photomodulation of ionic current
through hemithioindigo-modi ed gramicidin channels Org. Biomol. Chem. 2004, 2, 2798– 2801
 DOI: 10.1039/b408485c [Crossref], [PubMed], [CAS], [Google Scholar]

1779. Banghart, M. R.; Volgraf, M.; Trauner, D. Engineering light-gated ion channels Biochemistry 2006, 45,
15129– 15141 DOI: 10.1021/bi0618058 [ACS Full Text ], [CAS], [Google Scholar]

1780. Brieke, C.; Rohrbach, F.; Gottschalk, A.; Mayer, G.; Heckel, A. Light-Controlled Tools Angew. Chem., Int. Ed.
2012, 51, 8446– 8476 DOI: 10.1002/anie.201202134 [Crossref], [PubMed], [CAS], [Google Scholar]
1781. Shinbo, T.; Kurihara, K.; Kobatake, Y.; Kamo, N. Active transport of picrate anion through organic liquid
membrane Nature 1977, 270, 277– 278 DOI: 10.1038/270277a0 [Crossref], [PubMed], [CAS], [Google 
Scholar] 

1782. Grimaldi, J. J.; Boileau, S.; Lehn, J.-M. Light-driven, carrier-mediated electron transfer across arti cial
membranes Nature 1977, 265, 229– 230 DOI: 10.1038/265229a0 [Crossref], [PubMed], [CAS], [Google Scholar]

1783. Grimaldi, J. J.; Lehn, J.-M. Transport processes in organic chemistry. 5. Multicarrier transport: coupled
transport of electrons and metal cations mediated by an electron carrier and a selective cation carrier J. Am.
Chem. Soc. 1979, 101, 1333– 1334 DOI: 10.1021/ja00499a074 [ACS Full Text ], [CAS], [Google Scholar]

1784. Anderson, S. S.; Lyle, I. G.; Paterson, R. Electron transfer across membranes using vitamin K1 and
coenzyme Q10 as carrier molecules Nature 1976, 259, 147– 148 DOI: 10.1038/259147a0 [Crossref], [PubMed],
[CAS], [Google Scholar]

1785. Steinberg-Yfrach, G.; Liddell, P. A.; Hung, S.-C.; Moore, A. L.; Gust, D.; Moore, T. A. Conversion of light
energy to proton potential in liposomes by arti cial photosynthetic reaction centers Nature 1997, 385, 239– 241
 DOI: 10.1038/385239a0 [Crossref], [CAS], [Google Scholar]

1786. Bennett, I. M.; Farfano, H. M. V.; Bogani, F.; Primak, A.; Liddell, P. A.; Otero, L.; Sereno, L.; Silber, J. J.; Moore,
A. L.; Moore, T. A.; Gust, D. Active transport of Ca2+ by an arti cial photosynthetic membrane Nature 2002, 420,
398– 401 DOI: 10.1038/nature01209 [Crossref], [PubMed], [CAS], [Google Scholar]

1787. Meyer, T. J. Chemical approaches to arti cial photosynthesis Acc. Chem. Res. 1989, 22, 163– 170
 DOI: 10.1021/ar00161a001 [ACS Full Text ], [CAS], [Google Scholar]

1788. Wasielewski, M. R. Photoinduced electron transfer in supramolecular systems for arti cial
photosynthesis Chem. Rev. 1992, 92, 435– 461 DOI: 10.1021/cr00011a005 [ACS Full Text ], [CAS], [Google
Scholar]

1789. Gust, D.; Moore, T. A.; Moore, A. L. Molecular mimicry of photosynthetic energy and electron transfer Acc.
Chem. Res. 1993, 26, 198– 205 DOI: 10.1021/ar00028a010 [ACS Full Text ], [CAS], [Google Scholar]

1790. Kurreck, H.; Huber, M. Model Reactions for Photosynthesis—Photoinduced Charge and Energy Transfer
between Covalently Linked Porphyrin and Quinone Units Angew. Chem., Int. Ed. Engl. 1995, 34, 849– 866
 DOI: 10.1002/anie.199508491 [Crossref], [CAS], [Google Scholar]
1791. Balzani, V.; Credi, A.; Venturi, M. Photoprocesses Curr. Opin. Chem. Biol. 1997, 1, 506– 513
 DOI: 10.1016/S1367-5931(97)80045-2 [Crossref], [PubMed], [CAS], [Google Scholar]   

1792. Gust, D.; Moore, T. A.; Moore, A. L. Mimicking Photosynthetic Solar Energy Transduction Acc. Chem. Res.
2001, 34, 40– 48 DOI: 10.1021/ar9801301 [ACS Full Text ], [CAS], [Google Scholar]

1793. Dürr, H.; Bossmann, S. Ruthenium Polypyridine Complexes. On the Route to Biomimetic Assemblies as
Models for the Photosynthetic Reaction Center Acc. Chem. Res. 2001, 34, 905– 917 DOI: 10.1021/ar9901220
[ACS Full Text ], [CAS], [Google Scholar]

1794. Sun, L.; Hammarstrom, L.; Akermark, B.; Styring, S. Towards arti cial photosynthesis: ruthenium-
manganese chemistry for energy production Chem. Soc. Rev. 2001, 30, 36– 49 DOI: 10.1039/a801490f
[Crossref], [CAS], [Google Scholar]

1795. Holten, D.; Bocian, D. F.; Lindsey, J. S. Probing Electronic Communication in Covalently Linked
Multiporphyrin Arrays. A Guide to the Rational Design of Molecular Photonic Devices Acc. Chem. Res. 2002, 35,
57– 69 DOI: 10.1021/ar970264z [ACS Full Text ], [CAS], [Google Scholar]

1796. Imahori, H.; Mori, Y.; Matano, Y. Nanostructured arti cial photosynthesis J. Photochem. Photobiol., C
2003, 4, 51– 83 DOI: 10.1016/S1389-5567(03)00004-2 [Crossref], [CAS], [Google Scholar]

1797. Guldi, D. M. Fullerene-porphyrin architectures; photosynthetic antenna and reaction center models Chem.
Soc. Rev. 2002, 31, 22– 36 DOI: 10.1039/b106962b [Crossref], [PubMed], [CAS], [Google Scholar]

1798. Imahori, H. Porphyrin-fullerene linked systems as arti cial photosynthetic mimics Org. Biomol. Chem.
2004, 2, 1425– 1433 DOI: 10.1039/b403024a [Crossref], [PubMed], [CAS], [Google Scholar]

1799. Alstrum-Acevedo, J. H.; Brennaman, M. K.; Meyer, T. J. Chemical Approaches to Arti cial Photosynthesis.
2 Inorg. Chem. 2005, 44, 6802– 6827 DOI: 10.1021/ic050904r [ACS Full Text ], [CAS], [Google Scholar]

1800. Paddon-Row, M. N. Investigating long-range electron-transfer processes with rigid, covalently linked
donor-(norbornylogous bridge)-acceptor systems Acc. Chem. Res. 1994, 27, 18– 25 DOI: 10.1021/ar00037a003
[ACS Full Text ], [CAS], [Google Scholar]

1801. Nicoletta, F. P.; Cupelli, D.; Formoso, P.; De Filpo, G.; Colella, V.; Gugliuzza, A. Light Responsive Polymer
Membranes: A Review Membranes 2012, 2, 134– 197 DOI: 10.3390/membranes2010134 [Crossref], [PubMed],
[CAS], [Google Scholar]

  
1802. Limburg, B.; Laisné, G.; Bouwman, E.; Bonnet, S. Enhanced Photoinduced Electron Transfer at the Surface
of Charged Lipid Bilayers Chem. - Eur. J. 2014, 20, 8965– 8972 DOI: 10.1002/chem.201402712 [Crossref],
[PubMed], [CAS], [Google Scholar]

1803. Banerji, N.; Fürstenberg, A.; Bhosale, S.; Sisson, A. L.; Sakai, N.; Matile, S.; Vauthey, E. Ultrafast
Photoinduced Charge Separation in Naphthalene Diimide Based Multichromophoric Systems in Liquid Solutions
and in a Lipid Membrane J. Phys. Chem. B 2008, 112, 8912– 8922 DOI: 10.1021/jp801276p [ACS Full Text ],
[CAS], [Google Scholar]

1804. Numata, T.; Murakami, T.; Kawashima, F.; Morone, N.; Heuser, J. E.; Takano, Y.; Ohkubo, K.; Fukuzumi, S.;
Mori, Y.; Imahori, H. Utilization of Photoinduced Charge-Separated State of Donor–Acceptor-Linked Molecules for
Regulation of Cell Membrane Potential and Ion Transport J. Am. Chem. Soc. 2012, 134, 6092– 6095
 DOI: 10.1021/ja3007275 [ACS Full Text ], [CAS], [Google Scholar]

1805. Garg, V.; Kodis, G.; Chachisvilis, M.; Hambourger, M.; Moore, A. L.; Moore, T. A.; Gust, D. Conformationally
Constrained Macrocyclic Diporphyrin–Fullerene Arti cial Photosynthetic Reaction Center J. Am. Chem. Soc.
2011, 133, 2944– 2954 DOI: 10.1021/ja1083078 [ACS Full Text ], [CAS], [Google Scholar]

1806. Rutherford, A. W.; Moore, T. A. Mimicking photosynthesis, but just the best bits Nature 2008, 453, 449–
449 DOI: 10.1038/453449b [Crossref], [PubMed], [CAS], [Google Scholar]

1807. Bhosale, S.; Sisson, A. L.; Talukdar, P.; Fürstenberg, A.; Banerji, N.; Vauthey, E.; Bollot, G.; Mareda, J.; Röger,
C.; Würthner, F.; Sakai, N.; Matile, S. Photoproduction of Proton Gradients with π-Stacked Fluorophore Scaffolds in
Lipid Bilayers Science 2006, 313, 84– 86 DOI: 10.1126/science.1126524 [Crossref], [PubMed], [CAS], [Google
Scholar]

1808. Tan, S. C.; Crouch, L. I.; Jones, M. R.; Welland, M. Generation of Alternating Current in Response to
Discontinuous Illumination by Photoelectrochemical Cells Based on Photosynthetic Proteins Angew. Chem., Int.
Ed. 2012, 51, 6667– 6671 DOI: 10.1002/anie.201200466 [Crossref], [PubMed], [CAS], [Google Scholar]

1809. Xie, X.; Crespo, G. A.; Mistlberger, G.; Bakker, E. Photocurrent generation based on a light-driven proton
pump in an arti cial liquid membrane Nat. Chem. 2014, 6, 202– 207 DOI: 10.1038/nchem.1858 [Crossref],
[PubMed], [CAS], [Google Scholar]
1810. Simmel, F. C.; Yurke, B. Using DNA to construct and power a nanoactuator Phys. Rev. E: Stat. Phys.,
 
Plasmas, Fluids, Relat. Interdiscip. Top. 2001, 63, 041913 DOI: 10.1103/PhysRevE.63.041913 [Crossref], 
[CAS], [Google Scholar]

1811. Li, J. W. J.; Tan, W. H. A single DNA molecule nanomotor Nano Lett. 2002, 2, 315– 318
 DOI: 10.1021/nl015713+ [ACS Full Text ], [CAS], [Google Scholar]

1812. Niemeyer, C. M.; Adler, M. Nanomechanical devices based on DNA Angew. Chem., Int. Ed. 2002, 41,
3779– 3783 DOI: 10.1002/1521-3773(20021018)41:20<3779::AID-ANIE3779>3.0.CO;2-F [Crossref], [PubMed],
[CAS], [Google Scholar]

1813. Simmel, F. C.; Yurke, B. A DNA-based molecular device switchable between three distinct mechanical
states Appl. Phys. Lett. 2002, 80, 883– 885 DOI: 10.1063/1.1447008 [Crossref], [CAS], [Google Scholar]

1814. Yan, H.; Zhang, X. P.; Shen, Z. Y.; Seeman, N. C. A robust DNA mechanical device controlled by
hybridization topology Nature 2002, 415, 62– 65 DOI: 10.1038/415062a [Crossref], [PubMed], [CAS], [Google
Scholar]

1815. Seeman, N. C. Biochemistry and structural DNA nanotechnology: An evolving symbiotic relationship
Biochemistry 2003, 42, 7259– 7269 DOI: 10.1021/bi030079v [ACS Full Text ], [CAS], [Google Scholar]

1816. Feng, L. P.; Park, S. H.; Reif, J. H.; Yan, H. A two-state DNA lattice switched by DNA nanoactuator Angew.
Chem., Int. Ed. 2003, 42, 4342– 4346 DOI: 10.1002/anie.200351818 [Crossref], [PubMed], [CAS], [Google Scholar]

1817. Liu, D. S.; Balasubramanian, S. A proton-fueled DNA nanomachine Angew. Chem., Int. Ed. 2003, 42,
5734– 5736 DOI: 10.1002/anie.200352402 [Crossref], [PubMed], [CAS], [Google Scholar]

1818. Chen, Y.; Lee, S. H.; Mao, C. A DNA nanomachine based on a duplex-triplex transition Angew. Chem., Int.
Ed. 2004, 43, 5335– 5338 DOI: 10.1002/anie.200460789 [Crossref], [PubMed], [CAS], [Google Scholar]

1819. Chen, Y.; Wang, M. S.; Mao, C. D. An autonomous DNA nanomotor powered by a DNA enzyme Angew.
Chem., Int. Ed. 2004, 43, 3554– 3557 DOI: 10.1002/anie.200453779 [Crossref], [PubMed], [CAS], [Google Scholar]

1820. Chen, Y.; Mao, C. D. Putting a brake on an autonomous DNA nanomotor J. Am. Chem. Soc. 2004, 126,
8626– 8627 DOI: 10.1021/ja047991r [ACS Full Text ], [CAS], [Google Scholar]
1821. Seeman, N. C. From genes to machines: DNA nanomechanical devices Trends Biochem. Sci. 2005, 30,
119– 125 DOI: 10.1016/j.tibs.2005.01.007 [Crossref], [PubMed], [CAS], [Google Scholar]   

1822. Simmel, F. C.; Dittmer, W. U. DNA nanodevices Small 2005, 1, 284– 299 DOI: 10.1002/smll.200400111
[Crossref], [PubMed], [CAS], [Google Scholar]

1823. Beissenhirtz, M. K.; Willner, I. DNA-based machines Org. Biomol. Chem. 2006, 4, 3392– 3401
 DOI: 10.1039/b607033g [Crossref], [PubMed], [CAS], [Google Scholar]

1824. Bath, J.; Turber eld, A. J. DNA nanomachines Nat. Nanotechnol. 2007, 2, 275– 284
 DOI: 10.1038/nnano.2007.104 [Crossref], [PubMed], [CAS], [Google Scholar]

1825. Krishnan, Y.; Simmel, F. C. Nucleic Acid Based Molecular Devices Angew. Chem., Int. Ed. 2011, 50, 3124–
3156 DOI: 10.1002/anie.200907223 [Crossref], [PubMed], [CAS], [Google Scholar]

1826. Liu, X. Q.; Lu, C. H.; Willner, I. Switchable Recon guration of Nucleic Acid Nanostructures by Stimuli-
Responsive DNA Machines Acc. Chem. Res. 2014, 47, 1673– 1680 DOI: 10.1021/ar400316h [ACS Full Text ],
[CAS], [Google Scholar]

1827. Wang, F.; Willner, B.; Willner, I. DNA-based machines Top. Curr. Chem. 2014, 354, 279– 338
 DOI: 10.1007/128_2013_515 [Crossref], [PubMed], [CAS], [Google Scholar]

1828. Tang, Y.; Ge, B.; Sen, D.; Yu, H. Z. Functional DNA switches: rational design and electrochemical signaling
Chem. Soc. Rev. 2014, 43, 518– 529 DOI: 10.1039/C3CS60264H [Crossref], [PubMed], [CAS], [Google Scholar]

1829. Venkataraman, S.; Dirks, R. M.; Rothemund, P. W.; Winfree, E.; Pierce, N. A. An autonomous polymerization
motor powered by DNA hybridization Nat. Nanotechnol. 2007, 2, 490– 494 DOI: 10.1038/nnano.2007.225
[Crossref], [PubMed], [CAS], [Google Scholar]

1830. Tomov, T. E.; Tsukanov, R.; Liber, M.; Masoud, R.; Plavner, N.; Nir, E. Rational design of DNA motors: fuel
optimization through single-molecule uorescence J. Am. Chem. Soc. 2013, 135, 11935– 11941
 DOI: 10.1021/ja4048416 [ACS Full Text ], [CAS], [Google Scholar]

1831. Dittmer, W. U.; Simmel, F. C. Transcriptional control of DNA-based nanomachines Nano Lett. 2004, 4,
689– 691 DOI: 10.1021/nl049784v [ACS Full Text ], [CAS], [Google Scholar]
1832. Dittmer, W. U.; Kempter, S.; Radler, J. O.; Simmel, F. C. Using gene regulation to program DNA-based
  
molecular devices Small 2005, 1, 709– 712 DOI: 10.1002/smll.200500074 [Crossref], [PubMed], [CAS], [Google
Scholar]

1833. Shin, J. S.; Pierce, N. A. A synthetic DNA walker for molecular transport J. Am. Chem. Soc. 2004, 126,
10834– 10835 DOI: 10.1021/ja047543j [ACS Full Text ], [CAS], [Google Scholar]

1834. Sherman, W. B.; Seeman, N. C. A precisely controlled DNA biped walking device. (vol 4, pg 1801, 2004)
Nano Lett. 2004, 4, 1801– 1801 DOI: 10.1021/nl048887a [ACS Full Text ], [CAS], [Google Scholar]

1835. Kelly, T. R. Molecular motors: synthetic DNA-based walkers inspired by kinesin Angew. Chem., Int. Ed.
2005, 44, 4124– 4127 DOI: 10.1002/anie.200500568 [Crossref], [PubMed], [CAS], [Google Scholar]

1836. Omabegho, T.; Sha, R.; Seeman, N. C. A bipedal DNA Brownian motor with coordinated legs Science
2009, 324, 67– 71 DOI: 10.1126/science.1170336 [Crossref], [PubMed], [CAS], [Google Scholar]

1837. Reif, J. H. The design of autonomous DNA nanomechanical devices: Walking and rolling DNA DNA
Comp. 2003, 2568, 22– 37 DOI: 10.1007/3-540-36440-4_3 [Crossref], [Google Scholar]

1838. Yin, P.; Turber eld, A. J.; Reif, J. H. Designs of autonomous unidirectional walking DNA devices DNA
Comp. 2005, 3384, 410– 425 DOI: 10.1007/11493785_36 [Crossref], [Google Scholar]

1839. Yurke, B.; Turber eld, A. J.; Mills, A. P.; Simmel, F. C.; Neumann, J. L. A DNA-fueled molecular machine
made of DNA Nature 2000, 406, 605– 608 DOI: 10.1038/35020524 [Crossref], [PubMed], [CAS], [Google Scholar]

1840. Tian, Y.; Mao, C. D. Molecular gears: A pair of DNA circles continuously rolls against each other J. Am.
Chem. Soc. 2004, 126, 11410– 11411 DOI: 10.1021/ja046507h [ACS Full Text ], [CAS], [Google Scholar]

1841. Wang, C.; Huang, Z.; Lin, Y.; Ren, J.; Qu, X. Arti cial DNA nano-spring powered by protons Adv. Mater.
2010, 22, 2792– 2798 DOI: 10.1002/adma.201000445 [Crossref], [PubMed], [CAS], [Google Scholar]

1842. Lund, K.; Manzo, A. J.; Dabby, N.; Michelotti, N.; Johnson-Buck, A.; Nangreave, J.; Taylor, S.; Pei, R.;
Stojanovic, M. N.; Walter, N. G.; Winfree, E.; Yan, H. Molecular robots guided by prescriptive landscapes Nature
2010, 465, 206– 210 DOI: 10.1038/nature09012 [Crossref], [PubMed], [CAS], [Google Scholar]
1843. Muscat, R. A.; Bath, J.; Turber eld, A. J. A programmable molecular robot Nano Lett. 2011, 11, 982– 987
 DOI: 10.1021/nl1037165 [ACS Full Text ], [CAS], [Google Scholar]   

1844. Shelley, F. J.; Wickham, M. E.; Katsuda, Y.; Hidaka, K.; Bath, J.; Sugiyama, H.; Turber eld, A. J. Nat.
Nanotechnol. 2011, 6, 166– 169 DOI: 10.1038/nnano.2010.284 [Crossref], [PubMed], [Google Scholar]

1845. Ackermann, D.; Jester, S.-S.; Famulok, M. Design Strategy for DNA Rotaxanes with a Mechanically
Reinforced PX100 Axle Angew. Chem., Int. Ed. 2012, 51, 6771– 6775 DOI: 10.1002/anie.201202816 [Crossref],
[PubMed], [CAS], [Google Scholar]

1846. Ackermann, Damian; S, T. L.; Hannam, Jeffrey S.; Purohit, Chandra S.; Heckel, Alexander; Famulok,
Michael A Double-Stranded DNA Rotaxane Nat. Nanotechnol. 2010, 5, 436– 442 DOI: 10.1038/nnano.2010.65
[Crossref], [PubMed], [CAS], [Google Scholar]

1847. Schmidt, T. L.; Heckel, A. Construction of a structurally de ned double-stranded DNA catenane Nano Lett.
2011, 11, 1739– 1742 DOI: 10.1021/nl200303m [ACS Full Text ], [CAS], [Google Scholar]

1848. Lohmann, F.; Ackermann, D.; Famulok, M. Reversible light switch for macrocycle mobility in a DNA
rotaxane J. Am. Chem. Soc. 2012, 134, 11884– 11887 DOI: 10.1021/ja3042096 [ACS Full Text ], [CAS], [Google
Scholar]

1849. Elbaz, J.; Wang, Z. G.; Wang, F.; Willner, I. Programmed dynamic topologies in DNA catenanes Angew.
Chem., Int. Ed. 2012, 51, 2349– 2353 DOI: 10.1002/anie.201107591 [Crossref], [PubMed], [CAS], [Google Scholar]

1850. Onizuka, K.; Nagatsugi, F.; Ito, Y.; Abe, H. Automatic pseudorotaxane formation targeting on nucleic acids
using a pair of reactive oligodeoxynucleotides J. Am. Chem. Soc. 2014, 136, 7201– 7204
 DOI: 10.1021/ja5018283 [ACS Full Text ], [CAS], [Google Scholar]

1851. Niemeyer, C. M. DNA as a material for Nanotechnology Angew. Chem., Int. Ed. Engl. 1997, 36, 585– 587
 DOI: 10.1002/anie.199705851 [Crossref], [CAS], [Google Scholar]

1852. Seeman, N. C. Nucleic Acid Nanostructures and Topology Angew. Chem., Int. Ed. 1998, 37, 3220– 3238
 DOI: 10.1002/(SICI)1521-3773(19981217)37:23<3220::AID-ANIE3220>3.0.CO;2-C [Crossref], [CAS], [Google
Scholar]
1853. Seeman, N. C. DNA nanotechnology: novel DNA constructions Annu. Rev. Biophys. Biomol. Struct. 1998,
27, 225– 248 DOI: 10.1146/annurev.biophys.27.1.225 [Crossref], [PubMed], [CAS], [Google Scholar]  

1854. Seeman, N. C.; Wang, H.; Yang, X.; Liu, F.; Mao, C.; Sun, W.; Wenzler, L.; Shen, Z.; Sha, R.; Yan, H., et al. New
Motifs in DNA Nanotechnology Nanotechnology 1998, 9, 257– 273 DOI: 10.1088/0957-4484/9/3/018 [Crossref],
[CAS], [Google Scholar]

1855. Yang, X.; Vologodskii, A. V.; Liu, B.; Kemper, B.; Seeman, N. C. Torsional control of double-stranded DNA
branch migration Biopolymers 1998, 45, 69– 83
 DOI: 10.1002/(SICI)1097-0282(199801)45:1<69::AID-BIP6>3.0.CO;2-X [Crossref], [PubMed], [CAS], [Google
Scholar]

1856. Mao, C.; Sun, W.; Shen, Z.; Seeman, N. C. A nanomechanical device based on the B–Z transition of DNA
Nature 1999, 397, 144– 146 DOI: 10.1038/16437 [Crossref], [PubMed], [CAS], [Google Scholar]

1857. Niemeyer, C. M. Progress in “engineering up” nanotechnology devices utilizing DNA asa construction
material Appl. Phys. A: Mater. Sci. Process. 1999, 68, 119– 124 DOI: 10.1007/s003390050865 [Crossref],
[CAS], [Google Scholar]

1858. Seeman, N. C. DNA engineering and its application to nanotechnology Trends Biotechnol. 1999, 17, 437–
443 DOI: 10.1016/S0167-7799(99)01360-8 [Crossref], [PubMed], [CAS], [Google Scholar]

1859. Niemeyer, C. M. Self-assembled nanostructures based on DNA: towards the development of


nanobiotechnology Curr. Opin. Chem. Biol. 2000, 4, 609– 618 DOI: 10.1016/S1367-5931(00)00140-X [Crossref],
[PubMed], [CAS], [Google Scholar]

1860. Seeman, N. C.; Belcher, A. M. Emulating biology: building nanostructures from the bottom up Proc. Natl.
Acad. Sci. U. S. A. 2002, 99, 6451– 6455 DOI: 10.1073/pnas.221458298 [Crossref], [PubMed], [CAS], [Google
Scholar]

1861. Seeman, N. C. At the Crossroads of Chemistry, Biology, and Materials Chem. Biol. 2003, 10, 1151– 1159
 DOI: 10.1016/j.chembiol.2003.12.002 [Crossref], [PubMed], [CAS], [Google Scholar]

1862. Seeman, N. C. DNA in a Material World Nature 2003, 421, 427– 431 DOI: 10.1038/nature01406
[Crossref], [PubMed], [CAS], [Google Scholar]
1863. Seeman, N. C. Molecular Engineering: Nanotechnology and the Double Helix Sci. Am. 2004, 290, 34–
43 [Crossref], [PubMed], [Google Scholar]   

1864. Yan, H. Materials science. Nucleic acid nanotechnology Science 2004, 306, 2048– 2049
 DOI: 10.1126/science.1106754 [Crossref], [PubMed], [CAS], [Google Scholar]

1865. Brucale, M.; Zuccheri, G.; Samori, B. The dynamic properties of an intramolecular transition from DNA
duplex to cytosine-thymine motif triplex Org. Biomol. Chem. 2005, 3, 575– 577 DOI: 10.1039/b418353n
[Crossref], [PubMed], [CAS], [Google Scholar]

1866. Aldaye, F. A.; Palmer, A. L.; Sleiman, H. F. Assembling materials with DNA as the guide Science 2008, 321,
1795– 1799 DOI: 10.1126/science.1154533 [Crossref], [PubMed], [CAS], [Google Scholar]

1867. Gothelf, K. V. Materials science. LEGO-like DNA structures Science 2012, 338, 1159– 1160
 DOI: 10.1126/science.1229960 [Crossref], [PubMed], [CAS], [Google Scholar]

1868. Marras, A. E.; Zhou, L. F.; Su, H. J.; Castro, C. E. Programmable motion of DNA origami mechanisms Proc.
Natl. Acad. Sci. U. S. A. 2015, 112, 713– 718 DOI: 10.1073/pnas.1408869112 [Crossref], [PubMed],
[CAS], [Google Scholar]

1869. Zhang, D. Y.; Seelig, G. Dynamic DNA nanotechnology using strand-displacement reactions Nat. Chem.
2011, 3, 103– 113 DOI: 10.1038/nchem.957 [Crossref], [PubMed], [CAS], [Google Scholar]

1870. Hamada, S.; Murata, S. Substrate-Assisted Assembly of Interconnected Single-Duplex DNA


Nanostructures Angew. Chem., Int. Ed. 2009, 48, 6820– 6823 DOI: 10.1002/anie.200902662 [Crossref],
[PubMed], [CAS], [Google Scholar]

1871. Kalle Gehring, J.-L. L. a. M. G. A tetrameric DNA structure with protonated cytosine Nature 1993, 363,
561– 565 DOI: 10.1038/363561a0 [Crossref], [PubMed], [Google Scholar]

1872. Chen, L.; C, L.; Zhang, X.; Rich, A. Crystal structure of a four-stranded intercalated DNA: d(C4)
Biochemistry 1994, 33, 13540– 13546 DOI: 10.1021/bi00250a005 [ACS Full Text ], [CAS], [Google Scholar]

1873. Alberti, P.; Mergny, J. L. DNA duplex-quadruplex exchange as the basis for a nanomolecular machine
Proc. Natl. Acad. Sci. U. S. A. 2003, 100, 1569– 1573 DOI: 10.1073/pnas.0335459100 [Crossref], [PubMed],
[CAS], [Google Scholar]
  
1874. Dittmer, W. U.; Reuter, A.; Simmel, F. C. A DNA-based machine that can cyclically bind and release
thrombin Angew. Chem., Int. Ed. 2004, 43, 3550– 3553 DOI: 10.1002/anie.200353537 [Crossref], [PubMed],
[CAS], [Google Scholar]

1875. Davis, J. T.; Spada, G. P. Supramolecular architectures generated by self-assembly of guanosine


derivatives Chem. Soc. Rev. 2007, 36, 296– 313 DOI: 10.1039/B600282J [Crossref], [PubMed], [CAS], [Google
Scholar]

1876. Collie, G. W.; Parkinson, G. N. The application of DNA and RNA G-quadruplexes to therapeutic medicines
Chem. Soc. Rev. 2011, 40, 5867– 5892 DOI: 10.1039/c1cs15067g [Crossref], [PubMed], [CAS], [Google Scholar]

1877. Miyake, Y.; Togashi, H.; Tashiro, M.; Yamaguchi, M.; Oda, S.; Kudo, M.; Tanaka, Y.; Kondo, Y.; Sawa, R.;
Fujimoto, T., et al. MercuryII-Mediated Formation of Thymine–HgII–Thymine Base Pairs in DNA Duplexes J. Am.
Chem. Soc. 2006, 128, 2172– 2173 DOI: 10.1021/ja056354d [ACS Full Text ], [CAS], [Google Scholar]

1878. Tanaka, Y.; Oda, S.; Yamaguchi, H.; Kondo, Y.; Kojima, C.; Ono, A. 15N–15N J-Coupling Across HgII: Direct
Observation of HgII-Mediated T–T Base Pairs in a DNA Duplex J. Am. Chem. Soc. 2007, 129, 244– 245
 DOI: 10.1021/ja065552h [ACS Full Text ], [CAS], [Google Scholar]

1879. Yurke, B.; Tuber eld, A. J.; Mills, A. P., Jr.; Simmel, F. C.; Neumann, J. L. A DNA-fuelled molecular machine
made of DNA Nature 2000, 406, 605– 608 DOI: 10.1038/35020524 [Crossref], [PubMed], [CAS], [Google Scholar]

1880. Turber eld, A.; Mitchell, J.; Yurke, B.; Mills, A.; Blakey, M.; Simmel, F. DNA Fuel for Free-Running
Nanomachines Phys. Rev. Lett. 2003, 90, 118102 DOI: 10.1103/PhysRevLett.90.118102 [Crossref], [PubMed],
[CAS], [Google Scholar]

1881. Liu, D.; Bruckbauer, A.; Abell, C.; Balasubramanian, S.; Kang, D.-J.; Klenerman, D.; Zhou, D. A Reversible pH-
Driven DNA Nanoswitch Array J. Am. Chem. Soc. 2006, 128, 2067– 2071 DOI: 10.1021/ja0568300 [ACS Full Text
], [CAS], [Google Scholar]

1882. Liang, X.; Nishioka, H.; Takenaka, N.; Asanuma, H. A DNA nanomachine powered by light irradiation
ChemBioChem 2008, 9, 702– 705 DOI: 10.1002/cbic.200700649 [Crossref], [PubMed], [CAS], [Google Scholar]

1883. Li, D.; Wieckowska, A.; Willner, I. Optical Analysis of Hg2+ Ions by Oligonucleotide–Gold-Nanoparticle
Hybrids and DNA-Based Machines Angew. Chem. 2008, 120, 3991– 3995 DOI: 10.1002/ange.200705991
[Crossref], [Google Scholar]
  
1884. Song, G.; Chen, M.; Chen, C.; Wang, C.; Hu, D.; Ren, J.; Qu, X. Design of proton-fueled tweezers for
controlled, multi-function DNA-based molecular device Biochimie 2010, 92, 121– 127
 DOI: 10.1016/j.biochi.2009.10.007 [Crossref], [PubMed], [CAS], [Google Scholar]

1885. Saghatelian, A.; Volcker, N. H.; Guckian, K. M.; Lin, V. S. Y.; Ghadin, M. R. DNA-Based Photonic Logic Gates:
AND, NAND, and INHIBIT J. Am. Chem. Soc. 2003, 125, 346– 347 DOI: 10.1021/ja029009m [ACS Full Text ],
[CAS], [Google Scholar]

1886. Okamoto, A.; Tanaka, K.; Saito, I. DNA Logic Gates J. Am. Chem. Soc. 2004, 126, 9458– 9463
 DOI: 10.1021/ja047628k [ACS Full Text ], [CAS], [Google Scholar]

1887. Lee, J. F.; Stovall, G. M.; Ellington, A. D. Aptamer therapeutics advance Curr. Opin. Chem. Biol. 2006, 10,
282– 289 DOI: 10.1016/j.cbpa.2006.03.015 [Crossref], [PubMed], [CAS], [Google Scholar]

1888. Shlyahovsky, B.; Li, Y.; Lioubashevski, O.; Elbaz, J.; Willner, I. Logic Gates and Antisense DNA Devices
Operating on a Translator Nucleic Acid Scaffold ACS Nano 2009, 3, 1831– 1843 DOI: 10.1021/nn900085x [ACS
Full Text ], [CAS], [Google Scholar]

1889. Wilner, O. I.; Shimron, S.; Weizmann, Y.; Wang, Z.-G.; Willner, I. Self-Assembly of Enzymes on DNA
Scaffolds: En Route to Biocatalytic Cascades and the Synthesis of Metallic Nanowires Nano Lett. 2009, 9, 2040–
2043 DOI: 10.1021/nl900302z [ACS Full Text ], [CAS], [Google Scholar]

1890. Wilner, O. I.; Weizmann, Y.; Gill, R.; Lioubashevski, O.; Freeman, R.; Willner, I. Enzyme cascades activated
on topologically programmed DNA scaffolds Nat. Nanotechnol. 2009, 4, 249– 254 DOI: 10.1038/nnano.2009.50
[Crossref], [PubMed], [CAS], [Google Scholar]

1891. Teller, C.; Willner, I. Organizing protein-DNA hybrids as nanostructures with programmed functionalities
Trends Biotechnol. 2010, 28, 619– 628 DOI: 10.1016/j.tibtech.2010.09.005 [Crossref], [PubMed], [CAS], [Google
Scholar]

1892. Qian, L.; Winfree, E.; Bruck, J. Neural network computation with DNA strand displacement cascades
Nature 2011, 475, 368– 372 DOI: 10.1038/nature10262 [Crossref], [PubMed], [CAS], [Google Scholar]

1893. Winfree, L. Q. a. E. Scaling Up Digital Circuit Computation with DNA Strand Displacement Cascades
Science 2011, 332, 1196– 1201 DOI: 10.1126/science.1200520 [Crossref], [PubMed], [CAS], [Google Scholar]
  
1894. Pei, H.; Liang, L.; Yao, G.; Li, J.; Huang, Q.; Fan, C. Recon gurable Three-Dimensional DNA Nanostructures
for the Construction of Intracellular Logic Sensors Angew. Chem., Int. Ed. 2012, 51, 9020– 9024
 DOI: 10.1002/anie.201202356 [Crossref], [PubMed], [CAS], [Google Scholar]

1895. Fu, Y.; Zeng, D.; Chao, J.; Jin, Y.; Zhang, Z.; Liu, H.; Li, D.; Ma, H.; Huang, Q.; Gothelf, K. V.; Fan, C. Single-step
rapid assembly of DNA origami nanostructures for addressable nanoscale bioreactors J. Am. Chem. Soc. 2013,
135, 696– 702 DOI: 10.1021/ja3076692 [ACS Full Text ], [CAS], [Google Scholar]

1896. Miyamoto, T.; Razavi, S.; DeRose, R.; Inoue, T. Synthesizing biomolecule-based Boolean logic gates ACS
Synth. Biol. 2013, 2, 72– 82 DOI: 10.1021/sb3001112 [ACS Full Text ], [CAS], [Google Scholar]

1897. Gibbons, A.; Amos, M.; Hodgson, D. DNA computing Curr. Opin. Biotechnol. 1997, 8, 103– 106
 DOI: 10.1016/S0958-1669(97)80164-4 [Crossref], [PubMed], [CAS], [Google Scholar]

1898. Cox, J. C.; Cohen, D. S.; Ellington, A. D. The complexities of DNA computation Trends Biotechnol. 1999,
17, 151– 154 DOI: 10.1016/S0167-7799(99)01312-8 [Crossref], [PubMed], [CAS], [Google Scholar]

1899. Ruben, A. J.; Landweber, L. F. The past, present and future of molecular computing Nat. Rev. Mol. Cell
Biol. 2000, 1, 69– 72 DOI: 10.1038/35036086 [Crossref], [PubMed], [CAS], [Google Scholar]

1900. Gillmor, S. D.; Rugheimer, P. P.; Lagally, M. G. Computation with DNA on surfaces Surf. Sci. 2002, 500,
699– 721 DOI: 10.1016/S0039-6028(01)01524-2 [Crossref], [CAS], [Google Scholar]

1901. Reif, J. H. Computing: Successes and challenges Science 2002, 296, 478– 479
 DOI: 10.1126/science.1070978 [Crossref], [PubMed], [CAS], [Google Scholar]

1902. Livstone, M. S.; van Noort, D.; Landweber, L. F. Molecular computing revisited: a Moore’s law? Trends
Biotechnol. 2003, 21, 98– 101 DOI: 10.1016/S0167-7799(03)00007-6 [Crossref], [PubMed], [CAS], [Google
Scholar]

1903. Condon, A. Automata make antisense Nature 2004, 429, 351– 352 DOI: 10.1038/429351a [Crossref],
[PubMed], [CAS], [Google Scholar]

1904. Hasty, J.; McMillen, D.; Collins, J. J. Engineered gene circuits Nature 2002, 420, 224– 230
 DOI: 10.1038/nature01257 [Crossref], [PubMed], [CAS], [Google Scholar]
 2015,
1905. Kim, J.; Lee, J.; Hamada, S.; Murata, S.; Ha Park, S. Self-replication of DNA rings Nat. Nanotechnol.
10, 528– 533 DOI: 10.1038/nnano.2015.87 [Crossref], [PubMed], [CAS], [Google Scholar]

1906. Park, S. Y.; Lytton-Jean, A. K.; Lee, B.; Weigand, S.; Schatz, G. C.; Mirkin, C. A. DNA-programmable
nanoparticle crystallization Nature 2008, 451, 553– 556 DOI: 10.1038/nature06508 [Crossref], [PubMed],
[CAS], [Google Scholar]

1907. Nykypanchuk, D.; Maye, M. M.; van der Lelie, D.; Gang, O. DNA-guided crystallization of colloidal
nanoparticles Nature 2008, 451, 549– 552 DOI: 10.1038/nature06560 [Crossref], [PubMed], [CAS], [Google
Scholar]

1908. Wang, T.; Sha, R. J.; Dreyfus, R.; Leunissen, M. E.; Maass, C.; Pine, D. J.; Chaikin, P. M.; Seeman, N. C. Self-
replication of information-bearing nanoscale patterns Nature 2011, 478, 225– 228 DOI: 10.1038/nature10500
[Crossref], [PubMed], [CAS], [Google Scholar]

1909. Stojanovic, M. N.; Stefanovic, D. A deoxyribozyme-based molecular automaton Nat. Biotechnol. 2003, 21,
1069– 1074 DOI: 10.1038/nbt862 [Crossref], [PubMed], [CAS], [Google Scholar]

1910. Tabor, J. J.; Ellington, A. D. Playing to win at DNA computation Nat. Biotechnol. 2003, 21, 1013– 1015
 DOI: 10.1038/nbt0903-1013 [Crossref], [PubMed], [CAS], [Google Scholar]

1911. Macdonald, J.; Li, Y.; Sutovic, M.; Lederman, H.; Pendri, K.; Lu, W. H.; Andrews, B. L.; Stefanovic, D.;
Stojanovic, M. N. Medium scale integration of molecular logic gates in an automaton Nano Lett. 2006, 6, 2598–
2603 DOI: 10.1021/nl0620684 [ACS Full Text ], [CAS], [Google Scholar]

1912. Liu, M.; Fu, J.; Hejesen, C.; Yang, Y.; Woodbury, N. W.; Gothelf, K.; Liu, Y.; Yan, H. A DNA tweezer-actuated
enzyme nanoreactor Nat. Commun. 2013, 4, 2127 DOI: 10.1038/ncomms3127 [Crossref], [PubMed],
[CAS], [Google Scholar]

1913. Elbaz, J.; Wang, Z.-G.; Orbach, R.; Willner, I. pH-stimulated concurrent mechanical activation of two DNA
″tweezers″. A ″SET-RESET″ logic gate system Nano Lett. 2009, 9, 4510– 4514 DOI: 10.1021/nl902859m [ACS Full
Text ], [CAS], [Google Scholar]

1914. Wang, Z.-G.; Elbaz, J.; Remacle, F.; Levine, R. D.; Willner, I. All-DNA nite-state automata with nite
memory Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 21996– 22001 DOI: 10.1073/pnas.1015858107 [Crossref],
[PubMed], [CAS], [Google Scholar]

  
1915. Xin, L.; Zhou, C.; Yang, Z.; Liu, D. Regulation of an enzyme cascade reaction by a DNA machine Small
2013, 9, 3088– 3091 DOI: 10.1002/smll.201300019 [Crossref], [PubMed], [CAS], [Google Scholar]

1916. Elbaz, J.; Cecconello, A.; Fan, Z.; Govorov, A. O.; Willner, I. Powering the programmed nanostructure and
function of gold nanoparticles with catenated DNA machines Nat. Commun. 2013, 4, 2000
 DOI: 10.1038/ncomms3000 [Crossref], [PubMed], [CAS], [Google Scholar]

1917. Lange, H.; Juarez, B. H.; Carl, A.; Richter, M.; Bastus, N. G.; Weller, H.; Thomsen, C.; von Klitzing, R.; Knorr,
A. Tunable plasmon coupling in distance-controlled gold nanoparticles Langmuir 2012, 28, 8862– 8866
 DOI: 10.1021/la3001575 [ACS Full Text ], [CAS], [Google Scholar]

1918. Kuzyk, A.; Schreiber, R.; Fan, Z.; Pardatscher, G.; Roller, E. M.; Hogele, A.; Simmel, F. C.; Govorov, A. O.; Liedl,
T. DNA-based self-assembly of chiral plasmonic nanostructures with tailored optical response Nature 2012, 483,
311– 314 DOI: 10.1038/nature10889 [Crossref], [PubMed], [CAS], [Google Scholar]

1919. Liu, Y.; Kuzuya, A.; Sha, R.; Guillaume, J.; Wang, R.; Canary, J. W.; Seeman, N. C. Coupling Across a DNA
Helical Turn Yields a Hybrid DNA/Organic Catenane Doubly Tailed with Functional Termini J. Am. Chem. Soc.
2008, 130, 10882– 10883 DOI: 10.1021/ja8041096 [ACS Full Text ], [CAS], [Google Scholar]

1920. Han, D.; Pal, S.; Liu, Y.; Yan, H. Folding and cutting DNA into recon gurable topological nanostructures
Nat. Nanotechnol. 2010, 5, 712– 717 DOI: 10.1038/nnano.2010.193 [Crossref], [PubMed], [CAS], [Google Scholar]

1921. Lu, C. H.; Cecconello, A.; Elbaz, J.; Credi, A.; Willner, I. A three-station DNA catenane rotary motor with
controlled directionality Nano Lett. 2013, 13, 2303– 2308 DOI: 10.1021/nl401010e [ACS Full Text ],
[CAS], [Google Scholar]

1922. Samori, B.; Zuccheri, G. DNA codes for nanoscience Angew. Chem., Int. Ed. 2005, 44, 1166– 1181
 DOI: 10.1002/anie.200400652 [Crossref], [PubMed], [CAS], [Google Scholar]

1923. Chen, Y.; Mao, C. Reprogramming DNA-directed reactions on the basis of a DNA conformational change
J. Am. Chem. Soc. 2004, 126, 13240– 13241 DOI: 10.1021/ja045718j [ACS Full Text ], [CAS], [Google Scholar]

1924. Storhoff, J. J.; Mirkin, C. A. Programmed materials synthesis with DNA Chem. Rev. 1999, 99, 1849– 1862
 DOI: 10.1021/cr970071p [ACS Full Text ], [CAS], [Google Scholar]
1925. Braun, E.; Keren, K. From DNA to transistors Adv. Phys. 2004, 53, 441– 496   
 DOI: 10.1080/00018730412331294688 [Crossref], [CAS], [Google Scholar]

1926. Wengel, J. Nucleic acid nanotechnology - towards Angstrom-scale engineering Org. Biomol. Chem. 2004,
2, 277– 280 DOI: 10.1039/b313986g [Crossref], [PubMed], [CAS], [Google Scholar]

1927. Feldkamp, U.; Niemeyer, C. M. Rational design of DNA nanoarchitectures Angew. Chem., Int. Ed. 2006, 45,
1856– 1876 DOI: 10.1002/anie.200502358 [Crossref], [PubMed], [CAS], [Google Scholar]

1928. Broude, N. E. Stem-loop oligonucleotides: a robust tool for molecular biology and biotechnology Trends
Biotechnol. 2002, 20, 249– 256 DOI: 10.1016/S0167-7799(02)01942-X [Crossref], [PubMed], [CAS], [Google
Scholar]

1929. Tan, W. H.; Wang, K. M.; Drake, T. J. Molecular beacons Curr. Opin. Chem. Biol. 2004, 8, 547– 553
 DOI: 10.1016/j.cbpa.2004.08.010 [Crossref], [PubMed], [CAS], [Google Scholar]

1930. Shen, W. Q.; Bruist, M. F.; Goodman, S. D.; Seeman, N. C. A protein-driven DNA device that measures the
excess binding energy of proteins that distort DNA Angew. Chem., Int. Ed. 2004, 43, 4750– 4752
 DOI: 10.1002/anie.200460302 [Crossref], [PubMed], [CAS], [Google Scholar]

1931. Nutiu, R.; Li, Y. F. A DNA-protein nanoengine for ″On-Demand″ release and precise delivery of molecules
Angew. Chem., Int. Ed. 2005, 44, 5464– 5467 DOI: 10.1002/anie.200501214 [Crossref], [PubMed], [CAS], [Google
Scholar]

1932. Loh, I. Y.; Cheng, J.; Tee, S. R.; Efremov, A.; Wang, Z. S. From Bistate Molecular Switches to Self-Directed
Track-Walking Nanomotors ACS Nano 2014, 8, 10293– 10304 DOI: 10.1021/nn5034983 [ACS Full Text ],
[CAS], [Google Scholar]

1933. Leigh, D. A.; Lewandowska, U.; Lewandowski, B.; Wilson, M. R. Synthetic Molecular Walkers Top. Curr.
Chem. 2014, 354, 111– 138 DOI: 10.1007/128_2014_546 [Crossref], [PubMed], [CAS], [Google Scholar]

1934. Wang, C. Y.; Ren, J. S.; Qu, X. G. A stimuli responsive DNA walking device Chem. Commun. 2011, 47,
1428– 1430 DOI: 10.1039/C0CC04234J [Crossref], [PubMed], [CAS], [Google Scholar]
1935. Liedl, T.; Olapinski, M.; Simmel, F. C. A Surface-Bound DNA Switch Driven by a Chemical Oscillator Angew.
 
Chem., Int. Ed. 2006, 45, 5007– 5010 DOI: 10.1002/anie.200600353 [Crossref], [PubMed], [CAS], [Google 
Scholar]

1936. Bromley, E. H.; Kuwada, N. J.; Zuckermann, M. J.; Donadini, R.; Samii, L.; Blab, G. A.; Gemmen, G. J.; Lopez,
B. J.; Curmi, P. M.; Forde, N. R.; Woolfson, D. N.; Linke, H. The Tumbleweed: towards a synthetic proteinmotor
HFSP J. 2009, 3, 204– 212 DOI: 10.2976/1.3111282 [Crossref], [PubMed], [CAS], [Google Scholar]

1937. Rothemund, P. W. Folding DNA to create nanoscale shapes and patterns Nature 2006, 440, 297– 302
 DOI: 10.1038/nature04586 [Crossref], [PubMed], [CAS], [Google Scholar]

1938. Smith, L. M. Nanotechnology: Molecular robots on the move Nature 2010, 465, 167– 168
 DOI: 10.1038/465167a [Crossref], [PubMed], [CAS], [Google Scholar]

1939. Wickham, S. F.; Endo, M.; Katsuda, Y.; Hidaka, K.; Bath, J.; Sugiyama, H.; Turber eld, A. J. Direct
observation of stepwise movement of a synthetic molecular transporter Nat. Nanotechnol. 2011, 6, 166– 169
 DOI: 10.1038/nnano.2010.284 [Crossref], [PubMed], [CAS], [Google Scholar]

1940. Bath, J.; Green, S. J.; Turber eld, A. J. A free-running DNA motor powered by a nicking enzyme Angew.
Chem., Int. Ed. 2005, 44, 4358– 4361 DOI: 10.1002/anie.200501262 [Crossref], [PubMed], [CAS], [Google Scholar]

1941. Mai, J.; Sokolov, I. M.; Blumen, A. Directed particle diffusion under ″burnt bridges″ conditions Phys. Rev. E:
Stat. Phys., Plasmas, Fluids, Relat. Interdiscip. Top. 2001, 64, 011102 DOI: 10.1103/PhysRevE.64.011102
[Crossref], [CAS], [Google Scholar]

1942. Antal, T.; Krapivsky, P. L. “Burnt-bridge” mechanism of molecular motor motion. Phys. Rev. E 2005, 72,
 DOI: 10.1103/PhysRevE.72.046104 . [Crossref], [Google Scholar]

1943. Yin, P.; Choi, H. M.; Calvert, C. R.; Pierce, N. A. Programming biomolecular self-assembly pathways Nature
2008, 451, 318– 322 DOI: 10.1038/nature06451 [Crossref], [PubMed], [CAS], [Google Scholar]

1944. Green, S. J.; Bath, J.; Turber eld, A. J. Coordinated chemomechanical cycles: a mechanism for
autonomous molecular motion Phys. Rev. Lett. 2008, 101, 238101 DOI: 10.1103/PhysRevLett.101.238101
[Crossref], [PubMed], [CAS], [Google Scholar]

1945. Bath, J.; Green, S. J.; Allen, K. E.; Turber eld, A. J. Mechanism for a directional, processive, and reversible
DNA motor Small 2009, 5, 1513– 1516 DOI: 10.1002/smll.200900078 [Crossref], [PubMed], [CAS], [Google
Scholar]

  
1946. Pei, R.; Taylor, S. K.; Stefanovic, D.; Rudchenko, S.; Mitchell, T. E.; Stojanovic, M. N. Behavior of
polycatalytic assemblies in a substrate-displaying matrix J. Am. Chem. Soc. 2006, 128, 12693– 12699
 DOI: 10.1021/ja058394n [ACS Full Text ], [CAS], [Google Scholar]

1947. You, M.; Chen, Y.; Zhang, X.; Liu, H.; Wang, R.; Wang, K.; Williams, K. R.; Tan, W. An autonomous and
controllable light-driven DNA walking device Angew. Chem., Int. Ed. 2012, 51, 2457– 2460
 DOI: 10.1002/anie.201107733 [Crossref], [PubMed], [CAS], [Google Scholar]

1948. Kay, E. R.; Leigh, D. A. Rise of the Molecular Machines Angew. Chem., Int. Ed. 2015, 54, 10080– 10088
 DOI: 10.1002/anie.201503375 [Crossref], [PubMed], [CAS], [Google Scholar]

1949. Turing, A. M. Proc. London Math. Soc. 1937, 42, 230– 265 DOI: 10.1112/plms/s2-42.1.230
[Crossref], [Google Scholar]

Cited By

This article is cited by 916 publications.

1. Giacomo Prando, Jacopo Perego, Mattia Negroni, Mauro Riccò, Silvia Bracco, Angiolina Comotti, Piero
Sozzani, Pietro Carretta. Molecular Rotors in a Metal–Organic Framework: Muons on a Hyper-Fast Carousel.
Nano Letters 2020, 20 (10) , 7613-7618. https://doi.org/10.1021/acs.nanolett.0c03140

2. Maggie He, Timothy M. Swager. Aryl Migration on Graphene. Journal of the American Chemical Society
2020, Article ASAP.

3. Matt Guest, Roya Mir, Gregory Foran, Brianne Hickson, Aleksandar Necakov, Travis Dudding.
Trisaminocyclopropenium Cations as Small-Molecule Organic Fluorophores: Design Guidelines and
Bioimaging Applications. The Journal of Organic Chemistry 2020, Article ASAP.

4. Meng-Yue Guo, Peng Li, Shuai-Liang Yang, Ran Bu, Xian-Qing Piao, En-Qing Gao. Distinct and Selective
Amine- and Anion-Responsive Behaviors of an Electron-De cient and Anion-Exchangeable Metal–Organic
Framework. ACS Applied Materials & Interfaces 2020, 12 (39) , 43958-43966.
https://doi.org/10.1021/acsami.0c14648

5. Wei-Jian Li, Zhubin Hu, Lin Xu, Xu-Qing Wang, Wei Wang, Guang-Qiang Yin, Dan-Yang Zhang, Zhenrong Sun,
Xiaopeng Li, Haitao Sun, Hai-Bo Yang. Rotaxane-Branched Dendrimers with Enhanced Photosensitization.
Journal of the American Chemical Society 2020, 142 (39) , 16748-16756.
https://doi.org/10.1021/jacs.0c07292
6. Gregory B. Boursalian, Eise R. Nijboer, Ruth Dorel, Lukas Pfeifer, Omer Markovitch, Alex Blokhuis, Ben L.
Feringa. All-Photochemical Rotation of Molecular Motors with a Phosphorus Stereoelement. Journal of the
  
American Chemical Society 2020, 142 (39) , 16868-16876. https://doi.org/10.1021/jacs.0c08249

7. Shobhana Krishnaswamy, Soumyakanta Prusty, Daniel Chartrand, Garry S. Hanan, Dillip K. Chand.
Conformational Solvatomorphism in a [2]Catenane. Crystal Growth & Design 2020, 20 (9) , 5820-5833.
https://doi.org/10.1021/acs.cgd.0c00421

8. Zeng Yi, Xiangyu Chen, Guangcan Chen, Zhiwen Deng, Qiulan Tong, Zhe Sun, Xiaomin Ma, Wen Su, Lei Ma,
Yaqin Ran, Xudong Li. General Nanomedicine Platform by Solvent-Mediated Disassembly/Reassembly of
Scalable Natural Polyphenol Colloidal Spheres. ACS Applied Materials & Interfaces 2020, 12 (34) , 37914-
37928. https://doi.org/10.1021/acsami.0c11650

9. Jing Zhang, Benzhao He, Wenjie Wu, Parvej Alam, Han Zhang, Junyi Gong, Fengyan Song, Zaiyu Wang,
Herman H. Y. Sung, Ian D. Williams, Zhiming Wang, Jacky W. Y. Lam, Ben Zhong Tang. Molecular Motions in
AIEgen Crystals: Turning on Photoluminescence by Force-Induced Filament Sliding. Journal of the American
Chemical Society 2020, 142 (34) , 14608-14618. https://doi.org/10.1021/jacs.0c06305

10. Qing-Hui Guo, Yunyan Qiu, Xinyi Kuang, Jiaqi Liang, Yuanning Feng, Long Zhang, Yang Jiao, Dengke Shen,
R. Dean Astumian, J. Fraser Stoddart. Arti cial Molecular Pump Operating in Response to Electricity and Light.
Journal of the American Chemical Society 2020, 142 (34) , 14443-14449.
https://doi.org/10.1021/jacs.0c06663

11. José Augusto Berrocal, Lukas Pfeifer, Dorus Heijnen, Ben L. Feringa. Synthesis of Core-Modi ed Third-
Generation Light-Driven Molecular Motors. The Journal of Organic Chemistry 2020, 85 (16) , 10670-10680.
https://doi.org/10.1021/acs.joc.0c01235

12. Zheng Cui, Ye Lu, Xiang Gao, Hui-Jun Feng, Guo-Xin Jin. Stereoselective Synthesis of a Topologically
Chiral Solomon Link. Journal of the American Chemical Society 2020, 142 (32) , 13667-13671.
https://doi.org/10.1021/jacs.0c05366

13. Adrian Saura-Sanmartin, Alberto Martinez-Cuezva, Delia Bautista, Mara R. B. Marzari, Marcos A. P.
Martins, Mateo Alajarin, Jose Berna. Copper-Linked Rotaxanes for the Building of Photoresponsive Metal
Organic Frameworks with Controlled Cargo Delivery. Journal of the American Chemical Society 2020, 142 (31)
, 13442-13449. https://doi.org/10.1021/jacs.0c04477

14. Alberto Martinez-Cuezva, Adrian Saura-Sanmartin, Mateo Alajarin, Jose Berna. Mechanically Interlocked
Catalysts for Asymmetric Synthesis. ACS Catalysis 2020, 10 (14) , 7719-7733.
https://doi.org/10.1021/acscatal.0c02032

15. Ieva Liepuoniute, Chau Minh Huynh, Salvador Perez-Estrada, Yangyang Wang, Saeed Khan, K. N. Houk,
Miguel A. Garcia-Garibay. Enhanced Rotation by Ground State Destabilization in Amphidynamic Crystals of a
Dipolar 2,3-Di uorophenylene Rotator as Established by Solid State 2H NMR and Dielectric Spectroscopy. The
Journal of Physical Chemistry C 2020, 124 (28) , 15391-15398. https://doi.org/10.1021/acs.jpcc.0c05314

16. Wen-Xi Gao, Hui-Jun Feng, Bei-Bei Guo, Ye Lu, Guo-Xin Jin. Coordination-Directed Construction of
Molecular Links. Chemical Reviews 2020, 120 (13) , 6288-6325.
https://doi.org/10.1021/acs.chemrev.0c00321

17. Ieva Liepuoniute, Jacob N. Sanders, Miguel A. Garcia-Garibay, K. N. Houk. Computational Investigation
into Ligand Effects on Correlated Geared Dynamics in Dirhodium Supramolecular Gears—Insights Beyond the
NMR Experimental Window. The Journal of Organic Chemistry 2020, 85 (13) , 8695-8701.
https://doi.org/10.1021/acs.joc.0c01230
  
18. Aijaz A. Dar, Arshid A. Ganie. Irreversible Thermochromism in Organic Salts of Sulfonated Anils. Crystal
Growth & Design 2020, 20 (6) , 3888-3897. https://doi.org/10.1021/acs.cgd.0c00188

19. Wojciech Danowski, Fabio Castiglioni, Andy S. Sardjan, Simon Krause, Lukas Pfeifer, Diederik Roke,
Angiolina Comotti, Wesley R. Browne, Ben L. Feringa. Visible-Light-Driven Rotation of Molecular Motors in a
Dual-Function Metal–Organic Framework Enabled by Energy Transfer. Journal of the American Chemical
Society 2020, 142 (19) , 9048-9056. https://doi.org/10.1021/jacs.0c03063

20. Wei-Jian Li, Wei Wang, Xu-Qing Wang, Mu Li, Yubin Ke, Rui Yao, Jin Wen, Guang-Qiang Yin, Bo Jiang,
Xiaopeng Li, Panchao Yin, Hai-Bo Yang. Daisy Chain Dendrimers: Integrated Mechanically Interlocked
Molecules with Stimuli-Induced Dimension Modulation Feature. Journal of the American Chemical Society
2020, 142 (18) , 8473-8482. https://doi.org/10.1021/jacs.0c02475

21. Ming Chen, Xiaoyan Zhang, Junkai Liu, Feng Liu, Ruoyao Zhang, Peifa Wei, Haitao Feng, Mei Tu, Anjun Qin,
Jacky W. Y. Lam, Dan Ding, Ben Zhong Tang. Evoking Photothermy by Capturing Intramolecular Bond
Stretching Vibration-Induced Dark-State Energy. ACS Nano 2020, 14 (4) , 4265-4275.
https://doi.org/10.1021/acsnano.9b09625

22. Amanda Acevedo-Jake, Andrew T. Ball, Marzia Galli, Mikiembo Kukwikila, Mathieu Denis, Daniel G.
Singleton, Ali Tavassoli, Stephen M. Goldup. AT-CuAAC Synthesis of Mechanically Interlocked Oligonucleotides.
Journal of the American Chemical Society 2020, 142 (13) , 5985-5990.
https://doi.org/10.1021/jacs.0c01670

23. Shuangli Du, Han Liu, Xueguang Shao, Christophe Chipot, Wensheng Cai. Free-Energy Landscape of
Stepwise, Directional Motion in Multiple Molecular Switches. The Journal of Physical Chemistry C 2020, 124
(11) , 6448-6453. https://doi.org/10.1021/acs.jpcc.9b11872

24. Rui-Lian Lin, Ran Li, Hao Shi, Kun Zhang, Di Meng, Wen-Qi Sun, Kai Chen, Jing-Xin Liu. Symmetrical-
Tetramethyl-Cucurbit[6]uril-Driven Movement of Cucurbit[7]uril Gives Rise to Heterowheel [4]Pseudorotaxanes.
The Journal of Organic Chemistry 2020, 85 (5) , 3568-3575. https://doi.org/10.1021/acs.joc.9b03283

25. Toshikazu Takata. Switchable Polymer Materials Controlled by Rotaxane Macromolecular Switches. ACS
Central Science 2020, 6 (2) , 129-143. https://doi.org/10.1021/acscentsci.0c00002

26. Andrew W. Heard, Stephen M. Goldup. Simplicity in the Design, Operation, and Applications of
Mechanically Interlocked Molecular Machines. ACS Central Science 2020, 6 (2) , 117-128.
https://doi.org/10.1021/acscentsci.9b01185

27. Krishna Gopal Chattaraj, Rabindranath Paul, Sandip Paul. Switching of Self-Assembly to Solvent-Assisted
Assembly of Molecular Motor: Unveiling the Mechanisms of Dynamic Control on Solvent Exchange. Langmuir
2020, 36 (7) , 1773-1792. https://doi.org/10.1021/acs.langmuir.9b03718

28. Shunjie Liu, Yanhua Cheng, Yuanyuan Li, Ming Chen, Jacky W. Y. Lam, Ben Zhong Tang. Manipulating
Solid-State Intramolecular Motion toward Controlled Fluorescence Patterns. ACS Nano 2020, 14 (2) , 2090-
2098. https://doi.org/10.1021/acsnano.9b08761

29. Noriyuki Tanaka, Yusuke Inagaki, Kentaro Yamaguchi, Wataru Setaka. Gear Alignments Due to Hydrogen-
Bonded Networks in a Crystal Structure of Resorcyltriptycene Hydrate and Its Transformation to a Nongearing
Anhydrate Crystal by Heating. Crystal Growth & Design 2020, 20 (2) , 1097-1102.
https://doi.org/10.1021/acs.cgd.9b01424
  
30. Shuai Jiang, Anke Kaltbeitzel, Minghan Hu, Oksana Suraeva, Daniel Crespy, Katharina Landfester. One-
Step Preparation of Fuel-Containing Anisotropic Nanocapsules with Stimuli-Regulated Propulsion. ACS Nano
2020, 14 (1) , 498-508. https://doi.org/10.1021/acsnano.9b06408

31. Liu-Pan Yang, Xiaoping Wang, Huan Yao, Wei Jiang. Naphthotubes: Macrocyclic Hosts with a Biomimetic
Cavity Feature. Accounts of Chemical Research 2020, 53 (1) , 198-208.
https://doi.org/10.1021/acs.accounts.9b00415

32. Theodore Tyrikos-Ergas, Giulio Fittolani, Peter H. Seeberger, Martina Delbianco. Structural Studies Using
Unnatural Oligosaccharides: Toward Sugar Foldamers. Biomacromolecules 2020, 21 (1) , 18-29.
https://doi.org/10.1021/acs.biomac.9b01090

33. Abir Goswami, Suchismita Saha, Pronay Kumar Biswas, Michael Schmittel. (Nano)mechanical Motion
Triggered by Metal Coordination: from Functional Devices to Networked Multicomponent Catalytic Machinery.
Chemical Reviews 2020, 120 (1) , 125-199. https://doi.org/10.1021/acs.chemrev.9b00159

34. Víctor García-López, Dongdong Liu, James M. Tour. Light-Activated Organic Molecular Motors and Their
Applications. Chemical Reviews 2020, 120 (1) , 79-124. https://doi.org/10.1021/acs.chemrev.9b00221

35. Gadiel Saper, Henry Hess. Synthetic Systems Powered by Biological Molecular Motors. Chemical Reviews
2020, 120 (1) , 288-309. https://doi.org/10.1021/acs.chemrev.9b00249

36. Aidan I. Brown, David A. Sivak. Theory of Nonequilibrium Free Energy Transduction by Molecular
Machines. Chemical Reviews 2020, 120 (1) , 434-459. https://doi.org/10.1021/acs.chemrev.9b00254

37. Damien Dattler, Gad Fuks, Joakim Heiser, Emilie Moulin, Alexis Perrot, Xuyang Yao, Nicolas Giuseppone.
Design of Collective Motions from Synthetic Molecular Switches, Rotors, and Motors. Chemical Reviews 2020,
120 (1) , 310-433. https://doi.org/10.1021/acs.chemrev.9b00288

38. Massimo Baroncini, Serena Silvi, Alberto Credi. Photo- and Redox-Driven Arti cial Molecular Motors.
Chemical Reviews 2020, 120 (1) , 200-268. https://doi.org/10.1021/acs.chemrev.9b00291

39. Ryota Iino, , Kazushi Kinbara, , Zev Bryant. Introduction: Molecular Motors. Chemical Reviews 2020, 120
(1) , 1-4. https://doi.org/10.1021/acs.chemrev.9b00819

40. Hassan Al Sabea, Lucie Norel, Olivier Galangau, Hussein Hijazi, Remi Métivier, Thierry Roisnel, Olivier
Maury, Christophe Bucher, François Riobé, Stéphane Rigaut. Dual Light and Redox Control of NIR
Luminescence with Complementary Photochromic and Organometallic Antennae. Journal of the American
Chemical Society 2019, 141 (51) , 20026-20030. https://doi.org/10.1021/jacs.9b11318

41. Tatu Kumpulainen, Matthijs R. Panman, Bert H. Bakker, Michiel Hilbers, Sander Woutersen, Albert M.
Brouwer. Accelerating the Shuttling in Hydrogen-Bonded Rotaxanes: Active Role of the Axle and the End
Station. Journal of the American Chemical Society 2019, 141 (48) , 19118-19129.
https://doi.org/10.1021/jacs.9b10005

42. Jacob B. Smith, Andrew M. Camp, Alexandra H. Farquhar, Stewart H. Kerr, Chun-Hsing Chen, Alexander J.
M. Miller. Organometallic Elaboration as a Strategy for Tuning the Supramolecular Characteristics of Aza-
Crown Ethers. Organometallics 2019, 38 (22) , 4392-4398. https://doi.org/10.1021/acs.organomet.9b00462

  
43. Lingyun Zhou, Fengting Lv, Libing Liu, Shu Wang. Water-Soluble Conjugated Organic Molecules as Optical
and Electrochemical Materials for Interdisciplinary Biological Applications. Accounts of Chemical Research
2019, 52 (11) , 3211-3222. https://doi.org/10.1021/acs.accounts.9b00427

44. Kazuma Okamura, Yusuke Inagaki, Hiroyuki Momma, Eunsang Kwon, Wataru Setaka. Gear Slippage in
Molecular Bevel Gears Bridged with a Group 14 Element. The Journal of Organic Chemistry 2019, 84 (22) ,
14636-14643. https://doi.org/10.1021/acs.joc.9b02214

45. Arthur H. G. David, Raquel Casares, Juan M. Cuerva, Araceli G. Campaña, Victor Blanco. A [2]Rotaxane-
Based Circularly Polarized Luminescence Switch. Journal of the American Chemical Society 2019, 141 (45) ,
18064-18074. https://doi.org/10.1021/jacs.9b07143

46. María Eugenia Ochoa, Rafael Arcos-Ramos, Pedro I. Ramirez-Montes, Herbert Höp , Marco A. Leyva,
Norberto Farfán, Rosa Santillan. Asymmetric Molecular Rotors Based on Steroidal Fragments. Organic
Building Blocks Displaying Versatile Supramolecular Steroid-Stacking Interactions. Crystal Growth & Design
2019, 19 (11) , 6114-6126. https://doi.org/10.1021/acs.cgd.9b00238

47. Yunyan Qiu, Long Zhang, Cristian Pezzato, Yuanning Feng, Weixingyue Li, Minh T. Nguyen, Chuyang
Cheng, Dengke Shen, Qing-Hui Guo, Yi Shi, Kang Cai, Fehaid M. Alsubaie, R. Dean Astumian, J. Fraser Stoddart.
A Molecular Dual Pump. Journal of the American Chemical Society 2019, 141 (44) , 17472-17476.
https://doi.org/10.1021/jacs.9b08927

48. Massimo Cametti, Yoko Sakata, Javier Martí-Rujas, Shigehisa Akine. ON/OFF Control of the Flipping
Motion of Diuranyl Bis(Salophen) Macrocycle by Extremely Strong Binding with Fluoride Ion. Inorganic
Chemistry 2019, 58 (21) , 14871-14875. https://doi.org/10.1021/acs.inorgchem.9b02587

49. Daisuke Shimoyama, Takeharu Haino. Conformational Characteristics of Feet-to-Feet-Connected


Biscavitands. The Journal of Organic Chemistry 2019, 84 (21) , 13483-13489.
https://doi.org/10.1021/acs.joc.9b01730

50. Liuzhou Gao, Guoqiang Wang, Jia Cao, Hui Chen, Yuming Gu, Xueting Liu, Xu Cheng, Jing Ma, Shuhua Li.
Lewis Acid-Catalyzed Selective Reductive Decarboxylative Pyridylation of N-Hydroxyphthalimide Esters:
Synthesis of Congested Pyridine-Substituted Quaternary Carbons. ACS Catalysis 2019, 9 (11) , 10142-10151.
https://doi.org/10.1021/acscatal.9b03798

51. Riku Kubota, Yui Sasaki, Tsukuru Minamiki, Tsuyoshi Minami. Chemical Sensing Platforms Based on
Organic Thin-Film Transistors Functionalized with Arti cial Receptors. ACS Sensors 2019, 4 (10) , 2571-2587.
https://doi.org/10.1021/acssensors.9b01114

52. Min Zhang, Guillaume De Bo. Mechanical Susceptibility of a Rotaxane. Journal of the American Chemical
Society 2019, 141 (40) , 15879-15883. https://doi.org/10.1021/jacs.9b06960

53. Dawei Zhang, Tanya K. Ronson, Songül Güryel, John D. Thoburn, David J. Wales, Jonathan R. Nitschke.
Temperature Controls Guest Uptake and Release from Zn4L4 Tetrahedra. Journal of the American Chemical
Society 2019, 141 (37) , 14534-14538. https://doi.org/10.1021/jacs.9b07307

54. Xu-Qing Wang, Wei-Jian Li, Wei Wang, Jin Wen, Ying Zhang, Hongwei Tan, Hai-Bo Yang. Construction of
Type III-C Rotaxane-Branched Dendrimers and Their Anion-Induced Dimension Modulation Feature. Journal of
the American Chemical Society 2019, 141 (35) , 13923-13930. https://doi.org/10.1021/jacs.9b06739
55. Takehiro Hirao, Naoyuki Hisano, Shigehisa Akine, Shin-ichi Kihara, Takeharu Haino. Ring–Chain

 Cleft.
Competition in Supramolecular Polymerization Directed by Molecular Recognition of the Bisporphyrin

Macromolecules 2019, 52 (16) , 6160-6168. https://doi.org/10.1021/acs.macromol.9b01012

56. Morgan E. Howe, Miguel A. Garcia-Garibay. The Roles of Intrinsic Barriers and Crystal Fluidity in
Determining the Dynamics of Crystalline Molecular Rotors and Molecular Machines. The Journal of Organic
Chemistry 2019, 84 (16) , 9835-9849. https://doi.org/10.1021/acs.joc.9b00993

57. Arshid A. Ganie, Aadil A. Ahangar, Aijaz A. Dar. Sulfonate···Pyridinium Supramolecular Synthon: A Robust
Interaction Utilized to Design Molecular Assemblies. Crystal Growth & Design 2019, 19 (8) , 4650-4660.
https://doi.org/10.1021/acs.cgd.9b00555

58. Morgan E. Howe, Miguel A. Garcia-Garibay. Fluorescence and Rotational Dynamics of a Crystalline
Molecular Rotor Featuring an Aggregation-Induced Emission Fluorophore. The Journal of Organic Chemistry
2019, 84 (15) , 9570-9576. https://doi.org/10.1021/acs.joc.9b01201

59. Si Yue Guo, Matthew J. Timm, Kai Huang, John C. Polanyi. Electron Attachment Leads to Unidirectional
In-Plane Molecular Rotation of Para-Chlorostyrene on Si(100). The Journal of Physical Chemistry C 2019, 123
(30) , 18425-18431. https://doi.org/10.1021/acs.jpcc.9b04076

60. Zhi-Xu Zhang, Tie Zhang, Ping-Ping Shi, Wan-Ying Zhang, Qiong Ye, Da-Wei Fu. Anion-Regulated Molecular
Rotor Crystal: The First Case of a Stator–Rotator Double Switch with Relaxation Behavior. The Journal of
Physical Chemistry Letters 2019, 10 (15) , 4237-4244. https://doi.org/10.1021/acs.jpclett.9b01503

61. Nikolai Kh Petrov, Denis A. Ivanov, Michael V. Al mov. Ultrafast Dynamics of Electronically Excited Host–
Guest Complexes of Cucurbiturils with Styryl Dyes. ACS Omega 2019, 4 (7) , 11500-11507.
https://doi.org/10.1021/acsomega.9b01158

62. Haochuan Chen, Hong Zhang, Xueguang Shao, Wensheng Cai. Tumbling of Anisole Units in Calixarene
Promotes Its Shuttling in Rotaxanes. The Journal of Physical Chemistry C 2019, 123 (29) , 18050-18055.
https://doi.org/10.1021/acs.jpcc.9b03870

63. Paola Franchi, Cecilia Poderi, Elisabetta Mezzina, Chiara Biagini, Stefano Di Stefano, Marco Lucarini. 2-
Cyano-2-phenylpropanoic Acid Triggers the Back and Forth Motions of an Acid–Base-Operated Paramagnetic
Molecular Switch. The Journal of Organic Chemistry 2019, 84 (14) , 9364-9368.
https://doi.org/10.1021/acs.joc.9b01164

64. Jens Kügel, Tim Zenger, Markus Leisegang, Matthias Bode. On the Impact of Geometrical Factors on Hot
Electron-Induced Tautomerization. The Journal of Physical Chemistry C 2019, 123 (27) , 17056-17061.
https://doi.org/10.1021/acs.jpcc.9b05602

65. Jiří Kaleta, Guillaume Bastien, Jin Wen, Martin Dračínský, Edward Tortorici, Ivana Císařová, Paul D. Beale,
Charles T. Rogers, Josef Michl. Bulk Inclusions of Double Pyridazine Molecular Rotors in Hexagonal Tris(o-
phenylene)cyclotriphosphazene. The Journal of Organic Chemistry 2019, 84 (13) , 8449-8467.
https://doi.org/10.1021/acs.joc.9b00553

66. Yuanyuan Chang, Zhongyu Wu, Qianxin Sun, Ying Zhuo, Yaqin Chai, Ruo Yuan. Simply Constructed and
Highly E cient Classi ed Cargo-Discharge DNA Robot: A DNA Walking Nanomachine Platform for
Ultrasensitive Multiplexed Sensing. Analytical Chemistry 2019, 91 (13) , 8123-8128.
https://doi.org/10.1021/acs.analchem.9b00363
67. Hao Zheng, Cailing Ni, Hang Chen, Daijun Zha, Yu Hai, Hebo Ye, Lei You. Regulation of Axial Chirality
through Dynamic Covalent Bond Constrained Biaryls. ACS Omega 2019, 4 (6) , 10273-10278.   
https://doi.org/10.1021/acsomega.9b01273

68. Zhao Wang. Selective Conduction of Organic Molecules via Free-Standing Graphene. The Journal of
Physical Chemistry C 2019, 123 (24) , 15166-15170. https://doi.org/10.1021/acs.jpcc.9b04131

69. Stefano Corra, Christiaan de Vet, Jessica Groppi, Marcello La Rosa, Serena Silvi, Massimo Baroncini,
Alberto Credi. Chemical On/Off Switching of Mechanically Planar Chirality and Chiral Anion Recognition in a
[2]Rotaxane Molecular Shuttle. Journal of the American Chemical Society 2019, 141 (23) , 9129-9133.
https://doi.org/10.1021/jacs.9b00941

70. Agostino Galanti, Jasmin Santoro, Rajesh Mannancherry, Quentin Duez, Valentin Diez-Cabanes, Michal
Valášek, Julien De Winter, Jérôme Cornil, Pascal Gerbaux, Marcel Mayor, Paolo Samorì. A New Class of Rigid
Multi(azobenzene) Switches Featuring Electronic Decoupling: Unravelling the Isomerization in Individual
Photochromes. Journal of the American Chemical Society 2019, 141 (23) , 9273-9283.
https://doi.org/10.1021/jacs.9b02544

71. Kanagaraj Naveen, Paramasivan Thirumalai Perumal, Deug-Hee Cho. Domino Palladium-Catalyzed Double
Norbornene Insertion/Annulation Reaction: Expeditious Synthesis of Overcrowded Tetrasubstituted Ole ns.
Organic Letters 2019, 21 (11) , 4350-4354. https://doi.org/10.1021/acs.orglett.9b01543

72. Baihao Shao, Hai Qian, Quan Li, Ivan Aprahamian. Structure Property Analysis of the Solution and Solid-
State Properties of Bistable Photochromic Hydrazones. Journal of the American Chemical Society 2019, 141
(20) , 8364-8371. https://doi.org/10.1021/jacs.9b03932

73. Habiburrahman Zul kri, Mark A. J. Koenis, Michael M. Lerch, Mariangela Di Donato, Wiktor Szymański,
Claudia Filippi, Ben L. Feringa, Wybren Jan Buma. Taming the Complexity of Donor–Acceptor Stenhouse
Adducts: Infrared Motion Pictures of the Complete Switching Pathway. Journal of the American Chemical
Society 2019, 141 (18) , 7376-7384. https://doi.org/10.1021/jacs.9b00341

74. He-Ye Zhou, Ying Han, Qiang Shi, Chuan-Feng Chen. Directional Transportation of a Helic[6]arene along a
Nonsymmetric Molecular Axle. The Journal of Organic Chemistry 2019, 84 (9) , 5872-5876.
https://doi.org/10.1021/acs.joc.9b00229

75. Jing-Jing Yu, Li-Yang Zhao, Zhao-Tao Shi, Qi Zhang, Gabor London, Wen-Jing Liang, Chuan Gao, Ming-
Ming Li, Xiao-Ming Cao, He Tian, Ben L. Feringa, Da-Hui Qu. Pumping a Ring-Sliding Molecular Motion by a
Light-Powered Molecular Motor. The Journal of Organic Chemistry 2019, 84 (9) , 5790-5802.
https://doi.org/10.1021/acs.joc.9b00783

76. Hong Zhang, Xueguang Shao, Christophe Chipot, Wensheng Cai. pH-Controlled Fluorescence Probes for
Rotaxane Isomerization. The Journal of Physical Chemistry C 2019, 123 (17) , 11304-11309.
https://doi.org/10.1021/acs.jpcc.9b02028

77. Charlie T. McTernan, Tanya K. Ronson, Jonathan R. Nitschke. Post-assembly Modi cation of Phosphine
Cages Controls Host–Guest Behavior. Journal of the American Chemical Society 2019, 141 (17) , 6837-6842.
https://doi.org/10.1021/jacs.9b02604

78. Noushaba Nusrat Mafy, Yuna Kim, Reji Thomas, Takehito Akasaka, Nobuyuki Tamaoki. Molecular
Crankshaft Effect Converting Piston-like Molecular Motion to Continuous Rotation of Macro Objects. ACS
Applied Materials & Interfaces 2019, 11 (16) , 15097-15102. https://doi.org/10.1021/acsami.9b03706

  
79. Arshid A. Ganie, Pratap Vishnoi, Aijaz A. Dar. Utility of Bis-4-pyridines as Supramolecular Linkers for 5-
Sulfosalicylic Acid Centers: Structural and Optical Investigations. Crystal Growth & Design 2019, 19 (4) , 2289-
2297. https://doi.org/10.1021/acs.cgd.8b01914

80. Tomoyuki Ikai, Takumu Yoshida, Ken-ichi Shinohara, Tsuyoshi Taniguchi, Yuya Wada, Timothy M. Swager.
Triptycene-Based Ladder Polymers with One-Handed Helical Geometry. Journal of the American Chemical
Society 2019, 141 (11) , 4696-4703. https://doi.org/10.1021/jacs.8b13865

81. Yuree Oh, Kyoung Min Lee, Doyoung Jung, Ji Ae Chae, Hea Ji Kim, Mincheol Chang, Jong-Jin Park,
Hyungwoo Kim. Sustainable, Naringenin-Based Thermosets Show Reversible Macroscopic Shape Changes
and Enable Modular Recycling. ACS Macro Letters 2019, 8 (3) , 239-244.
https://doi.org/10.1021/acsmacrolett.9b00008

82. Soumalya Bhattacharyya, Aniket Chowdhury, Rupak Saha, Partha Sarathi Mukherjee. Multifunctional Self-
Assembled Macrocycles with Enhanced Emission and Reversible Photochromic Behavior. Inorganic Chemistry
2019, 58 (6) , 3968-3981. https://doi.org/10.1021/acs.inorgchem.9b00039

83. Salvador Pérez-Estrada, Braulio Rodríguez-Molina, Emily F. Maverick, Saeed I. Khan, Miguel A. Garcia-
Garibay. Throwing in a Monkey Wrench to Test and Determine Geared Motion in the Dynamics of a Crystalline
One-Dimensional (1D) Columnar Rotor Array. Journal of the American Chemical Society 2019, 141 (6) , 2413-
2420. https://doi.org/10.1021/jacs.8b11385

84. Tainára Orlando, Paulo R. S. Salbego, Carmen L. R. Taschetto, Helio G. Bonacorso, Nilo Zanatta, Manfredo
Hoerner, Marcos A. P. Martins. Polymorphism in a Rotaxane Molecule: Intra- and Intermolecular
Understanding. Crystal Growth & Design 2019, 19 (2) , 1021-1030. https://doi.org/10.1021/acs.cgd.8b01560

85. Martina Cirulli, Amanpreet Kaur, James E. M. Lewis, Zhihui Zhang, Jonathan A. Kitchen, Stephen M.
Goldup, Maxie M. Roessler. Rotaxane-Based Transition Metal Complexes: Effect of the Mechanical Bond on
Structure and Electronic Properties. Journal of the American Chemical Society 2019, 141 (2) , 879-889.
https://doi.org/10.1021/jacs.8b09715

86. Wei Zheng, Wei Wang, Shu-Ting Jiang, Guang Yang, Zhen Li, Xu-Qing Wang, Guang-Qiang Yin, Ying Zhang,
Hongwei Tan, Xiaopeng Li, Hongming Ding, Guosong Chen, Hai-Bo Yang. Supramolecular Transformation of
Metallacycle-linked Star Polymers Driven by Simple Phosphine Ligand-Exchange Reaction. Journal of the
American Chemical Society 2019, 141 (1) , 583-591. https://doi.org/10.1021/jacs.8b11642

87. Kai-Jen Chen, Ann Chen Tan, Chi-Hsien Wang, Ting-Shen Kuo, Pei-Lin Chen, Masaki Horie. Photoinduced
Mechanical Motions of Biferrocene-Containing Pseudorotaxane Crystals. Crystal Growth & Design 2019, 19 (1)
, 17-22. https://doi.org/10.1021/acs.cgd.8b01169

88. Henry Hess, Gadiel Saper. Engineering with Biomolecular Motors. Accounts of Chemical Research 2018,
51 (12) , 3015-3022. https://doi.org/10.1021/acs.accounts.8b00296

89. Naoki Ousaka, Kaori Shimizu, Yoshimasa Suzuki, Takuya Iwata, Manabu Itakura, Daisuke Taura, Hiroki
Iida, Yoshio Furusho, Tadashi Mori, Eiji Yashima. Spiroborate-Based Double-Stranded Helicates: Meso-to-
Racemo Isomerization and Ion-Triggered Springlike Motion of the Racemo-Helicate. Journal of the American
Chemical Society 2018, 140 (49) , 17027-17039. https://doi.org/10.1021/jacs.8b08268
90. Stefano F. Pizzolato, Peter Štacko, Jos C. M. Kistemaker, Thomas van Leeuwen, Edwin Otten, Ben L.
Feringa. Central-to-Helical-to-Axial-to-Central Transfer of Chirality with a Photoresponsive Catalyst. Journal of
the American Chemical Society 2018, 140 (49) , 17278-17289. https://doi.org/10.1021/jacs.8b10816
  

91. Shinji Ikejiri, Yoshinori Takashima, Motofumi Osaki, Hiroyasu Yamaguchi, Akira Harada. Solvent-Free
Photoresponsive Arti cial Muscles Rapidly Driven by Molecular Machines. Journal of the American Chemical
Society 2018, 140 (49) , 17308-17315. https://doi.org/10.1021/jacs.8b11351

92. Aaron Gerwien, Peter Mayer, Henry Dube. Photon-Only Molecular Motor with Reverse Temperature-
Dependent E ciency. Journal of the American Chemical Society 2018, 140 (48) , 16442-16445.
https://doi.org/10.1021/jacs.8b10660

93. Agostino Galanti, Valentin Diez-Cabanes, Jasmin Santoro, Michal Valášek, Andrea Minoia, Marcel Mayor,
Jérôme Cornil, Paolo Samorì. Electronic Decoupling in C3-Symmetrical Light-Responsive Tris(Azobenzene)
Scaffolds: Self-Assembly and Multiphotochromism. Journal of the American Chemical Society 2018, 140 (47) ,
16062-16070. https://doi.org/10.1021/jacs.8b06324

94. Zoran Kokan, Michał J. Chmielewski. A Photoswitchable Heteroditopic Ion-Pair Receptor. Journal of the
American Chemical Society 2018, 140 (47) , 16010-16014. https://doi.org/10.1021/jacs.8b08689

95. Pei Li, Ganhua Xie, Pei Liu, Xiang-Yu Kong, Yanlin Song, Liping Wen, Lei Jiang. Light-Driven ATP
Transmembrane Transport Controlled by DNA Nanomachines. Journal of the American Chemical Society
2018, 140 (47) , 16048-16052. https://doi.org/10.1021/jacs.8b10527

96. José F. Rodríguez, Katherine I. Burton, Ivan Franzoni, David A. Petrone, Ina Scheipers, Mark Lautens.
Palladium-Catalyzed Hydride Addition/C–H Bond Activation Cascade: Cycloisomerization of 1,6-Diynes.
Organic Letters 2018, 20 (21) , 6915-6919. https://doi.org/10.1021/acs.orglett.8b03057

97. Miyako Tsurunaga, Yusuke Inagaki, Hiroyuki Momma, Eunsang Kwon, Kentaro Yamaguchi, Kenji Yoza,
Wataru Setaka. Dielectric Relaxation of Powdered Molecular Gyrotops Having a Thiophene Dioxide-diyl as a
Dipolar Rotor. Organic Letters 2018, 20 (21) , 6934-6937. https://doi.org/10.1021/acs.orglett.8b03087

98. Mark J. Moran, Mitchell Magrini, David M. Walba, Ivan Aprahamian. Driving a Liquid Crystal Phase
Transition Using a Photochromic Hydrazone. Journal of the American Chemical Society 2018, 140 (42) ,
13623-13627. https://doi.org/10.1021/jacs.8b09622

99. Yansong Ren, Sheng Xie, Erik Svensson Grape, A. Ken Inge, Olof Ramström. Multistimuli-Responsive
Enaminitrile Molecular Switches Displaying H+-Induced Aggregate Emission, Metal Ion-Induced Turn-On
Fluorescence, and Organogelation Properties. Journal of the American Chemical Society 2018, 140 (42) ,
13640-13643. https://doi.org/10.1021/jacs.8b09843

100. Reguram Arumugaperumal, Putikam Raghunath, Ming-Chang Lin, Wen-Sheng Chung. Distinct
Nanostructures and Organogel Driven by Reversible Molecular Switching of a Tetraphenylethene-Involved
Calix[4]arene-Based Amphiphilic [2]Rotaxane. Chemistry of Materials 2018, 30 (20) , 7221-7233.
https://doi.org/10.1021/acs.chemmater.8b03286

101. Youzhi Xu, Ramandeep Kaur, Bingzhe Wang, Martin B. Minameyer, Sebastian Gsänger, Bernd Meyer,
Thomas Drewello, Dirk M. Guldi, Max von Delius. Concave–Convex π–π Template Approach Enables the
Synthesis of [10]Cycloparaphenylene–Fullerene [2]Rotaxanes. Journal of the American Chemical Society 2018,
140 (41) , 13413-13420. https://doi.org/10.1021/jacs.8b08244
102. Huan Yao, Hua Ke, Xiaobin Zhang, San-Jiang Pan, Ming-Shuang Li, Liu-Pan Yang, Georg Schreckenbach,
Wei Jiang. Molecular Recognition of Hydrophilic Molecules in Water by Combining the Hydrophobic Effect with
Hydrogen Bonding. Journal of the American Chemical Society 2018, 140 (41) , 13466-13477.
  
https://doi.org/10.1021/jacs.8b09157

103. Anouk S. Lubbe, Christian Böhmer, Filippo Tosi, Wiktor Szymanski, Ben L. Feringa. Molecular Motors in
Aqueous Environment. The Journal of Organic Chemistry 2018, 83 (18) , 11008-11018.
https://doi.org/10.1021/acs.joc.8b01627

104. Michele Bruschini, Gianfranco Ercolani, Stefano Gallina, Paolo Mencarelli. Record Rate Enhancements
for Tetrathiafulvalene Guests in the Formation of Bipyridinium- and Diazapyrenium-Based [2]Pseudorotaxanes.
The Journal of Organic Chemistry 2018, 83 (18) , 11446-11449. https://doi.org/10.1021/acs.joc.8b01786

105. Jizhuang Wang, Ze Xiong, Jing Zheng, Xiaojun Zhan, Jinyao Tang. Light-Driven Micro/Nanomotor for
Promising Biomedical Tools: Principle, Challenge, and Prospect. Accounts of Chemical Research 2018, 51 (9) ,
1957-1965. https://doi.org/10.1021/acs.accounts.8b00254

106. Lindsay J. Schafer, Kevin J. Garcia, Andrew W. Baggett, Taylor M. Lord, Peter M. Findeis, Robert D. Pike,
Robert A. Stockland, Jr.. Synthesis of Spirocyclic Diphosphite-Supported Gold Metallomacrocycles via a
Protodeauration/Cyclization Strategy: Mechanistic and Binding Studies. Inorganic Chemistry 2018, 57 (18) ,
11662-11672. https://doi.org/10.1021/acs.inorgchem.8b01805

107. Ming Zhang, Yong-Guang Jia, Lingyan Liu, Jing Li, X. X. Zhu. Soluble–Insoluble–Soluble Transitions of
Thermoresponsive Cryptand-Containing Graft Copolymers. ACS Omega 2018, 3 (8) , 10172-10179.
https://doi.org/10.1021/acsomega.8b01308

108. Jens Kügel, Markus Leisegang, Matthias Bode. Imprinting Directionality into Proton Transfer Reactions
of an Achiral Molecule. ACS Nano 2018, 12 (8) , 8733-8738. https://doi.org/10.1021/acsnano.8b04868

109. Rashmi J. Das, Kingsuk Mahata. Synthesis, Photophysical, Electrochemical, and Halochromic Properties
of peri-Naphthoindigo. Organic Letters 2018, 20 (16) , 5027-5031.
https://doi.org/10.1021/acs.orglett.8b02178

110. Carlos Romero-Muñiz, Denís Paredes-Roibás, Concepción López, Antonio Hernanz, José María Gavira-
Vallejo. Assignment of the Raman Spectrum of Benzylic Amide [2]Catenane: Raman Microscopy Experiments
and First-Principles Calculations. The Journal of Physical Chemistry C 2018, 122 (31) , 18102-18109.
https://doi.org/10.1021/acs.jpcc.8b04904

111. Ruizi Peng, Xiaofang Zheng, Yifan Lyu, Liujun Xu, Xiaobing Zhang, Guoliang Ke, Qiaoling Liu, Changjun
You, Shuangyan Huan, Weihong Tan. Engineering a 3D DNA-Logic Gate Nanomachine for Bispeci c
Recognition and Computing on Target Cell Surfaces. Journal of the American Chemical Society 2018, 140 (31)
, 9793-9796. https://doi.org/10.1021/jacs.8b04319

112. Lais S. Matos, Ronaldo C. Amaral, Neyde Y. Murakami Iha. Visible Photosensitization of trans-
Styrylpyridine Coordinated to fac-[Re(CO)3(dcbH2)]+: New Insights. Inorganic Chemistry 2018, 57 (15) , 9316-
9326. https://doi.org/10.1021/acs.inorgchem.8b01304

113. Ronaldo C. Amaral, Lais S. Matos, Kassio P. S. Zanoni, Neyde Y. Murakami Iha. Photoreversible
Molecular Motion of stpyCN Coordinated to fac-[Re(CO)3(NN)]+ Complexes. The Journal of Physical
Chemistry A 2018, 122 (29) , 6071-6080. https://doi.org/10.1021/acs.jpca.8b02630
114. Paramjyothi C. Nandajan, Hyeong-Ju Kim, Santiago Casado, Soo Young Park, Johannes Gierschner.
Insight into Water-Soluble Highly Fluorescent Low-Dimensional Host–Guest Supramolecular Polymers:
  
Structure and Energy-Transfer Dynamics Revealed by Polarized Fluorescence Spectroscopy. The Journal of
Physical Chemistry Letters 2018, 9 (14) , 3870-3877. https://doi.org/10.1021/acs.jpclett.8b01562

115. Ruth Dorel, Carla Miró, Yuchen Wei, Sander J. Wezenberg, Ben L. Feringa. Cation-Modulated Rotary
Speed in a Light-Driven Crown Ether Functionalized Molecular Motor. Organic Letters 2018, 20 (13) , 3715-
3718. https://doi.org/10.1021/acs.orglett.8b00969

116. Yoshinori Takashima, Yuki Hayashi, Motofumi Osaki, Fumitoshi Kaneko, Hiroyasu Yamaguchi, Akira
Harada. A Photoresponsive Polymeric Actuator Topologically Cross-Linked by Movable Units Based on a
[2]Rotaxane. Macromolecules 2018, 51 (12) , 4688-4693. https://doi.org/10.1021/acs.macromol.8b00939

117. Sergey Simonov, Leokadiya Zorina, Pawel Wzietek, Antonio Rodríguez-Fortea, Enric Canadell, Cécile
Mézière, Guillaume Bastien, Cyprien Lemouchi, Miguel A. Garcia-Garibay, Patrick Batail. Static Modulation
Wave of Arrays of Halogen Interactions Transduced to a Hierarchy of Nanoscale Change Stimuli of Crystalline
Rotors Dynamics. Nano Letters 2018, 18 (6) , 3780-3784. https://doi.org/10.1021/acs.nanolett.8b00956

118. Ren-Tsung Wu, Xiaodong Chi, Takehiro Hirao, Vincent M. Lynch, Jonathan L. Sessler. Supramolecular
Properties of a Monocarboxylic Acid-Functionalized “Texas-Sized” Molecular Box. Journal of the American
Chemical Society 2018, 140 (22) , 6823-6831. https://doi.org/10.1021/jacs.7b12957

119. Jiawen Chen, Jérôme Vachon, Ben L. Feringa. Design, Synthesis, and Isomerization Studies of Light-
Driven Molecular Motors for Single Molecular Imaging. The Journal of Organic Chemistry 2018, 83 (11) , 6025-
6034. https://doi.org/10.1021/acs.joc.8b00654

120. Hang Yin, Roselyne Rosas, Didier Gigmes, Olivier Ouari, Ruibing Wang, Anthony Kermagoret, David
Bardelang. Metal Actuated Ring Translocation Switches in Water. Organic Letters 2018, 20 (11) , 3187-3191.
https://doi.org/10.1021/acs.orglett.8b01019

121. Tania Neva, Thais Carmona, Juan M. Benito, Cédric Przybylski, Carmen Ortiz Mellet, Francisco
Mendicuti, José M. García Fernández. Xylylene Clips for the Topology-Guided Control of the Inclusion and Self-
Assembling Properties of Cyclodextrins. The Journal of Organic Chemistry 2018, 83 (10) , 5588-5597.
https://doi.org/10.1021/acs.joc.8b00602

122. Christian Petermayer, Henry Dube. Indigoid Photoswitches: Visible Light Responsive Molecular Tools.
Accounts of Chemical Research 2018, 51 (5) , 1153-1163. https://doi.org/10.1021/acs.accounts.7b00638

123. Kazuto Takaishi, Makoto Yasui, Tadashi Ema. Binaphthyl–Bipyridyl Cyclic Dyads as a Chiroptical Switch.
Journal of the American Chemical Society 2018, 140 (16) , 5334-5338.
https://doi.org/10.1021/jacs.8b01860

124. Shunsuke Kuwahara, Yuri Suzuki, Naoya Sugita, Mari Ikeda, Fumi Nagatsugi, Nobuyuki Harada, Yoichi
Habata. Thermal E/Z Isomerization in First Generation Molecular Motors. The Journal of Organic Chemistry
2018, 83 (8) , 4800-4804. https://doi.org/10.1021/acs.joc.7b03264

125. Roland Wilcken, Monika Schildhauer, Florian Rott, Ludwig Alexander Huber, Manuel Guentner, Stefan
Thumser, Kerstin Hoffmann, Sven Oesterling, Regina de Vivie-Riedle, Eberhard Riedle, Henry Dube. Complete
Mechanism of Hemithioindigo Motor Rotation. Journal of the American Chemical Society 2018, 140 (15) ,
5311-5318. https://doi.org/10.1021/jacs.8b02349
126. James E. M. Lewis, Florian Modicom, Stephen M. Goldup. E cient Multicomponent Active Template
Synthesis of Catenanes. Journal of the American Chemical Society 2018, 140 (14) , 4787-4791.
  
https://doi.org/10.1021/jacs.8b01602

127. Toru Amaya, Hayato Fujimoto, Takahiro Tanaka, Toshiyuki Moriuchi. Synthesis and Isomerization
Behavior of a Macrocycle with Four Photoresponsive Moieties. Organic Letters 2018, 20 (7) , 2055-2058.
https://doi.org/10.1021/acs.orglett.8b00598

128. Nikita Mittal, Merve S. Özer, Michael Schmittel. Four-Component Catalytic Machinery: Reversible Three-
State Control of Organocatalysis by Walking Back and Forth on a Track. Inorganic Chemistry 2018, 57 (7) ,
3579-3586. https://doi.org/10.1021/acs.inorgchem.7b02703

129. Paul M. Bogie, Lauren R. Holloway, Yana Lyon, Nicole C. Onishi, Gregory J. O. Beran, Ryan R. Julian,
Richard J. Hooley. A Springloaded Metal-Ligand Mesocate Allows Access to Trapped Intermediates of Self-
Assembly. Inorganic Chemistry 2018, 57 (7) , 4155-4163. https://doi.org/10.1021/acs.inorgchem.8b00370

130. Richard J. Hooley. Rings and Things: The Magic of Building Self-Assembled Cages and Macrocycles.
Inorganic Chemistry 2018, 57 (7) , 3497-3499. https://doi.org/10.1021/acs.inorgchem.8b00553

131. Jinqiao Dong, Xu Li, Kang Zhang, Yi Di Yuan, Yuxiang Wang, Linzhi Zhai, Guoliang Liu, Daqiang Yuan,
Jianwen Jiang, Dan Zhao. Con nement of Aggregation-Induced Emission Molecular Rotors in Ultrathin Two-
Dimensional Porous Organic Nanosheets for Enhanced Molecular Recognition. Journal of the American
Chemical Society 2018, 140 (11) , 4035-4046. https://doi.org/10.1021/jacs.7b13069

132. Ian G. Powers, John M. Andjaba, Xuyi Luo, Jianguo Mei, Christopher Uyeda. Catalytic Azoarene
Synthesis from Aryl Azides Enabled by a Dinuclear Ni Complex. Journal of the American Chemical Society
2018, 140 (11) , 4110-4118. https://doi.org/10.1021/jacs.8b00503

133. Tomoyuki Ikai, Takumu Yoshida, Seiya Awata, Yuya Wada, Katsuhiro Maeda, Motohiro Mizuno, Timothy
M. Swager. Circularly Polarized Luminescent Triptycene-Based Polymers. ACS Macro Letters 2018, 7 (3) , 364-
369. https://doi.org/10.1021/acsmacrolett.8b00106

134. Xiao-Dong Yang, Rui Zhu, Li Sun, Rui-Yun Guo, Jie Zhang. Phototriggered Mechanical Movement in A
Bipyridinium-based Coordination Polymer Powered by Electron Transfer. Inorganic Chemistry 2018, 57 (5) ,
2724-2729. https://doi.org/10.1021/acs.inorgchem.7b03108

135. Amit Ghosh, Indrajit Paul, Matthias Adlung, Claudia Wickleder, and Michael Schmittel . Oscillating
Emission of [2]Rotaxane Driven by Chemical Fuel. Organic Letters 2018, 20 (4) , 1046-1049.
https://doi.org/10.1021/acs.orglett.7b03996

136. Yoshimitsu Sagara, Marc Karman, Ester Verde-Sesto, Kazuya Matsuo, Yuna Kim, Nobuyuki Tamaoki, and
Christoph Weder . Rotaxanes as Mechanochromic Fluorescent Force Transducers in Polymers. Journal of the
American Chemical Society 2018, 140 (5) , 1584-1587. https://doi.org/10.1021/jacs.7b12405

137. Timothy A. Barendt, Ilija Rašović, Maria A. Lebedeva, George A. Farrow, Alexander Auty, Dimitri
Chekulaev, Igor V. Sazanovich, Julia A. Weinstein, Kyriakos Porfyrakis, and Paul D. Beer . Anion-Mediated
Photophysical Behavior in a C60 Fullerene [3]Rotaxane Shuttle. Journal of the American Chemical Society
2018, 140 (5) , 1924-1936. https://doi.org/10.1021/jacs.7b12819

138. Michael M. Lerch, Miroslav Medved′, Andrea Lapini, Adèle D. Laurent, Alessandro Iagatti, Laura Bussotti,
Wiktor Szymański, Wybren Jan Buma, Paolo Foggi, Mariangela Di Donato, and Ben L. Feringa . Tailoring
Photoisomerization Pathways in Donor–Acceptor Stenhouse Adducts: The Role of the Hydroxy Group. The
Journal of Physical Chemistry A 2018, 122 (4) , 955-964. https://doi.org/10.1021/acs.jpca.7b10255
  
139. Antonio Rodríguez-Fortea, Jiří Kaleta, Cécile Mézière, Magali Allain, Enric Canadell, Pawel Wzietek, Josef
Michl, and Patrick Batail . Asymmetric Choreography in Pairs of Orthogonal Rotors. ACS Omega 2018, 3 (1) ,
1293-1297. https://doi.org/10.1021/acsomega.7b01580

140. Ning-Ning Yang, Jia-Jia Fang, Qi Sui, and En-Qing Gao . Incorporating Electron-De cient Bipyridinium
Chromorphores to Make Multiresponsive Metal–organic Frameworks. ACS Applied Materials & Interfaces
2018, 10 (3) , 2735-2744. https://doi.org/10.1021/acsami.7b17381

141. Oleksandr Stetsovych, Pingo Mutombo, Martin Švec, Michal Šámal, Jindřich Nejedlý, Ivana Císařová,
Héctor Vázquez, María Moro-Lagares, Jan Berger, Jaroslav Vacek, Irena G. Stará, Ivo Starý, and Pavel Jelínek .
Large Converse Piezoelectric Effect Measured on a Single Molecule on a Metallic Surface. Journal of the
American Chemical Society 2018, 140 (3) , 940-946. https://doi.org/10.1021/jacs.7b08729

142. Guocan Yu, Bryant C. Yung, Zijian Zhou, Zhengwei Mao, and Xiaoyuan Chen . Arti cial Molecular
Machines in Nanotheranostics. ACS Nano 2018, 12 (1) , 7-12. https://doi.org/10.1021/acsnano.7b07851

143. Shao-Chi Cheng, Kai-Jen Chen, Yuji Suzaki, Yoshitaka Tsuchido, Ting-Shen Kuo, Kohtaro Osakada, and
Masaki Horie . Reversible Laser-Induced Bending of Pseudorotaxane Crystals. Journal of the American
Chemical Society 2018, 140 (1) , 90-93. https://doi.org/10.1021/jacs.7b10998

144. Anna Walczak and Artur R. Stefankiewicz . pH-Induced Linkage Isomerism of Pd(II) Complexes: A
Pathway to Air- and Water-Stable Suzuki–Miyaura-Reaction Catalysts. Inorganic Chemistry 2018, 57 (1) , 471-
477. https://doi.org/10.1021/acs.inorgchem.7b02711

145. Emmanuel Di Piazza, Corentin Boilleau, Antoine Vacher, Khalissa Merahi, Lucie Norel, Karine Costuas,
Thierry Roisnel, Sylvie Choua, Philippe Turek, and Stéphane Rigaut . Ruthenium Carbon-Rich Group as a Redox-
Switchable Metal Coupling Unit in Linear Trinuclear Complexes. Inorganic Chemistry 2017, 56 (23) , 14540-
14555. https://doi.org/10.1021/acs.inorgchem.7b02288

146. Fu Huang, Guangxia Wang, Lishuang Ma, Ying Wang, Xuebo Chen, Yanke Che, and Hua Jiang . Molecular
Spur Gears Based on a Switchable Quinquepyridine Foldamer Acting as a Stator. The Journal of Organic
Chemistry 2017, 82 (23) , 12106-12111. https://doi.org/10.1021/acs.joc.7b01864

147. Laurianne Wojcik, François Michaud, Sébastien Gauthier, Nolwenn Cabon, Pascal Le Poul, Frederic
Gloaguen, and Nicolas Le Poul . Reversible Redox Switching of Chromophoric Phenylmethylenepyrans by
Carbon–Carbon Bond Making/Breaking. The Journal of Organic Chemistry 2017, 82 (23) , 12395-12405.
https://doi.org/10.1021/acs.joc.7b02199

148. Hitoshi Ube, Ryo Yamada, Jun-ichi Ishida, Hiroyasu Sato, Motoo Shiro, and Mitsuhiko Shionoya . A
Circularly Arranged Sextuple Triptycene Gear Molecule. Journal of the American Chemical Society 2017, 139
(46) , 16470-16473. https://doi.org/10.1021/jacs.7b09439

149. Shuangshuang Wang, Xueguang Shao, and Wensheng Cai . Solvent and Structure Effects on the
Shuttling in Pillar[5]arene/Triazole Rotaxanes. The Journal of Physical Chemistry C 2017, 121 (45) , 25547-
25553. https://doi.org/10.1021/acs.jpcc.7b07279

150. Mariangela Di Donato, Michael M. Lerch, Andrea Lapini, Adèle D. Laurent, Alessandro Iagatti, Laura
Bussotti, Svante P. Ihrig, Miroslav Medved’, Denis Jacquemin, Wiktor Szymański, Wybren Jan Buma, Paolo
Foggi, and Ben L. Feringa . Shedding Light on the Photoisomerization Pathway of Donor–Acceptor Stenhouse
Adducts. Journal of the American Chemical Society 2017, 139 (44) , 15596-15599.
https://doi.org/10.1021/jacs.7b09081   

151. Zarina M. Bikbaeva, Daniil M. Ivanov, Alexander S. Novikov, Ivan V. Ananyev, Nadezhda A. Bokach, and
Vadim Yu. Kukushkin . Electrophilic–Nucleophilic Dualism of Nickel(II) toward Ni···I Noncovalent Interactions:
Semicoordination of Iodine Centers via Electron Belt and Halogen Bonding via σ-Hole. Inorganic Chemistry
2017, 56 (21) , 13562-13578. https://doi.org/10.1021/acs.inorgchem.7b02224

152. Qi Sui, Ning-Ning Yang, Teng Gong, Peng Li, Ye Yuan, En-Qing Gao, Lin Wang. Impact of Lattice Water on
Solid-State Electron Transfer in Viologen Pseudopolymorphs: Modulation of Photo- and Piezochromism. The
Journal of Physical Chemistry Letters 2017, 8 (21) , 5450-5455. https://doi.org/10.1021/acs.jpclett.7b02452

153. Fan Hong, Fei Zhang, Yan Liu, and Hao Yan . DNA Origami: Scaffolds for Creating Higher Order
Structures. Chemical Reviews 2017, 117 (20) , 12584-12640. https://doi.org/10.1021/acs.chemrev.6b00825

154. Antoine Goujon, Thomas Lang, Giacomo Mariani, Emilie Moulin, Gad Fuks, Jesus Raya, Eric Buhler,
Nicolas Giuseppone. Bistable [c2] Daisy Chain Rotaxanes as Reversible Muscle-like Actuators in Mechanically
Active Gels. Journal of the American Chemical Society 2017, 139 (42) , 14825-14828.
https://doi.org/10.1021/jacs.7b06710

155. Elton Oyarzua, Jens Honore Walther, Constantine M Megaridis, Petros Koumoutsakos, and Harvey A.
Zambrano . Carbon Nanotubes as Thermally Induced Water Pumps. ACS Nano 2017, 11 (10) , 9997-10002.
https://doi.org/10.1021/acsnano.7b04177

156. Elizabeth Elacqua, Xiaolong Zheng, and Marcus Weck . Light-Mediated Reversible Assembly of
Polymeric Colloids. ACS Macro Letters 2017, 6 (10) , 1060-1065.
https://doi.org/10.1021/acsmacrolett.7b00539

157. Ting Wang, Cheng Wang, Shuai Zhou, JianHua Xu, Wei Jiang, LingHua Tan, and JiaJun Fu . Nanovalves-
Based Bacteria-Triggered, Self-Defensive Antibacterial Coating: Using Combination Therapy, Dual Stimuli-
Responsiveness, and Multiple Release Modes for Treatment of Implant-Associated Infections. Chemistry of
Materials 2017, 29 (19) , 8325-8337. https://doi.org/10.1021/acs.chemmater.7b02678

158. Yuping Wang, Marco Frasconi, and J. Fraser Stoddart . Introducing Stable Radicals into Molecular
Machines. ACS Central Science 2017, 3 (9) , 927-935. https://doi.org/10.1021/acscentsci.7b00219

159. Dustin Chen and Qibing Pei . Electronic Muscles and Skins: A Review of Soft Sensors and Actuators.
Chemical Reviews 2017, 117 (17) , 11239-11268. https://doi.org/10.1021/acs.chemrev.7b00019

160. Ivica Cvrtila, Hugo Fanlo-Virgós, Gaël Schaeffer, Guillermo Monreal Santiago, and Sijbren Otto . Redox
Control over Acyl Hydrazone Photoswitches. Journal of the American Chemical Society 2017, 139 (36) ,
12459-12465. https://doi.org/10.1021/jacs.7b03724

161. Ruizi Peng, Huijing Wang, Yifan Lyu, Liujun Xu, Hui Liu, Hailan Kuai, Qiaoling Liu, and Weihong Tan .
Facile Assembly/Disassembly of DNA Nanostructures Anchored on Cell-Mimicking Giant Vesicles. Journal of
the American Chemical Society 2017, 139 (36) , 12410-12413. https://doi.org/10.1021/jacs.7b07485

162. Daut R. Islamov, Valery G. Shtyrlin, Nikita Yu. Serov, Ivan V. Fedyanin, and Konstantin A. Lyssenko .
Symmetry In uence on the Rotation of Molecules in Crystals. Crystal Growth & Design 2017, 17 (9) , 4703-
4709. https://doi.org/10.1021/acs.cgd.7b00588
163. Arvin Sain Tanwar and Parameswar Krishnan Iyer . Fluorescence “Turn-On” Indicator Displacement
and
Assay-Based Sensing of Nitroexplosive 2,4,6-Trinitrophenol in Aqueous Media via a Polyelectrolyte Dye

Complex. ACS Omega 2017, 2 (8) , 4424-4430. https://doi.org/10.1021/acsomega.7b00765

164. Christopher J. Martin, Alan T. L. Lee, Ralph W. Adams, and David A. Leigh . Enzyme-Mediated Directional
Transport of a Small-Molecule Walker With Chemically Identical Feet. Journal of the American Chemical
Society 2017, 139 (34) , 11998-12002. https://doi.org/10.1021/jacs.7b06503

165. Chuhan Zong, Michelle J. Wu, Jason Z. Qin, and A. James Link . Lasso Peptide Benenodin-1 Is a
Thermally Actuated [1]Rotaxane Switch. Journal of the American Chemical Society 2017, 139 (30) , 10403-
10409. https://doi.org/10.1021/jacs.7b04830

166. Jiří Kaleta, Jiawen Chen, Guillaume Bastien, Martin Dračínský, Milan Mašát, Charles T. Rogers, Ben L.
Feringa, and Josef Michl . Surface Inclusion of Unidirectional Molecular Motors in Hexagonal Tris(o-
phenylene)cyclotriphosphazene. Journal of the American Chemical Society 2017, 139 (30) , 10486-10498.
https://doi.org/10.1021/jacs.7b05404

167. Chuan Gao, Zhou-Lin Luan, Qi Zhang, Si-Jia Rao, Da-Hui Qu, and He Tian . A Braided Hetero[2]
(3)rotaxane. Organic Letters 2017, 19 (14) , 3931-3934. https://doi.org/10.1021/acs.orglett.7b01853

168. Abdulselam Adam and Gebhard Haberhauer . Switching Process Consisting of Three Isomeric States of
an Azobenzene Unit. Journal of the American Chemical Society 2017, 139 (28) , 9708-9713.
https://doi.org/10.1021/jacs.7b05316

169. Benjamin Doistau, Lorien Benda, Jean-Louis Cantin, Lise-Marie Chamoreau, Eliseo Ruiz, Valérie Marvaud,
Bernold Hasenknopf, and Guillaume Vives . Six States Switching of Redox-Active Molecular Tweezers by Three
Orthogonal Stimuli. Journal of the American Chemical Society 2017, 139 (27) , 9213-9220.
https://doi.org/10.1021/jacs.7b02945

170. Xiang Wang, Barbara Wicher, Yann Ferrand, and Ivan Huc . Orchestrating Directional Molecular Motions:
Kinetically Controlled Supramolecular Pathways of a Helical Host on Rodlike Guests. Journal of the American
Chemical Society 2017, 139 (27) , 9350-9358. https://doi.org/10.1021/jacs.7b04884

171. Katarzyna Eichstaedt, Javier Jaramillo-Garcia, David A. Leigh, Vanesa Marcos, Simone Pisano, and
Thomas A. Singleton . Switching between Anion-Binding Catalysis and Aminocatalysis with a Rotaxane Dual-
Function Catalyst. Journal of the American Chemical Society 2017, 139 (27) , 9376-9381.
https://doi.org/10.1021/jacs.7b04955

172. Timothy A. Barendt, Liliana Ferreira, Igor Marques, Vítor Félix, and Paul D. Beer . Anion- and Solvent-
Induced Rotary Dynamics and Sensing in a Perylene Diimide [3]Catenane. Journal of the American Chemical
Society 2017, 139 (26) , 9026-9037. https://doi.org/10.1021/jacs.7b04295

173. Wang Li, Chun-Ting He, Ying Zeng, Cheng-Min Ji, Zi-Yi Du, Wei-Xiong Zhang, and Xiao-Ming Chen .
Crystalline Supramolecular Gyroscope with a Water Molecule as an Ultrasmall Polar Rotator Modulated by
Charge-Assisted Hydrogen Bonds. Journal of the American Chemical Society 2017, 139 (24) , 8086-8089.
https://doi.org/10.1021/jacs.7b02981

174. Qiang Shi and Chuan-Feng Chen . Switchable Complexation between (O-Methyl)6-2,6-helic[6]arene and
Protonated Pyridinium Salts Controlled by Acid/Base and Photoacid. Organic Letters 2017, 19 (12) , 3175-
3178. https://doi.org/10.1021/acs.orglett.7b01296
175. Lara S. R. Small, Marc Bruning, Andrew R. Thomson, Aimee L. Boyle, Roberta B. Davies, Paul M. G.
 of a
Curmi, Nancy R. Forde, Heiner Linke, Derek N. Woolfson, and Elizabeth H. C. Bromley . Construction 
Chassis for a Tripartite Protein-Based Molecular Motor. ACS Synthetic Biology 2017, 6 (6) , 1096-1102.
https://doi.org/10.1021/acssynbio.7b00037

176. Krishna Kanti Dey and Ayusman Sen . Chemically Propelled Molecules and Machines. Journal of the
American Chemical Society 2017, 139 (23) , 7666-7676. https://doi.org/10.1021/jacs.7b02347

177. Jae Woo Park and Toru Shiozaki . Analytical Derivative Coupling for Multistate CASPT2 Theory. Journal
of Chemical Theory and Computation 2017, 13 (6) , 2561-2570. https://doi.org/10.1021/acs.jctc.7b00018

178. Pronay Kumar Biswas, Suchismita Saha, Yerramsetti Nanaji, Anup Rana, and Michael Schmittel .
In uence of Rotator Design on the Speed of Self-Assembled Four-Component Nanorotors: Coordinative
Versus Dispersive Interactions. Inorganic Chemistry 2017, 56 (11) , 6662-6670.
https://doi.org/10.1021/acs.inorgchem.7b00740

179. Liu-Pan Yang, Fei Jia, Jie-Shun Cui, Song-Bo Lu, and Wei Jiang . Light-Controlled Switching of a Non-
photoresponsive Molecular Shuttle. Organic Letters 2017, 19 (11) , 2945-2948.
https://doi.org/10.1021/acs.orglett.7b01184

180. Giuseppe Romanazzi, Leonardo Degennaro, Piero Mastrorilli, and Renzo Luisi . Chiral Switchable
Catalysts for Dynamic Control of Enantioselectivity. ACS Catalysis 2017, 7 (6) , 4100-4114.
https://doi.org/10.1021/acscatal.7b00827

181. Bartosz Krajnik, Jiawen Chen, Matthew A. Watson, Scott L. Cockroft, Ben L. Feringa, and Johan Hofkens
. Defocused Imaging of UV-Driven Surface-Bound Molecular Motors. Journal of the American Chemical
Society 2017, 139 (21) , 7156-7159. https://doi.org/10.1021/jacs.7b02758

182. David A. Leigh, Vanesa Marcos, Tugrul Nalbantoglu, Iñigo J. Vitorica-Yrezabal, F. Tuba Yasar, and
Xiaokang Zhu . Pyridyl-Acyl Hydrazone Rotaxanes and Molecular Shuttles. Journal of the American Chemical
Society 2017, 139 (20) , 7104-7109. https://doi.org/10.1021/jacs.7b03307

183. Margherita De Rosa, Carmen Talotta, Carmine Gaeta, Annunziata Soriente, Placido Neri, Sebastiano
Pappalardo, Giuseppe Gattuso, Anna Notti, Melchiorre F. Parisi, and Ilenia Pisagatti . Calix[5]arene Through-the-
Annulus Threading of Dialkylammonium Guests Weakly Paired to the TFPB Anion. The Journal of Organic
Chemistry 2017, 82 (10) , 5162-5168. https://doi.org/10.1021/acs.joc.7b00406

184. Thomas van Leeuwen, Wojciech Danowski, Edwin Otten, Sander J. Wezenberg, and Ben L. Feringa .
Asymmetric Synthesis of Second-Generation Light-Driven Molecular Motors. The Journal of Organic
Chemistry 2017, 82 (10) , 5027-5033. https://doi.org/10.1021/acs.joc.7b00852

185. Josef M. Maier, Ping Li, Erik C. Vik, Christopher J. Yehl, Sharon M. S. Strickland, and Ken D. Shimizu .
Measurement of Solvent OH−π Interactions Using a Molecular Balance. Journal of the American Chemical
Society 2017, 139 (19) , 6550-6553. https://doi.org/10.1021/jacs.7b02349

186. Toma E. Tomov, Roman Tsukanov, Yair Glick, Yaron Berger, Miran Liber, Dorit Avrahami, Doron Gerber,
and Eyal Nir . DNA Bipedal Motor Achieves a Large Number of Steps Due to Operation Using Micro uidics-
Based Interface. ACS Nano 2017, 11 (4) , 4002-4008. https://doi.org/10.1021/acsnano.7b00547
187. Denhy Hernández-Melo, Ruy Cervantes, and Jorge Tiburcio . Shuttling Motion in a Host–Guest Complex
Triggered by Spiropyran to Merocyanine Reversible Chemical Transformation. The Journal of Organic
  
Chemistry 2017, 82 (8) , 4484-4488. https://doi.org/10.1021/acs.joc.7b00509

188. Chuan Gao, Zhou-Lin Luan, Qi Zhang, Shun Yang, Si-Jia Rao, Da-Hui Qu, and He Tian . Triggering a
[2]Rotaxane Molecular Shuttle by a Photochemical Bond-Cleavage Strategy. Organic Letters 2017, 19 (7) ,
1618-1621. https://doi.org/10.1021/acs.orglett.7b00393

189. Qingqing Su, Yuanying Li, Bin Wang, Minjuan Liu, Hongjuan Wang, Wenliang Wang, and Fengyi Liu .
Combining the Advantages of Alkene and Azo E–Z Photoisomerizations: Mechanistic Insights into Ketoimine
Photoswitches. The Journal of Physical Chemistry A 2017, 121 (13) , 2588-2596.
https://doi.org/10.1021/acs.jpca.7b01674

190. Antoine Goujon, Giacomo Mariani, Thomas Lang, Emilie Moulin, Michel Rawiso, Eric Buhler, Nicolas
Giuseppone. Controlled Sol–Gel Transitions by Actuating Molecular Machine Based Supramolecular Polymers.
Journal of the American Chemical Society 2017, 139 (13) , 4923-4928.
https://doi.org/10.1021/jacs.7b00983

191. Nikita Mittal, Susnata Pramanik, Indrajit Paul, Soumen De, and Michael Schmittel . Networking
Nanoswitches for ON/OFF Control of Catalysis. Journal of the American Chemical Society 2017, 139 (12) ,
4270-4273. https://doi.org/10.1021/jacs.6b12951

192. Dawei Zhang, Alexandre Martinez, and Jean-Pierre Dutasta . Emergence of Hemicryptophanes: From
Synthesis to Applications for Recognition, Molecular Machines, and Supramolecular Catalysis. Chemical
Reviews 2017, 117 (6) , 4900-4942. https://doi.org/10.1021/acs.chemrev.6b00847

193. Thomas van Leeuwen, Jasper Pol, Diederik Roke, Sander J. Wezenberg, and Ben L. Feringa . Visible-Light
Excitation of a Molecular Motor with an Extended Aromatic Core. Organic Letters 2017, 19 (6) , 1402-1405.
https://doi.org/10.1021/acs.orglett.7b00317

194. Alexandra Vuillamy, Soumaila Zebret, Céline Besnard, Virginie Placide, Stéphane Petoud, and Josef
Hamacek . Functionalized Triptycene-Derived Tripodal Ligands: Privileged Formation of Tetranuclear Cage
Assemblies with Larger Ln(III). Inorganic Chemistry 2017, 56 (5) , 2742-2749.
https://doi.org/10.1021/acs.inorgchem.6b02900

195. Xiaojuan Pang, Xueyan Cui, Deping Hu, Chenwei Jiang, Di Zhao, Zhenggang Lan, and Fuli Li . “Watching”
the Dark State in Ultrafast Nonadiabatic Photoisomerization Process of a Light-Driven Molecular Rotary Motor.
The Journal of Physical Chemistry A 2017, 121 (6) , 1240-1249. https://doi.org/10.1021/acs.jpca.6b12253

196. Teng Gong, Xiao Yang, Jia-Jia Fang, Qi Sui, Fu-Gui Xi, and En-Qing Gao . Distinct Chromic and Magnetic
Properties of Metal–Organic Frameworks with a Redox Ligand. ACS Applied Materials & Interfaces 2017, 9 (6)
, 5503-5512. https://doi.org/10.1021/acsami.6b15540

197. Stefano Di Stefano and Gianfranco Ercolani . Statistical Ring Catenation under Thermodynamic Control:
Should the Jacobson–Stockmayer Cyclization Theory Take into Account Catenane Formation?. The Journal of
Physical Chemistry B 2017, 121 (3) , 649-656. https://doi.org/10.1021/acs.jpcb.6b12323

198. Aaron J. Teator, Huiling Shao, Gang Lu, Peng Liu, and Christopher W. Bielawski . A Photoswitchable
Ole n Metathesis Catalyst. Organometallics 2017, 36 (2) , 490-497.
https://doi.org/10.1021/acs.organomet.6b00913
199. Luca Catalano, Salvador Perez-Estrada, Hsin-Hua Wang, Anoklase J.-L. Ayitou, Saeed I. Khan, Giancarlo
Terraneo, Pierangelo Metrangolo, Stuart Brown, and Miguel A. Garcia-Garibay . Rotational Dynamics of
  
Diazabicyclo[2.2.2]octane in Isomorphous Halogen-Bonded Co-crystals: Entropic and Enthalpic Effects.
Journal of the American Chemical Society 2017, 139 (2) , 843-848. https://doi.org/10.1021/jacs.6b10780

200. Sudhakar Gaikwad and Michael Schmittel . Five-State Rotary Nanoswitch. The Journal of Organic
Chemistry 2017, 82 (1) , 343-352. https://doi.org/10.1021/acs.joc.6b02436

201. Arnaud Tron, Isabelle Pianet, Alberto Martinez-Cuezva, James H. R. Tucker, Luca Pisciottani, Mateo
Alajarin, Jose Berna, and Nathan D. McClenaghan . Remote Photoregulated Ring Gliding in a [2]Rotaxane via a
Molecular Effector. Organic Letters 2017, 19 (1) , 154-157. https://doi.org/10.1021/acs.orglett.6b03457

202. Aaron S. Rury, Shayne A. Sorenson, and Jahan M. Dawlaty . Evidence of Ultrafast Charge Transfer Driven
by Coherent Lattice Vibrations. The Journal of Physical Chemistry Letters 2017, 8 (1) , 181-187.
https://doi.org/10.1021/acs.jpclett.6b02523

203. Alex Saywell, Anne Bakker, Johannes Mielke, Takashi Kumagai, Martin Wolf, Víctor García-López, Pinn-
Tsong Chiang, James M. Tour, and Leonhard Grill . Light-Induced Translation of Motorized Molecules on a
Surface. ACS Nano 2016, 10 (12) , 10945-10952. https://doi.org/10.1021/acsnano.6b05650

204. Xiaoyan He, Lucie Norel, Yves-Marie Hervault, Remi Métivier, Anthony D’Aléo, Olivier Maury, and
Stéphane Rigaut . Modulation of Eu(III) and Yb(III) Luminescence Using a DTE Photochromic Ligand. Inorganic
Chemistry 2016, 55 (24) , 12635-12643. https://doi.org/10.1021/acs.inorgchem.6b01797

205. Chengyuan Yu, Lishuang Ma, Jiaojiao He, Junfeng Xiang, Xuebin Deng, Ying Wang, Xuebo Chen, and Hua
Jiang . Flexible, Linear Chains Act as Ba es To Inhibit the Intramolecular Rotation of Molecular Turnstiles.
Journal of the American Chemical Society 2016, 138 (49) , 15849-15852.
https://doi.org/10.1021/jacs.6b10816

206. Carson J. Bruns, Hanwei Liu, and Matthew B. Francis . Near-Quantitative Aqueous Synthesis of
Rotaxanes via Bioconjugation to Oligopeptides and Proteins. Journal of the American Chemical Society 2016,
138 (47) , 15307-15310. https://doi.org/10.1021/jacs.6b10231

207. Susnata Pramanik and Ivan Aprahamian . Hydrazone Switch-Based Negative Feedback Loop. Journal of
the American Chemical Society 2016, 138 (46) , 15142-15145. https://doi.org/10.1021/jacs.6b10542

208. Yuliia Vyborna, Mykhailo Vybornyi, and Robert Häner . Pathway Diversity in the Self-Assembly of DNA-
Derived Bioconjugates. Bioconjugate Chemistry 2016, 27 (11) , 2755-2761.
https://doi.org/10.1021/acs.bioconjchem.6b00517

209. Qishui Chen, Junling Sun, Peng Li, Idan Hod, Peyman Z. Moghadam, Zachary S. Kean, Randall Q. Snurr,
Joseph T. Hupp, Omar K. Farha, and J. Fraser Stoddart . A Redox-Active Bistable Molecular Switch Mounted
inside a Metal–Organic Framework. Journal of the American Chemical Society 2016, 138 (43) , 14242-14245.
https://doi.org/10.1021/jacs.6b09880

210. Adele Faulkner, Thomas van Leeuwen, Ben L. Feringa, and Sander J. Wezenberg . Allosteric Regulation
of the Rotational Speed in a Light-Driven Molecular Motor. Journal of the American Chemical Society 2016,
138 (41) , 13597-13603. https://doi.org/10.1021/jacs.6b06467

211. Mee-Kyung Chung, Peter S. White, Stephen J. Lee, Michel R. Gagné, and Marcey L. Waters . Investigation
of a Catenane with a Responsive Noncovalent Network: Mimicking Long-Range Responses in Proteins.
Journal of the American Chemical Society 2016, 138 (40) , 13344-13352.
https://doi.org/10.1021/jacs.6b07833
  
212. Xing Jiang, Hai-Bao Duan, Saeed I. Khan, and Miguel A. Garcia-Garibay . Diffusion-Controlled Rotation of
Triptycene in a Metal–Organic Framework (MOF) Sheds Light on the Viscosity of MOF-Con ned Solvent. ACS
Central Science 2016, 2 (9) , 608-613. https://doi.org/10.1021/acscentsci.6b00168

213. Munechika Nakamura, Kazuki Kishimoto, Yasuhiro Kobori, Tomoka Abe, Kenji Yoza, and Kenji Kobayashi
. Self-Assembled Molecular Gear: A 4:1 Complex of Rh(III)Cl Tetraarylporphyrin and Tetra(p-pyridyl)cavitand.
Journal of the American Chemical Society 2016, 138 (38) , 12564-12577.
https://doi.org/10.1021/jacs.6b07284

214. Sandra Wiedbrauk, Benjamin Maerz, Elena Samoylova, Anne Reiner, Florian Trommer, Peter Mayer,
Wolfgang Zinth, and Henry Dube . Twisted Hemithioindigo Photoswitches: Solvent Polarity Determines the
Type of Light-Induced Rotations. Journal of the American Chemical Society 2016, 138 (37) , 12219-12227.
https://doi.org/10.1021/jacs.6b05981

215. Dario Cambié, Cecilia Bottecchia, Natan J. W. Straathof, Volker Hessel, and Timothy Noël . Applications
of Continuous-Flow Photochemistry in Organic Synthesis, Material Science, and Water Treatment. Chemical
Reviews 2016, 116 (17) , 10276-10341. https://doi.org/10.1021/acs.chemrev.5b00707

216. Atsuro Takai, Takashi Kajitani, Takanori Fukushima, Keiki Kishikawa, Takeshi Yasuda, and Masayuki
Takeuchi . Supramolecular Assemblies of Ferrocene-Hinged Naphthalenediimides: Multiple Conformational
Changes in Film States. Journal of the American Chemical Society 2016, 138 (35) , 11245-11253.
https://doi.org/10.1021/jacs.6b05824

217. Alejandro López-Moreno, Belén Nieto-Ortega, Maria Moffa, Alberto de Juan, M. Mar Bernal, Juan P.
Fernández-Blázquez, Juan J. Vilatela, Dario Pisignano, and Emilio M. Pérez . Threading through Macrocycles
Enhances the Performance of Carbon Nanotubes as Polymer Fillers. ACS Nano 2016, 10 (8) , 8012-8018.
https://doi.org/10.1021/acsnano.6b04028

218. Toshiyuki Masuda, Junko Arase, Yusuke Inagaki, Masatoshi Kawahata, Kentaro Yamaguchi, Takashi
Ohhara, Akiko Nakao, Hiroyuki Momma, Eunsang Kwon, and Wataru Setaka . Molecular Gyrotops with a Five-
Membered Heteroaromatic Ring: Synthesis, Temperature-Dependent Orientation of Dipolar Rotors inside the
Crystal, and its Birefringence Change. Crystal Growth & Design 2016, 16 (8) , 4392-4401.
https://doi.org/10.1021/acs.cgd.6b00508

219. Nuno Basílio, Luís Cruz, Victor de Freitas, and Fernando Pina . A Multistate Molecular Switch Based on
the 6,8-Rearrangement in Bromo-apigeninidin Operated with pH and Host–Guest Inputs. The Journal of
Physical Chemistry B 2016, 120 (29) , 7053-7061. https://doi.org/10.1021/acs.jpcb.6b03694

220. Andrés Aguilar-Granda, Salvador Pérez-Estrada, Arian E. Roa, Joelis Rodríguez-Hernández, Simón
Hernández-Ortega, Mario Rodriguez, and Braulio Rodríguez-Molina . Synthesis of a Carbazole-[pi]-carbazole
Molecular Rotor with Fast Solid State Intramolecular Dynamics and Crystallization-Induced Emission. Crystal
Growth & Design 2016, 16 (6) , 3435-3442. https://doi.org/10.1021/acs.cgd.6b00395

221. Michael M. Lerch, Sander J. Wezenberg, Wiktor Szymanski, and Ben L. Feringa . Unraveling the
Photoswitching Mechanism in Donor–Acceptor Stenhouse Adducts. Journal of the American Chemical
Society 2016, 138 (20) , 6344-6347. https://doi.org/10.1021/jacs.6b01722
222. Víctor García-López, Jonathan Jeffet, Shunsuke Kuwahara, Angel A. Martí, Yuval Ebenstein, and James
M. Tour . Synthesis and Photostability of Unimolecular Submersible Nanomachines: Toward Single-Molecule
  
Tracking in Solution. Organic Letters 2016, 18 (10) , 2343-2346. https://doi.org/10.1021/acs.orglett.6b00506

223. Frank Biedermann and Hans-Jörg Schneider . Experimental Binding Energies in Supramolecular
Complexes. Chemical Reviews 2016, 116 (9) , 5216-5300. https://doi.org/10.1021/acs.chemrev.5b00583

224. Zheng Meng, Jun-Feng Xiang, and Chuan-Feng Chen . Directional Molecular Transportation Based on a
Catalytic Stopper-Leaving Rotaxane System. Journal of the American Chemical Society 2016, 138 (17) , 5652-
5658. https://doi.org/10.1021/jacs.6b01852

225. Yoshimasa Suzuki, Taiki Nakamura, Hiroki Iida, Naoki Ousaka, and Eiji Yashima . Allosteric Regulation of
Unidirectional Spring-like Motion of Double-Stranded Helicates. Journal of the American Chemical Society
2016, 138 (14) , 4852-4859. https://doi.org/10.1021/jacs.6b00787

226. Xing Jiang, Zachary J. O’Brien, Song Yang, Lan Huong Lai, Jeffrey Buena or, Colleen Tan, Saeed Khan, K.
N. Houk, and Miguel A. Garcia-Garibay . Crystal Fluidity Re ected by Fast Rotational Motion at the Core,
Branches, and Peripheral Aromatic Groups of a Dendrimeric Molecular Rotor. Journal of the American
Chemical Society 2016, 138 (13) , 4650-4656. https://doi.org/10.1021/jacs.6b01398

227. Dorota Czajkowska-Szczykowska, Andrés Aguilar-Granda, Jadwiga Maj, Agnieszka Z. Wilczewska,


Stanisław Witkowski, Rosa Santillan, Miguel A. Garcia-Garibay, Jacek W. Morzycki, and Braulio Rodríguez-
Molina . Solid State Characterization of Bridged Steroidal Molecular Rotors: Effect of the Rotator Fluorination
on Their Crystallization. Crystal Growth & Design 2016, 16 (3) , 1599-1605.
https://doi.org/10.1021/acs.cgd.5b01705

228. Yusuf Cakmak, Sundus Erbas-Cakmak, and David A. Leigh . Asymmetric Catalysis with a Mechanically
Point-Chiral Rotaxane. Journal of the American Chemical Society 2016, 138 (6) , 1749-1751.
https://doi.org/10.1021/jacs.6b00303

229. Jiang-Fei Xu, Zehuan Huang, Linghui Chen, Bo Qin, Qiao Song, Zhiqiang Wang, and Xi Zhang .
Supramolecular Polymerization Controlled by Reversible Conformational Modulation. ACS Macro Letters 2015,
4 (12) , 1410-1414. https://doi.org/10.1021/acsmacrolett.5b00831

230. Maximilian J. Urban, Chao Zhou, Xiaoyang Duan, and Na Liu . Optically Resolving the Dynamic Walking
of a Plasmonic Walker Couple. Nano Letters 2015, 15 (12) , 8392-8396.
https://doi.org/10.1021/acs.nanolett.5b04270

231. Federico Nicoli, Erica Paltrinieri, Marina Tran ć Bakić, Massimo Baroncini, Serena Silvi, Alberto Credi.
Binary logic operations with arti cial molecular machines. Coordination Chemistry Reviews 2021, 428 ,
213589. https://doi.org/10.1016/j.ccr.2020.213589

232. Jin Dai, Xubiao Peng, Antti J Niemi. Autonomous topological time crystals and knotty molecular motors.
Journal of Physics: Condensed Matter 2021, 33 (1) , 015702. https://doi.org/10.1088/1361-648X/abb682

233. Hiroshi Masai, Takuya Yokoyama, Hiromichi V. Miyagishi, Maning Liu, Yasuhiro Tachibana, Tetsuaki
Fujihara, Yasushi Tsuji, Jun Terao. Insulated conjugated bimetallopolymer with sigmoidal response by dual
self-controlling system as a biomimetic material. Nature Communications 2020, 11 (1)
https://doi.org/10.1038/s41467-019-14271-2
234. Ruizi Peng, Liujun Xu, Huijing Wang, Yifan Lyu, Dan Wang, Cheng Bi, Cheng Cui, Chunhai Fan, Qiaoling
Liu, Xiaobing Zhang, Weihong Tan. DNA-based arti cial molecular signaling system that mimics basic
  
elements of reception and response. Nature Communications 2020, 11 (1) https://doi.org/10.1038/s41467-
020-14739-6

235. Anthonius H. J. Engwerda, Stephen P. Fletcher. A molecular assembler that produces polymers. Nature
Communications 2020, 11 (1) https://doi.org/10.1038/s41467-020-17814-0

236. Bahareh Taghavi Shahraki, Saeid Maghsoudi, Yousef Fatahi, Navid Rabiee, Saeed Bahadorikhalili,
Rassoul Dinarvand, Mojtaba Bagherzadeh, Francis Verpoort. The owering of Mechanically Interlocked
Molecules: Novel approaches to the synthesis of rotaxanes and catenanes. Coordination Chemistry Reviews
2020, 423 , 213484. https://doi.org/10.1016/j.ccr.2020.213484

237. Pei‐Ming Cheng, Li‐Xuan Cai, Shao‐Chuan Li, Shao‐Jun Hu, Dan‐Ni Yan, Li‐Peng Zhou, Qing‐Fu Sun.
Guest‐Reaction Driven Cage to Conjoined Twin‐Cage Mitosis‐Like Host Transformation. Angewandte Chemie
International Edition 2020, https://doi.org/10.1002/anie.202011474

238. Katsuhiko Ariga. The evolution of molecular machines through interfacial nanoarchitectonics: from toys
to tools. Chemical Science 2020, 11 (39) , 10594-10604. https://doi.org/10.1039/D0SC03164J

239. Chanjuan Liu, Yuan Liu, Enqiang Zhu, Qiang Zhang, Xiaopeng Wei, Bin Wang. Cross-Inhibitor: a time-
sensitive molecular circuit based on DNA strand displacement. Nucleic Acids Research 2020, 350
https://doi.org/10.1093/nar/gkaa835

240. Ashwani Kumar, Subodh Kumar, Pil Seok Chae. A novel anthrapyridone diamine-based probe for
selective and distinctive Cu2+ and Hg2+ sensing in aqueous solution; utility as molecular logic gates. Dyes and
Pigments 2020, 181 , 108522. https://doi.org/10.1016/j.dyepig.2020.108522

241. Chak-Shing Kwan, Ken Cham-Fai Leung. Development and advancement of rotaxane dendrimers as
switchable macromolecular machines. Materials Chemistry Frontiers 2020, 4 (10) , 2825-2844.
https://doi.org/10.1039/D0QM00368A

242. Ya-Fang Gao, Zhi-Xu Zhang, Tie Zhang, Chang-Yuan Su, Wan-Ying Zhang, Da-Wei Fu. Regulated
molecular rotor in phase transition materials with switchable dielectric and SHG effect. Materials Chemistry
Frontiers 2020, 4 (10) , 3003-3012. https://doi.org/10.1039/D0QM00360C

243. Charlie T. McTernan, Guillaume De Bo, David A. Leigh. A Track-Based Molecular Synthesizer that Builds a
Single-Sequence Oligomer through Iterative Carbon-Carbon Bond Formation. Chem 2020,
https://doi.org/10.1016/j.chempr.2020.09.021

244. Yan-Hom Li, Shao-Chun Chen. The dynamics of a planar beating micro-swimmer constructed using
functional uid. Journal of Intelligent Material Systems and Structures 2020, 54 , 1045389X2095945.
https://doi.org/10.1177/1045389X20959454

245. Fabrizio Medici, Nawel Goual, Vincent Delattre, Arnaud Voituriez, Angela Marinetti. Photoswitchable
phosphines in catalysis. ChemCatChem 2020, 5 https://doi.org/10.1002/cctc.202000620

246. Roland Wilcken, Ludwig Huber, Kerstin Grill, Manuel Guentner, Monika Schildhauer, Stefan Thumser,
Eberhard Riedle, Henry Dube. Tuning the Ground and Excited State Dynamics of Hemithioindigo Molecular
Motors by Changing Substituents. Chemistry – A European Journal 2020,
https://doi.org/10.1002/chem.202003096
247. Toshiki Shimizu, Dominik Lungerich, Joshua Stuckner, Mitsuhiro Murayama, Koji Harano, Eiichi
Nakamura. Real-Time Video Imaging of Mechanical Motions of a Single Molecular Shuttle with Sub-
  
Millisecond Sub-Angstrom Precision #. Bulletin of the Chemical Society of Japan 2020, 93 (9) , 1079-1085.
https://doi.org/10.1246/bcsj.20200134

248. Jing‐Xiang Lin, Andrea Daolio, Patrick Scilabra, Giancarlo Terraneo, Hongfan Li, Giuseppe Resnati, Rong
Cao. The Relevance of Size Matching in Self‐assembly: Impact on Regio‐ and Chemoselective
Cocrystallizations. Chemistry – A European Journal 2020, 26 (51) , 11701-11704.
https://doi.org/10.1002/chem.202002264

249. Pei-Ming Cheng, Li-Xuan Cai, Shao-Chuan Li, Shao-Jun Hu, Dan-Ni Yan, Li-Peng Zhou, Qing-Fu Sun.
Guest‐Reaction Driven Cage to Conjoined Twin‐Cage Mitosis‐Like Host Transformation. Angewandte Chemie
2020, https://doi.org/10.1002/ange.202011474

250. Julián Valero, Marko Škugor. Mechanisms, Methods of Tracking and Applications of DNA Walkers: A
Review. ChemPhysChem 2020, 21 (17) , 1971-1988. https://doi.org/10.1002/cphc.202000235

251. Yu Zhang, Zhe Chang, Heng Zhao, Stefano Crespi, Ben L. Feringa, Depeng Zhao. A Chemically Driven
Rotary Molecular Motor Based on Reversible Lactone Formation with Perfect Unidirectionality. Chem 2020, 6
(9) , 2420-2429. https://doi.org/10.1016/j.chempr.2020.07.025

252. Lucie Čechová, Juraj Filo, Martin Dračínský, Chavdar Slavov, Dazhong Sun, Zlatko Janeba, Tomáš
Slanina, Josef Wachtveitl, Eliška Procházková, Marek Cigáň. Polysubstituted 5‐Phenylazopyrimidines:
Extremely Fast Non‐ionic Photochromic Oscillators. Angewandte Chemie International Edition 2020, 59 (36) ,
15590-15594. https://doi.org/10.1002/anie.202007065

253. Lucie Čechová, Juraj Filo, Martin Dračínský, Chavdar Slavov, Dazhong Sun, Zlatko Janeba, Tomáš
Slanina, Josef Wachtveitl, Eliška Procházková, Marek Cigáň. Polysubstituted 5‐Phenylazopyrimidines:
Extremely Fast Non‐ionic Photochromic Oscillators. Angewandte Chemie 2020, 132 (36) , 15720-15724.
https://doi.org/10.1002/ange.202007065

254. He-Ye Zhou, Qian-Shou Zong, Ying Han, Chuan-Feng Chen. Recent advances in higher order rotaxane
architectures. Chemical Communications 2020, 56 (69) , 9916-9936. https://doi.org/10.1039/D0CC03057K

255. Jessica Groppi, Lorenzo Casimiro, Martina Canton, Stefano Corra, Mina Jafari‐Nasab, Gloria Tabacchi,
Luigi Cavallo, Massimo Baroncini, Serena Silvi, Ettore Fois, Alberto Credi. Precision Molecular
Threading/Dethreading. Angewandte Chemie 2020, 132 (35) , 14935-14944.
https://doi.org/10.1002/ange.202003064

256. Jessica Groppi, Lorenzo Casimiro, Martina Canton, Stefano Corra, Mina Jafari‐Nasab, Gloria Tabacchi,
Luigi Cavallo, Massimo Baroncini, Serena Silvi, Ettore Fois, Alberto Credi. Precision Molecular
Threading/Dethreading. Angewandte Chemie International Edition 2020, 59 (35) , 14825-14834.
https://doi.org/10.1002/anie.202003064

257. Mark A. J. Koenis, C. S. Chibueze, M. A. Jinks, Valentin P. Nicu, Lucas Visscher, S. M. Goldup, Wybren J.
Buma. Vibrational circular dichroism spectroscopy for probing the expression of chirality in mechanically
planar chiral rotaxanes. Chemical Science 2020, 11 (32) , 8469-8475. https://doi.org/10.1039/D0SC02485F

258. Jeff M. Van Raden, Nanette N. Jarenwattananon, Lev N. Zakharov, Ramesh Jasti. Active Metal Template
Synthesis and Characterization of a Nanohoop [ c 2]Daisy Chain Rotaxane. Chemistry – A European Journal
2020, 26 (45) , 10205-10209. https://doi.org/10.1002/chem.202001389

  
259. Stefano Di Stefano, Daniele Del Giudice, Emanuele Spatola, Roberta Cacciapaglia, Alessandro Casnati,
Laura Baldini, Gianfranco Ercolani. Time Programmable Locking/Unlocking of the Calix[4]arene Scaffold by
Means of Chemical Fuels. Chemistry – A European Journal 2020, https://doi.org/10.1002/chem.202002574

260. Yoko Sakata, Takaya Ogura, Shigehisa Akine. E cient formation of [3]pseudorotaxane based on
cooperative complexation of dibenzo-24-crown-8 with diphenylviologen axle. Chemical Communications 2020,
56 (62) , 8735-8738. https://doi.org/10.1039/D0CC03131C

261. Chun‐Wei Chiu, Jye‐Shane Yang. Photoluminescent and Photoresponsive Iptycene‐Incorporated π‐


Conjugated Systems: Fundamentals and Applications. ChemPhotoChem 2020, 4 (8) , 538-563.
https://doi.org/10.1002/cptc.201900300

262. Yunyan Qiu, Yuanning Feng, Qing-Hui Guo, R. Dean Astumian, J. Fraser Stoddart. Pumps through the
Ages. Chem 2020, 6 (8) , 1952-1977. https://doi.org/10.1016/j.chempr.2020.07.009

263. John R.J. Maynard, Stephen M. Goldup. Strategies for the Synthesis of Enantiopure Mechanically Chiral
Molecules. Chem 2020, 6 (8) , 1914-1932. https://doi.org/10.1016/j.chempr.2020.07.012

264. Marta Olivares, Martin Albrecht. Switchable iridium hydride catalysts for controlling selectivity of alcohol
oxidation. Journal of Organometallic Chemistry 2020, 920 , 121290.
https://doi.org/10.1016/j.jorganchem.2020.121290

265. Qi Zhang, Da-Hui Qu, He Tian, Ben L. Feringa. Bottom-Up: Can Supramolecular Tools Deliver
Responsiveness from Molecular Motors to Macroscopic Materials?. Matter 2020, 3 (2) , 355-370.
https://doi.org/10.1016/j.matt.2020.05.014

266. Yingmei Han, Cameron Nickle, Ziyu Zhang, Hippolyte P. A. G. Astier, Thorin J. Du n, Dongchen Qi, Zhe
Wang, Enrique del Barco, Damien Thompson, Christian A. Nijhuis. Electric- eld-driven dual-functional
molecular switches in tunnel junctions. Nature Materials 2020, 19 (8) , 843-848.
https://doi.org/10.1038/s41563-020-0697-5

267. Xiaochuan Li, Qiongyan Cai, Junna Zhang, Hyorim Kim, Young-A Son. An “electron lock” toward the
photochromic activity of phenylacetylene appended bisthienylethene. Molecular Crystals and Liquid Crystals
2020, 706 (1) , 141-149. https://doi.org/10.1080/15421406.2020.1743450

268. Erik C. Vik, Ping Li, Josef M. Maier, Daniel O. Madukwe, Vitaly A. Rassolov, Perry J. Pellechia, Eric
Masson, Ken D. Shimizu. Large transition state stabilization from a weak hydrogen bond. Chemical Science
2020, 11 (28) , 7487-7494. https://doi.org/10.1039/D0SC02806A

269. Julián Valero, Michael Famulok. Regeneration of Burnt Bridges on a DNA Catenane Walker. Angewandte
Chemie 2020, 119 https://doi.org/10.1002/ange.202004447

270. Damien Thompson, Enrique del Barco, Christian A. Nijhuis. Design principles of dual-functional
molecular switches in solid-state tunnel junctions. Applied Physics Letters 2020, 117 (3) , 030502.
https://doi.org/10.1063/5.0016280

271. Julián Valero, Michael Famulok. Regeneration of Burnt Bridges on a DNA Catenane Walker. Angewandte
Chemie International Edition 2020, 119 https://doi.org/10.1002/anie.202004447
272. Alejandra M. Arango, Julien Wist, Javier Ellena, Richard D'Vries, Manuel N. Chaur. Multiple Reversible
Dynamics of Pyrimidine Based Acylhydrazones. European Journal of Organic Chemistry 2020, 2020 (26) , 
 
4009-4017. https://doi.org/10.1002/ejoc.202000553

273. Iago Neira, Olaya Domarco, Jose L. Barriada, Paola Franchi, Marco Lucarini, Marcos D. García, Carlos
Peinador. An electrochemically controlled supramolecular zip tie based on host–guest chemistry of CB[8].
Organic & Biomolecular Chemistry 2020, 18 (27) , 5228-5233. https://doi.org/10.1039/D0OB01085E

274. Benjamin Doistau, Lorien Benda, Jean-Louis Cantin, Olivier Cador, Fabrice Pointillart, Wolfgang
Wernsdorfer, Lise-Marie Chamoreau, Valérie Marvaud, Bernold Hasenknopf, Guillaume Vives. Dual switchable
molecular tweezers incorporating anisotropic Mn III –salphen complexes. Dalton Transactions 2020, 49 (26) ,
8872-8882. https://doi.org/10.1039/D0DT01465F

275. Bareld Wit, Ole Bunjes, Martin Wenderoth, Claus Ropers. Structure and Nonequilibrium Heat‐Transfer of
a Physisorbed Molecular Layer on Graphene. Advanced Materials Interfaces 2020, 7 (13) , 2000473.
https://doi.org/10.1002/admi.202000473

276. Mulinde R. Ruyonga, Oscar Mendoza, Michael Browne, Vyacheslav V. Samoshin. Exploration of trans ‐2‐
(azaarylsulfanyl)‐cyclohexanols as potential pH‐triggered conformational switches. Journal of Physical
Organic Chemistry 2020, 33 (7) https://doi.org/10.1002/poc.4068

277. Mikhail Yu. Belikov, Mikhail Yu. Ievlev, Sergey V. Fedoseev, Oleg V. Ershov. Synthesis and ne-tuning of
thermal stability of the negative nitrile-rich photochromes of hydroxytricyanopyrrole (HTCP) series. Research
on Chemical Intermediates 2020, 46 (7) , 3477-3490. https://doi.org/10.1007/s11164-020-04157-0

278. Benjamin H. Wilson, Stephen J. Loeb. Integrating the Mechanical Bond into Metal-Organic Frameworks.
Chem 2020, 6 (7) , 1604-1612. https://doi.org/10.1016/j.chempr.2020.06.016

279. Qianqian Gong, Yazhen Li, Fengyi Liu. Trajectory surface-hopping dynamics of light-driven imine
switches. Chemical Physics Letters 2020, 751 , 137506. https://doi.org/10.1016/j.cplett.2020.137506

280. Kateřina Bártová, Ivana Císařová, Antonín Lyčka, Martin Dračínský. Tautomerism of azo dyes in the solid
state studied by 15N, 14N, 13C and 1H NMR spectroscopy, X-ray diffraction and quantum-chemical
calculations. Dyes and Pigments 2020, 178 , 108342. https://doi.org/10.1016/j.dyepig.2020.108342

281. Fabio Castiglioni, Wojciech Danowski, Jacopo Perego, Franco King-Chi Leung, Piero Sozzani, Silvia
Bracco, Sander J. Wezenberg, Angiolina Comotti, Ben L. Feringa. Modulation of porosity in a solid material
enabled by bulk photoisomerization of an overcrowded alkene. Nature Chemistry 2020, 12 (7) , 595-602.
https://doi.org/10.1038/s41557-020-0493-5

282. Nils E. Strand, Rueih-Sheng Fu, Todd R. Gingrich. Current inversion in a periodically driven two-
dimensional Brownian ratchet. Physical Review E 2020, 102 (1)
https://doi.org/10.1103/PhysRevE.102.012141

283. Yuming Zhao. Special Issue “New Studies of Conjugated Compounds”. Molecules 2020, 25 (14) , 3220.
https://doi.org/10.3390/molecules25143220

284. Zhen-Hua Wang, Yu-Yuan Lu, Hui Jin, Chuan-Fu Luo, Li-Jia An. Diffusion of a Ring Threaded on a Linear
Chain. Chinese Journal of Polymer Science 2020, 18 https://doi.org/10.1007/s10118-020-2448-0
285. Aidan Kerckhoffs, Matthew J. Langton. Reversible photo-control over transmembrane anion transport
using visible-light responsive supramolecular carriers. Chemical Science 2020, 11 (24) , 6325-6331.
  
https://doi.org/10.1039/D0SC02745F

286. Yosuke Tani, Mao Komura, Takuji Ogawa. Mechanoresponsive turn-on phosphorescence by a
desymmetrization approach. Chemical Communications 2020, 56 (50) , 6810-6813.
https://doi.org/10.1039/D0CC01949F

287. Yuichiro Kobayashi, Kenji Kohara, Yusuke Kiuchi, Hiroki Onoda, Osami Shoji, Hiroyasu Yamaguchi.
Control of microenvironment around enzymes by hydrogels. Chemical Communications 2020, 56 (49) , 6723-
6726. https://doi.org/10.1039/D0CC01332C

288. Yunmeng Jiang, Yanhua Cheng, Shunjie Liu, Haoke Zhang, Xiaoyan Zheng, Ming Chen, Michidmaa
Khorloo, Hengxue Xiang, Ben Zhong Tang, Meifang Zhu. Solid-state intramolecular motions in continuous
bers driven by ambient humidity for uorescent sensors. National Science Review 2020, 288
https://doi.org/10.1093/nsr/nwaa135

289. Chao Chen, Xiang Ni, Han‐Wen Tian, Qian Liu, Dong‐Sheng Guo, Dan Ding. Calixarene‐Based
Supramolecular AIE Dots with Highly Inhibited Nonradiative Decay and Intersystem Crossing for Ultrasensitive
Fluorescence Image‐Guided Cancer Surgery. Angewandte Chemie 2020, 132 (25) , 10094-10098.
https://doi.org/10.1002/ange.201916430

290. Chao Chen, Xiang Ni, Han‐Wen Tian, Qian Liu, Dong‐Sheng Guo, Dan Ding. Calixarene‐Based
Supramolecular AIE Dots with Highly Inhibited Nonradiative Decay and Intersystem Crossing for Ultrasensitive
Fluorescence Image‐Guided Cancer Surgery. Angewandte Chemie International Edition 2020, 59 (25) , 10008-
10012. https://doi.org/10.1002/anie.201916430

291. Yunyan Qiu, Bo Song, Cristian Pezzato, Dengke Shen, Wenqi Liu, Long Zhang, Yuanning Feng, Qing-Hui
Guo, Kang Cai, Weixingyue Li, Hongliang Chen, Minh T. Nguyen, Yi Shi, Chuyang Cheng, R. Dean Astumian,
Xiaopeng Li, J. Fraser Stoddart. A precise polyrotaxane synthesizer. Science 2020, 368 (6496) , 1247-1253.
https://doi.org/10.1126/science.abb3962

292. Jiannan Tang, Louis William Rogowski, Xiao Zhang, Min Jun Kim. Flagellar nanorobot with kinetic
behavior investigation and 3D motion. Nanoscale 2020, 12 (22) , 12154-12164.
https://doi.org/10.1039/D0NR02496A

293. Bibhuti Bhusan Rath, Goutam Kumar Kole, Samuel Alexander Morris, Jagadese J. Vittal. Rotation of a
helical coordination polymer by mechanical grinding. Chemical Communications 2020, 56 (46) , 6289-6292.
https://doi.org/10.1039/D0CC02158J

294. Abraham Colin-Molina, Diego Velázquez-Chávez, Marcus J. Jellen, Lizbeth A. Rodríguez-Cortés, Miguel
Eduardo Cifuentes-Quintal, Gabriel Merino, Braulio Rodríguez-Molina. Dynamic characterization of crystalline
uorophores with conformationally exible tetrahydrocarbazole frameworks. CrystEngComm 2020, 22 (22) ,
3789-3796. https://doi.org/10.1039/D0CE00423E

295. Sonia La Cognata, Ana Miljkovic, Riccardo Mobili, Greta Bergamaschi, Valeria Amendola. Organic Cages
as Building Blocks for Mechanically Interlocked Molecules: Towards Molecular Machines. ChemPlusChem
2020, 85 (6) , 1145-1155. https://doi.org/10.1002/cplu.202000274

296. Stefano F. Pizzolato, Peter Štacko, Jos C. M. Kistemaker, Thomas van Leeuwen, Ben L. Feringa.
Phosphoramidite-based photoresponsive ligands displaying multifold transfer of chirality in dynamic
enantioselective metal catalysis. Nature Catalysis 2020, 3 (6) , 488-496. https://doi.org/10.1038/s41929-020-
0452-y
  
297. Kelong Zhu, Stephen J. Loeb. A hydrogen-bonded polymer constructed from mechanically interlocked,
suit[1]ane monomers. Canadian Journal of Chemistry 2020, 98 (6) , 285-291. https://doi.org/10.1139/cjc-
2020-0002

298. Chiara Biagini, Giorgio Capocasa, Daniele Del Giudice, Valerio Cataldi, Luigi Mandolini, Stefano Di
Stefano. Controlling the liberation rate of the in situ release of a chemical fuel for the operationally
autonomous motions of molecular machines. Organic & Biomolecular Chemistry 2020, 18 (20) , 3867-3873.
https://doi.org/10.1039/D0OB00669F

299. Monika Stolar. Organic electrochromic molecules: synthesis, properties, applications and impact. Pure
and Applied Chemistry 2020, 92 (5) , 717-731. https://doi.org/10.1515/pac-2018-1208

300. Chiara Biagini, Stefano Di Stefano. Abiotic Chemical Fuels for the Operation of Molecular Machines.
Angewandte Chemie 2020, 132 (22) , 8420-8430. https://doi.org/10.1002/ange.201912659

301. Chiara Biagini, Stefano Di Stefano. Abiotic Chemical Fuels for the Operation of Molecular Machines.
Angewandte Chemie International Edition 2020, 59 (22) , 8344-8354.
https://doi.org/10.1002/anie.201912659

302. Ana María Sánchez-León, Pedro Cintas, Mark E. Light, Juan Carlos Palacios. Thermal and
Photochemical Switching of Chiral Sugar Azoalkenes: A Mechanistic Interrogation. European Journal of
Organic Chemistry 2020, 2020 (19) , 2827-2841. https://doi.org/10.1002/ejoc.202000029

303. Hikaru Aramoto, Motofumi Osaki, Subaru Konishi, Chiharu Ueda, Yuichiro Kobayashi, Yoshinori
Takashima, Akira Harada, Hiroyasu Yamaguchi. Redox-responsive supramolecular polymeric networks having
double-threaded inclusion complexes. Chemical Science 2020, 11 (17) , 4322-4331.
https://doi.org/10.1039/C9SC05589D

304. Chang‐Bo Huang, Artur Ciesielski, Paolo Samorì. Molecular Springs: Integration of Complex Dynamic
Architectures into Functional Devices. Angewandte Chemie 2020, 132 (19) , 7387-7398.
https://doi.org/10.1002/ange.201914931

305. Chang‐Bo Huang, Artur Ciesielski, Paolo Samorì. Molecular Springs: Integration of Complex Dynamic
Architectures into Functional Devices. Angewandte Chemie International Edition 2020, 59 (19) , 7319-7330.
https://doi.org/10.1002/anie.201914931

306. Andreas Walther. Viewpoint: From Responsive to Adaptive and Interactive Materials and Materials
Systems: A Roadmap. Advanced Materials 2020, 32 (20) , 1905111.
https://doi.org/10.1002/adma.201905111

307. Emilie Moulin, Lara Faour, Christian C. Carmona‐Vargas, Nicolas Giuseppone. From Molecular Machines
to Stimuli‐Responsive Materials. Advanced Materials 2020, 32 (20) , 1906036.
https://doi.org/10.1002/adma.201906036

308. Stefano Corra, Massimiliano Curcio, Massimo Baroncini, Serena Silvi, Alberto Credi. Photoactivated
Arti cial Molecular Machines that Can Perform Tasks. Advanced Materials 2020, 32 (20) , 1906064.
https://doi.org/10.1002/adma.201906064
309. Durga Prasad Karothu, Jad Mahmoud Halabi, Liang Li, Abraham Colin‐Molina, Braulio Rodríguez‐Molina,
Panče Naumov. Global Performance Indices for Dynamic Crystals as Organic Thermal Actuators. Advanced
Materials 2020, 32 (20) , 1906216. https://doi.org/10.1002/adma.201906216
  

310. Shuhui Yin, Ming Tien, Haw Yang. Prior-Apprised Unsupervised Learning of Subpixel Curvilinear Features
in Low Signal/Noise Images. Biophysical Journal 2020, 118 (10) , 2458-2469.
https://doi.org/10.1016/j.bpj.2020.04.009

311. V. Deneva, N.G. Vassilev, S. Hristova, D. Yordanov, Y. Hayashi, S. Kawauchi, F. Fennel, T. Völzer, S.
Lochbrunner, L. Antonov. Chercher de l'eau: The switching mechanism of the rotary switch ethyl-2-(2-(quinolin-
8-yl)hydrazono)-2-(pyridin-2-yl)acetate. Computational Materials Science 2020, 177 , 109570.
https://doi.org/10.1016/j.commatsci.2020.109570

312. Daniela F. Báez, Gabriel Ramos, Alejandro Corvalán, María Luisa Cordero, Soledad Bollo, Marcelo J.
Kogan. Effects of preparation on catalytic, magnetic and hybrid micromotors on their functional features and
application in gastric cancer biomarker detection. Sensors and Actuators B: Chemical 2020, 310 , 127843.
https://doi.org/10.1016/j.snb.2020.127843

313. Masahide Tominaga, Tadashi Hyodo, Yumi Maekawa, Masatoshi Kawahata, Kentaro Yamaguchi. One‐
Step Synthesis of Cyclophanes as Crystalline Sponge and Their [2]Catenanes through S N Ar Reactions.
Chemistry – A European Journal 2020, 26 (23) , 5157-5161. https://doi.org/10.1002/chem.201905854

314. Antonio Rodríguez-Fortea, Enric Canadell, Pawel Wzietek, Cyprien Lemouchi, Magali Allain, Leokadiya
Zorina, Patrick Batail. Nanoscale rotational dynamics of four independent rotators con ned in crowded
crystalline layers. Nanoscale 2020, 12 (15) , 8294-8302. https://doi.org/10.1039/D0NR00858C

315. Saikat Santra, Pradyut Ghosh. Fluorophoric [2]rotaxanes: post-synthetic functionalization,


conformational uxionality and metal ion chelation. New Journal of Chemistry 2020, 44 (15) , 5947-5964.
https://doi.org/10.1039/D0NJ00353K

316. Krzysztof M. Bąk, Kyriakos Porfyrakis, Jason J. Davis, Paul D. Beer. Exploiting the mechanical bond for
molecular recognition and sensing of charged species. Materials Chemistry Frontiers 2020, 4 (4) , 1052-1073.
https://doi.org/10.1039/C9QM00698B

317. Iwona Nierengarten, Jean‐François Nierengarten. Diversity Oriented Preparation of Pillar[5]arene‐


Containing [2]Rotaxanes by a Stopper Exchange Strategy. ChemistryOpen 2020, 9 (4) , 393-400.
https://doi.org/10.1002/open.202000035

318. Jakia Jannat Keya, Arif Md. Rashedul Kabir, Akira Kakugo. Synchronous operation of biomolecular
engines. Biophysical Reviews 2020, 12 (2) , 401-409. https://doi.org/10.1007/s12551-020-00651-2

319. Ana Miljkovic, Sonia La Cognata, Greta Bergamaschi, Mauro Freccero, Antonio Poggi, Valeria Amendola.
Towards Building Blocks for Supramolecular Architectures Based on Azacryptates. Molecules 2020, 25 (7) ,
1733. https://doi.org/10.3390/molecules25071733

320. Tsukuru Minamiki, Yuki Ichikawa, Ryoji Kurita. The Power of Assemblies at Interfaces: Nanosensor
Platforms Based on Synthetic Receptor Membranes. Sensors 2020, 20 (8) , 2228.
https://doi.org/10.3390/s20082228

321. Edgar Uhl, Peter Mayer, Henry Dube. Active and Unidirectional Acceleration of Biaryl Rotation by a
Molecular Motor. Angewandte Chemie 2020, 132 (14) , 5779-5786. https://doi.org/10.1002/ange.201913798
322. Edgar Uhl, Peter Mayer, Henry Dube. Active and Unidirectional Acceleration of Biaryl Rotation by a
  
Molecular Motor. Angewandte Chemie International Edition 2020, 59 (14) , 5730-5737.
https://doi.org/10.1002/anie.201913798

323. Zahra Ahmadian Dehaghani, Iurii Chubak, Christos N. Likos, Mohammad Reza Ejtehadi. Effects of
topological constraints on linked ring polymers in solvents of varying quality. Soft Matter 2020, 16 (12) , 3029-
3038. https://doi.org/10.1039/C9SM02374G

324. Maryam F. Abdollahi, Yuming Zhao. Recent advances in dithiafulvenyl-functionalized organic conjugated
materials. New Journal of Chemistry 2020, 44 (12) , 4681-4693. https://doi.org/10.1039/C9NJ06430C

325. Adam Mames, Dariusz Gołowicz, Mariusz Pietrzak, Krzysztof Kazimierczuk, Sławomir Szymański,
Tomasz Ratajczyk. Blue‐Shift Hydrogen Bonds in Silyltriptycene Derivatives: Antibonding σ* Orbitals of the
Si−C Bond as Effective Acceptors of Electron Density. ChemPhysChem 2020, 21 (6) , 540-545.
https://doi.org/10.1002/cphc.201901141

326. Daiki Inamori, Hiroshi Masai, Takashi Tamaki, Jun Terao. Macroscopic Change in Luminescent Color by
Thermally Driven Sliding Motion in [3]Rotaxanes. Chemistry – A European Journal 2020, 26 (15) , 3385-3389.
https://doi.org/10.1002/chem.201905342

327. Hang Chen, Hebo Ye, Yu Hai, Ling Zhang, Lei You. n → π* interactions as a versatile tool for controlling
dynamic imine chemistry in both organic and aqueous media. Chemical Science 2020, 11 (10) , 2707-2715.
https://doi.org/10.1039/C9SC05698J

328. Andrew W. Heard, Stephen M. Goldup. Synthesis of a Mechanically Planar Chiral Rotaxane Ligand for
Enantioselective Catalysis. Chem 2020, https://doi.org/10.1016/j.chempr.2020.02.006

329. Ashwani Kumar, Pil Seok Chae, Subodh Kumar. A dual-responsive anthrapyridone-triazole-based probe
for selective detection of Ni2+ and Cu2+: A mimetic system for molecular logic gates based on color change.
Dyes and Pigments 2020, 174 , 108092. https://doi.org/10.1016/j.dyepig.2019.108092

330. Lydia Chabane, Raphaël Chétrite, Gatien Verley. Periodically driven jump processes conditioned on large
deviations. Journal of Statistical Mechanics: Theory and Experiment 2020, 2020 (3) , 033208.
https://doi.org/10.1088/1742-5468/ab74c4

331. Zhihui Zhang, Graham J. Tizzard, J. A. Gareth Williams, Stephen M. Goldup. Rotaxane Pt II -complexes:
mechanical bonding for chemically robust luminophores and stimuli responsive behaviour. Chemical Science
2020, 11 (7) , 1839-1847. https://doi.org/10.1039/C9SC05507J

332. Maria A. Cardona, Leonard J. Prins. ATP-fuelled self-assembly to regulate chemical reactivity in the time
domain. Chemical Science 2020, 11 (6) , 1518-1522. https://doi.org/10.1039/C9SC05188K

333. Takanori Nakamura, Yuka Mori, Masaya Naito, Yukari Okuma, Shinobu Miyagawa, Hikaru Takaya,
Tsuneomi Kawasaki, Yuji Tokunaga. Rotaxanes comprising cyclic phenylenedioxydiacetamides and secondary
mono- and bis-dialkylammonium ions: effect of macrocyclic ring size on pseudorotaxane formation. Organic
Chemistry Frontiers 2020, 7 (3) , 513-524. https://doi.org/10.1039/C9QO01359H

334. Junya Adachi, Taizo Mori, Ryo Inoue, Masaya Naito, Ngoc Ha‐Thu Le, Soichiro Kawamorita, Jonathan P.
Hill, Takeshi Naota, Katsuhiko Ariga. Emission Control by Molecular Manipulation of Double‐Paddled Binuclear
Pt II Complexes at the Air‐Water Interface. Chemistry – An Asian Journal 2020, 15 (3) , 406-414.
https://doi.org/10.1002/asia.201901691
  
335. Baihao Shao, Ivan Aprahamian. Planarization‐Induced Activation Wavelength Red‐Shift and Thermal
Half‐Life Acceleration in Hydrazone Photoswitches. ChemistryOpen 2020, 9 (2) , 191-194.
https://doi.org/10.1002/open.201900340

336. Guang Liu, Jiajin Hong, Kefeng Ma, Ying Wan, Xutong Zhang, Yaqi Huang, Kai Kang, Meng Yang, Jialiang
Chen, Shengyuan Deng. Porphyrin Trio−Pendant fullerene guest as an In situ universal probe of high ECL
e ciency for sensitive miRNA detection. Biosensors and Bioelectronics 2020, 150 , 111963.
https://doi.org/10.1016/j.bios.2019.111963

337. Pritam Mondal, Sankar Prasad Rath. Cyclic metalloporphyrin dimers: Conformational exibility,
applications and future prospects. Coordination Chemistry Reviews 2020, 405 , 213117.
https://doi.org/10.1016/j.ccr.2019.213117

338. Edison Matamala-Cea, Fabián Valenzuela-Godoy, Déborah González, Rodrigo Arancibia, Vincent Dorcet,
Jean-René Hamon, Néstor Novoa. E cient preparation of 5,10,15,20-tetrakis(4-bromophenyl)porphyrin.
Microwave assisted v/s conventional synthetic method, X-ray and hirshfeld surface structural analysis. Journal
of Molecular Structure 2020, 1201 , 127139. https://doi.org/10.1016/j.molstruc.2019.127139

339. Nikita Mittal, Indrajit Paul, Susnata Pramanik, Michael Schmittel. Remote control of the reversible
assembly/disassembly of supramolecular aggregates. Supramolecular Chemistry 2020, 32 (2) , 133-138.
https://doi.org/10.1080/10610278.2020.1711907

340. Yong-Fu Li, Zheng Li, Qi Lin, Ying-Wei Yang. Functional supramolecular gels based on pillar[ n ]arene
macrocycles. Nanoscale 2020, 12 (4) , 2180-2200. https://doi.org/10.1039/C9NR09532B

341. Dirk Schlüter, Florian Kleemiss, Malte Fugel, Enno Lork, Kunihisa Sugimoto, Simon Grabowsky, Jeffrey R.
Harmer, Matthias Vogt. Non‐Oxido‐Vanadium(IV) Metalloradical Complexes with Bidentate 1,2‐Dithienylethene
Ligands: Observation of Reversible Cyclization of the Ligand Scaffold in Solution. Chemistry – A European
Journal 2020, 26 (6) , 1335-1343. https://doi.org/10.1002/chem.201904103

342. Aaron D. W. Kennedy, Isolde Sandler, Joakim Andréasson, Junming Ho, Jonathon E. Beves. Visible‐Light
Photoswitching by Azobenzazoles. Chemistry – A European Journal 2020, 26 (5) , 1103-1110.
https://doi.org/10.1002/chem.201904309

343. Enrico Berardo, Rebecca L. Greenaway, Marcin Miklitz, Andrew I. Cooper, Kim E. Jelfs. Computational
screening for nested organic cage complexes. Molecular Systems Design & Engineering 2020, 5 (1) , 186-196.
https://doi.org/10.1039/C9ME00085B

344. Youzhi Xu, Max Delius. Supramolekulare Chemie von gespannten Kohlenstoffnanoreifen. Angewandte
Chemie 2020, 132 (2) , 567-582. https://doi.org/10.1002/ange.201906069

345. Youzhi Xu, Max Delius. The Supramolecular Chemistry of Strained Carbon Nanohoops. Angewandte
Chemie International Edition 2020, 59 (2) , 559-573. https://doi.org/10.1002/anie.201906069

346. Christian Joachim. From the Anthycytera Astronomical Clock to Single Molecule Scale Machinery.
2020,,, 1-7. https://doi.org/10.1007/978-3-030-56777-4_1
347. Michael Schmittel, Abir Goswami, Indrajit Paul, Pronay Kumar Biswas. Motion and Nanomechanical
Effects in Supramolecular Catalysts. 2020,,, 195-218. https://doi.org/10.1007/978-3-030-56777-4_13
  
348. Yohan Gisbert, Agnès M. Sirven, Gwénaël Rapenne, Claire Kammerer. Design and Synthesis of a Nano-
winch. 2020,,, 81-98. https://doi.org/10.1007/978-3-030-56777-4_6

349. Cai-Xin Zhao, Qi Zhang, Gábor London, Da-Hui Qu. Functional Rotaxanes. 2020,,, 277-310.
https://doi.org/10.1007/978-981-15-2686-2_12

350. Bin Li, Yiliang Wang, Chunju Li. Biphen[n]arenes: Synthesis and Host–Guest Properties. 2020,,, 311-339.
https://doi.org/10.1007/978-981-15-2686-2_13

351. Wensheng Cai, Haohao Fu. Molecular Simulations of Supramolecular Architectures. 2020,,, 1107-1133.
https://doi.org/10.1007/978-981-15-2686-2_45

352. Jonathan Ainsley Iggo, Konstantin V. Luzyanin. . 2020,,https://doi.org/10.1016/B978-0-08-102688-


5.00002-7

353. Marcello La Rosa, Massimo Baroncini, Serena Silvi, Alberto Credi. Manufacturing at nanoscale. 2020,,,
41-63. https://doi.org/10.1016/B978-0-12-816865-3.00002-0

354. Juan Bueno. Autopoietic symbiogenesis in secondary metabolism. 2020,,, 75-90.


https://doi.org/10.1016/B978-0-12-817613-9.00005-5

355. Filip Novotný, Hong Wang, Martin Pumera. Nanorobots: Machines Squeezed between Molecular Motors
and Micromotors. Chem 2020, https://doi.org/10.1016/j.chempr.2019.12.028

356. Hamid Ramezani, Hendrik Dietz. Building machines with DNA molecules. Nature Reviews Genetics
2020, 21 (1) , 5-26. https://doi.org/10.1038/s41576-019-0175-6

357. He-Ye Zhou, Ying Han, Chuan-Feng Chen. pH-Controlled motions in mechanically interlocked molecules.
Materials Chemistry Frontiers 2020, 4 (1) , 12-28. https://doi.org/10.1039/C9QM00546C

358. David T. Hogan, Benjamin S. Gelfand, Denis M. Spasyuk, Todd C. Sutherland. Subtle substitution controls
the rainbow chromatic behaviour of multi-stimuli responsive core-expanded pyrenes. Materials Chemistry
Frontiers 2020, 4 (1) , 268-276. https://doi.org/10.1039/C9QM00710E

359. Surbhi Grewal, Saonli Roy, Himanshu Kumar, Mayank Saraswat, Naimat K. Bari, Sharmistha Sinha,
Sugumar Venkataramani. Temporal control in tritylation reactions through light-driven variation in chloride ion
binding catalysis – a proof of concept. Catalysis Science & Technology 2020, 105
https://doi.org/10.1039/D0CY01090A

360. Nadia Hoyas Pérez, James E. M. Lewis. Synthetic strategies towards mechanically interlocked
oligomers and polymers. Organic & Biomolecular Chemistry 2020, 83 https://doi.org/10.1039/D0OB01583K

361. Enqiang Zhu, Congzhou Chen, Yongsheng Rao, Weicheng Xiong. Biochemical Logic Circuits Based on
DNA Combinatorial Displacement. IEEE Access 2020, 8 , 34096-34103.
https://doi.org/10.1109/ACCESS.2020.2974024
362. Raffaello Papadakis. Mono- and Di-Quaternized 4,4′-Bipyridine Derivatives as Key Building Blocks for
Medium- and Environment-Responsive Compounds and Materials. Molecules 2020, 25 (1) , 1.
https://doi.org/10.3390/molecules25010001
  

363. Vishnu Verman Rajasekaran, Indrajit Paul, Michael Schmittel. Reversible switching from a three- to a
nine-fold degenerate dynamic slider-on-deck through catenation. Chemical Communications 2020, 333
https://doi.org/10.1039/D0CC05578F

364. Thomas S. C. MacDonald, William S. Price, R. Dean Astumian, Jonathon E. Beves. Enhanced Diffusion of
Molecular Catalysts is Due to Convection. Angewandte Chemie 2019, 131 (52) , 19040-19043.
https://doi.org/10.1002/ange.201910968

365. Thomas S. C. MacDonald, William S. Price, R. Dean Astumian, Jonathon E. Beves. Enhanced Diffusion of
Molecular Catalysts is Due to Convection. Angewandte Chemie International Edition 2019, 58 (52) , 18864-
18867. https://doi.org/10.1002/anie.201910968

366. Thomas Bridge, Saher A. Shaikh, Paul Thomas, Joaquin Botta, Peter J. McCormick, Amit Sachdeva. Site‐
Speci c Encoding of Photoactivity in Antibodies Enables Light‐Mediated Antibody–Antigen Binding on Live
Cells. Angewandte Chemie 2019, 131 (50) , 18154-18161. https://doi.org/10.1002/ange.201908655

367. Thomas Bridge, Saher A. Shaikh, Paul Thomas, Joaquin Botta, Peter J. McCormick, Amit Sachdeva. Site‐
Speci c Encoding of Photoactivity in Antibodies Enables Light‐Mediated Antibody–Antigen Binding on Live
Cells. Angewandte Chemie International Edition 2019, 58 (50) , 17986-17993.
https://doi.org/10.1002/anie.201908655

368. Suchismita Saha, Pronay Kumar Biswas, Indrajit Paul, Michael Schmittel. Selective and reversible
interconversion of nanosliders commanded by remote control via metal-ion signaling. Chemical
Communications 2019, 55 (98) , 14733-14736. https://doi.org/10.1039/C9CC07415E

369. Zheng Zhao, Chao Chen, Wenting Wu, Fenfen Wang, Lili Du, Xiaoyan Zhang, Yu Xiong, Xuewen He,
Yuanjing Cai, Ryan T. K. Kwok, Jacky W. Y. Lam, Xike Gao, Pingchuan Sun, David Lee Phillips, Dan Ding, Ben
Zhong Tang. Highly e cient photothermal nanoagent achieved by harvesting energy via excited-state
intramolecular motion within nanoparticles. Nature Communications 2019, 10 (1)
https://doi.org/10.1038/s41467-019-08722-z

370. Patrick M. J. Szell, Scott Zablotny, David L. Bryce. Halogen bonding as a supramolecular dynamics
catalyst. Nature Communications 2019, 10 (1) https://doi.org/10.1038/s41467-019-08878-8

371. Zheng Zhao, Xiaoyan Zheng, Lili Du, Yu Xiong, Wei He, Xiuxiu Gao, Chunli Li, Yingjie Liu, Bin Xu, Jing
Zhang, Fengyan Song, Ying Yu, Xueqian Zhao, Yuanjing Cai, Xuewen He, Ryan T. K. Kwok, Jacky W. Y. Lam,
Xuhui Huang, David Lee Phillips, Hua Wang, Ben Zhong Tang. Non-aromatic annulene-based aggregation-
induced emission system via aromaticity reversal process. Nature Communications 2019, 10 (1)
https://doi.org/10.1038/s41467-019-10818-5

372. Ru-Qiang Lu, Xiao-Yun Yan, Lei Zhu, Lin-Lin Yang, Hang Qu, Xin-Chang Wang, Ming Luo, Yu Wang, Rui
Chen, Xiao-Ye Wang, Yu Lan, Jian Pei, Wengui Weng, Haiping Xia, Xiao-Yu Cao. Unveiling how intramolecular
stacking modes of covalently linked dimers dictate photoswitching properties. Nature Communications 2019,
10 (1) https://doi.org/10.1038/s41467-019-13428-3

373. Lorien Benda, Benjamin Doistau, Caroline Rossi-Gendron, Lise-Marie Chamoreau, Bernold Hasenknopf,
Guillaume Vives. Substrate-dependent allosteric regulation by switchable catalytic molecular tweezers.
Communications Chemistry 2019, 2 (1) https://doi.org/10.1038/s42004-019-0246-9

  
374. Elena Prigorchenko, Lukas Ustrnul, Victor Borovkov, Riina Aav. Heterocomponent ternary supramolecular
complexes of porphyrins: A review. Journal of Porphyrins and Phthalocyanines 2019, 23 (11n12) , 1308-1325.
https://doi.org/10.1142/S108842461930026X

375. Aaron Gerwien, Peter Mayer, Henry Dube. Green light powered molecular state motor enabling eight-
shaped unidirectional rotation. Nature Communications 2019, 10 (1) https://doi.org/10.1038/s41467-019-
12463-4

376. Lukas Ustrnul, Sandra Kaabel, Tatsiana Burankova, Jevgenija Martõnova, Jasper Adamson, Nele Konrad,
Peeter Burk, Victor Borovkov, Riina Aav. Supramolecular chirogenesis in zinc porphyrins by enantiopure
hemicucurbit[ n ]urils ( n = 6, 8). Chemical Communications 2019, 55 (96) , 14434-14437.
https://doi.org/10.1039/C9CC07150D

377. Chiara Biagini, Giorgio Capocasa, Valerio Cataldi, Daniele Del Giudice, Luigi Mandolini, Stefano Di
Stefano. The Hydrolysis of the Anhydride of 2‐Cyano‐2‐phenylpropanoic Acid Triggers the Repeated Back and
Forth Motions of an Acid–Base Operated Molecular Switch. Chemistry – A European Journal 2019, 25 (66) ,
15205-15211. https://doi.org/10.1002/chem.201904048

378. . Fundamental Mechanisms of Intelligent Responsiveness. 2019,,, 27-68.


https://doi.org/10.1002/9783527816354.ch2

379. Yan-Cen Liu, Werner M. Nau, Andreas Hennig. A supramolecular ve-component relay switch that
exposes the mechanistic competition of dissociative versus associative binding to cucurbiturils by ratiometric
uorescence monitoring. Chemical Communications 2019, 55 (94) , 14123-14126.
https://doi.org/10.1039/C9CC07165B

380. Masaya Naito, Takaaki Fujino, Shinya Tajima, Shinobu Miyagawa, Kazuyuki Yoshida, Hajime Inoue,
Hiroaki Takagawa, Tsuneomi Kawasaki, Yuji Tokunaga. Ring size affects the kinetic and thermodynamic
formation of [2]rotaxanes featuring an unsymmetric bis-crown ether component. Materials Chemistry
Frontiers 2019, 3 (12) , 2702-2706. https://doi.org/10.1039/C9QM00441F

381. Márcia Pessêgo, Johan Mendoza, José Paulo da Silva, Nuno Basílio, Luis Garcia-Rio. Unveiling the
formation 1 : 2 supramolecular complexes between cucurbit[7]uril and a cationic calix[4]arene derivative.
Chemical Communications 2019, 55 (92) , 13828-13831. https://doi.org/10.1039/C9CC07280B

382. Jiong Zhou, Guocan Yu, Yang Li, Jie Shen, Mengbin Wang, Zhengtao Li, Peifa Wei, Jianbin Tang, Feihe
Huang. [2]Pseudorotaxane‐Based Supramolecular Optical Indicator for the Visual Detection of Cellular Cyanide
Excretion. Chemistry – A European Journal 2019, 25 (63) , 14447-14453.
https://doi.org/10.1002/chem.201903577

383. Rubi A. Luna‐Ixmatlahua, Anayeli Carrasco‐Ruiz, Ruy Cervantes, Alberto Vela, Jorge Tiburcio. An Anionic
Ring Locked into an Anionic Axle: A Metastable Rotaxane with Chemically Activated Electrostatic Stoppers.
Chemistry – A European Journal 2019, 25 (62) , 14042-14047. https://doi.org/10.1002/chem.201902735

384. Marius Gaedke, Felix Witte, Jana Anhäuser, Henrik Hupatz, Hendrik V. Schröder, Arto Valkonen, Kari
Rissanen, Arne Lützen, Beate Paulus, Christoph A. Schalley. Chiroptical inversion of a planar chiral redox-
switchable rotaxane. Chemical Science 2019, 10 (43) , 10003-10009. https://doi.org/10.1039/C9SC03694F
385. Guang‐Tao Xu, Liang‐Liang Wu, Xiao‐Yong Chang, Tim Wai Hung Ang, Wai‐Yeung Wong, Jie‐Sheng
Huang, Chi‐Ming Che. Solvent‐Induced Cluster‐to‐Cluster Transformation of Homoleptic Gold(I) Thiolates
  
between Catenane and Ring‐in‐Ring Structures. Angewandte Chemie 2019, 131 (45) , 16443-16452.
https://doi.org/10.1002/ange.201909980

386. Guang‐Tao Xu, Liang‐Liang Wu, Xiao‐Yong Chang, Tim Wai Hung Ang, Wai‐Yeung Wong, Jie‐Sheng
Huang, Chi‐Ming Che. Solvent‐Induced Cluster‐to‐Cluster Transformation of Homoleptic Gold(I) Thiolates
between Catenane and Ring‐in‐Ring Structures. Angewandte Chemie International Edition 2019, 58 (45) ,
16297-16306. https://doi.org/10.1002/anie.201909980

387. Indrajit Paul, Amit Ghosh, Michael Bolte, Michael Schmittel. Remote Control of the Synthesis of a
[2]Rotaxane and its Shuttling via Metal‐Ion Translocation. ChemistryOpen 2019, 8 (11) , 1355-1360.
https://doi.org/10.1002/open.201900293

388. Hongqian Sang, David Abbasi-Pérez, José Manuel Recio, Lev Kantorovich. Externally driven molecular
ratchets on a periodic potential surface: a rate equations approach. Physical Chemistry Chemical Physics
2019, 21 (42) , 23310-23319. https://doi.org/10.1039/C9CP03478A

389. Hendrik V. Schröder, Christoph A. Schalley. Electrochemically switchable rotaxanes: recent strides in new
directions. Chemical Science 2019, 10 (42) , 9626-9639. https://doi.org/10.1039/C9SC04118D

390. Yang Hu, Wei Wang, Rui Yao, Xu-Qing Wang, Yu-Xuan Wang, Bin Sun, Li-Jun Chen, Ying Zhang, Xiao-Li
Zhao, Lin Xu, Hong-Wei Tan, Yihua Yu, Xiaopeng Li, Hai-Bo Yang. Facile synthesis of diverse rotaxanes via
successive supramolecular transformations. Materials Chemistry Frontiers 2019, 3 (11) , 2397-2402.
https://doi.org/10.1039/C9QM00430K

391. Masaki Horie, Chi-Hsien Wang. Stimuli-responsive dynamic pseudorotaxane crystals. Materials
Chemistry Frontiers 2019, 3 (11) , 2258-2269. https://doi.org/10.1039/C9QM00483A

392. Jessica Groppi, Massimo Baroncini, Margherita Venturi, Serena Silvi, Alberto Credi. Design of photo-
activated molecular machines: highlights from the past ten years. Chemical Communications 2019, 55 (84) ,
12595-12602. https://doi.org/10.1039/C9CC06516D

393. Marcel Dommaschk, Javier Echavarren, David A. Leigh, Vanesa Marcos, Thomas A. Singleton. Dynamic
Control of Chiral Space Through Local Symmetry Breaking in a Rotaxane Organocatalyst. Angewandte Chemie
2019, 131 (42) , 15097-15100. https://doi.org/10.1002/ange.201908330

394. Li‐Shuo Zheng, Jie‐Shun Cui, Wei Jiang. Biomimetic Synchronized Motion of Two Interacting
Macrocycles in [3]Rotaxane‐Based Molecular Shuttles. Angewandte Chemie 2019, 131 (42) , 15280-15285.
https://doi.org/10.1002/ange.201910318

395. Marcel Dommaschk, Javier Echavarren, David A. Leigh, Vanesa Marcos, Thomas A. Singleton. Dynamic
Control of Chiral Space Through Local Symmetry Breaking in a Rotaxane Organocatalyst. Angewandte Chemie
International Edition 2019, 58 (42) , 14955-14958. https://doi.org/10.1002/anie.201908330

396. Li‐Shuo Zheng, Jie‐Shun Cui, Wei Jiang. Biomimetic Synchronized Motion of Two Interacting
Macrocycles in [3]Rotaxane‐Based Molecular Shuttles. Angewandte Chemie International Edition 2019, 58 (42)
, 15136-15141. https://doi.org/10.1002/anie.201910318

397. Xiao‐Dong Yang, Rui Zhu, Jian‐Ping Yin, Shuai Ma, Jing‐Wang Cui, Jie Zhang. Synergy of Electron
Transfer and Charge Transfer in the Control of Photodynamic Behavior of Coordination Polymers. Chemistry –
A European Journal 2019, 25 (57) , 13152-13156. https://doi.org/10.1002/chem.201902300

  
398. Daisuke Taura, Kaori Shimizu, Chiaki Yokota, Riho Ikeda, Yoshimasa Suzuki, Hiroki Iida, Naoki Ousaka,
Eiji Yashima. Fluorescent molecular spring that visualizes the extension and contraction motions of a double-
stranded helicate bearing terminal pyrene units triggered by release and binding of alkali metal ions. Chemical
Communications 2019, 55 (80) , 12084-12087. https://doi.org/10.1039/C9CC06126F

399. Lukas Pfeifer, Maximilian Scherübl, Maximilian Fellert, Wojciech Danowski, Jinling Cheng, Jasper Pol,
Ben L. Feringa. Photoe cient 2 nd generation molecular motors responsive to visible light. Chemical Science
2019, 10 (38) , 8768-8773. https://doi.org/10.1039/C9SC02150G

400. Shinji Toyota, Kenta Kawahata, Kota Sugahara, Tomohiro Oki, Tetsuo Iwanaga. Quadruple and Sextuple
Triptycene Gears in Macrocyclic Frameworks. Asian Journal of Organic Chemistry 2019, 8 (10) , 1919-1923.
https://doi.org/10.1002/ajoc.201900421

401. Andrea Sabatino, Emanuele Penocchio, Giulio Ragazzon, Alberto Credi, Diego Frezzato. Individual‐
Molecule Perspective Analysis of Chemical Reaction Networks: The Case of a Light‐Driven Supramolecular
Pump. Angewandte Chemie 2019, 131 (40) , 14479-14486. https://doi.org/10.1002/ange.201908026

402. Andrea Sabatino, Emanuele Penocchio, Giulio Ragazzon, Alberto Credi, Diego Frezzato. Individual‐
Molecule Perspective Analysis of Chemical Reaction Networks: The Case of a Light‐Driven Supramolecular
Pump. Angewandte Chemie International Edition 2019, 58 (40) , 14341-14348.
https://doi.org/10.1002/anie.201908026

403. Wai Ling Cheung-Lee, A. James Link. Genome mining for lasso peptides: past, present, and future.
Journal of Industrial Microbiology & Biotechnology 2019, 46 (9-10) , 1371-1379.
https://doi.org/10.1007/s10295-019-02197-z

404. Boris Atenas, Sergio Curilef. A solvable problem in statistical mechanics: The dipole-type Hamiltonian
mean eld model. Annals of Physics 2019, 409 , 167926. https://doi.org/10.1016/j.aop.2019.167926

405. Zhao Wang. Chirality-dependent motion transmission between aligned carbon nanotubes. Carbon 2019,
151 , 130-135. https://doi.org/10.1016/j.carbon.2019.05.051

406. Julián Valero, Mathias Centola, Yinzhou Ma, Marko Škugor, Ze Yu, Michael W. Haydell, Daniel Keppner,
Michael Famulok. Design, assembly, characterization, and operation of double-stranded interlocked DNA
nanostructures. Nature Protocols 2019, 14 (10) , 2818-2855. https://doi.org/10.1038/s41596-019-0198-7

407. Martijn van Galen, Ruben Higler, Joris Sprakel. Allosteric pathway selection in templated assembly.
Science Advances 2019, 5 (10) , eaaw3353. https://doi.org/10.1126/sciadv.aaw3353

408. Sofía Mena-Hernando, Emilio M. Pérez. Mechanically interlocked materials. Rotaxanes and catenanes
beyond the small molecule. Chemical Society Reviews 2019, 48 (19) , 5016-5032.
https://doi.org/10.1039/C8CS00888D

409. Chunying Fan, Jiabin Yao, Guojuan Li, Cheng Guo, Wanhua Wu, Dan Su, Zhihui Zhong, Dayang Zhou,
Yuliang Wang, Jason J. Chruma, Cheng Yang. Precise Manipulation of Temperature‐Driven Chirality Switching
of Molecular Universal Joints through Solvent Mixing. Chemistry – A European Journal 2019, 25 (54) , 12526-
12537. https://doi.org/10.1002/chem.201902676
410. M. Alexander Ardagh, Turan Birol, Qi Zhang, Omar A. Abdelrahman, Paul J. Dauenhauer. Catalytic
resonance theory: superVolcanoes, catalytic molecular pumps, and oscillatory steady state. Catalysis Science
& Technology 2019, 9 (18) , 5058-5076. https://doi.org/10.1039/C9CY01543D
  

411. Carlos Romero-Muñiz, Denís Paredes-Roibás, Antonio Hernanz, José María Gavira-Vallejo. A
comprehensive study of the molecular vibrations in solid-state benzylic amide [2]catenane. Physical Chemistry
Chemical Physics 2019, 21 (35) , 19538-19547. https://doi.org/10.1039/C9CP03053K

412. Shilin Yu, Nathan D. McClenaghan, Jean-Luc Pozzo. Photochromic rotaxanes and pseudorotaxanes.
Photochemical & Photobiological Sciences 2019, 18 (9) , 2102-2111. https://doi.org/10.1039/C9PP00057G

413. Yaping Feng, Haoyu Dai, Jianjun Chen, Xian Kong, Jinlei Yang, Lei Jiang. Geometric structure-guided
photo-driven ion current through asymmetric graphene oxide membranes. Journal of Materials Chemistry A
2019, 7 (35) , 20182-20186. https://doi.org/10.1039/C9TA07570D

414. Sudheer S. Kurup, Duleeka Wannipurage, Richard L. Lord, Stanislav Groysman. Tying the alkoxides
together: an iron complex of a new chelating bulky bis(alkoxide) demonstrates selectivity for coupling of non-
bulky aryl nitrenes. Chemical Communications 2019, 55 (72) , 10780-10783.
https://doi.org/10.1039/C9CC05319K

415. Yong‐Fei Yin, Meng‐Yan Yun, Lin Wu, Hong‐Ying Duan, Xia‐Min Jiang, Tian‐Guang Zhan, Jiecheng Cui, Li‐
Juan Liu, Kang‐Da Zhang. A Visible‐Light‐Induced Dynamic Mechanical Bond as a Linkage for Dynamic
Materials. Angewandte Chemie 2019, 131 (36) , 12835-12840. https://doi.org/10.1002/ange.201906761

416. Yong‐Fei Yin, Meng‐Yan Yun, Lin Wu, Hong‐Ying Duan, Xia‐Min Jiang, Tian‐Guang Zhan, Jiecheng Cui, Li‐
Juan Liu, Kang‐Da Zhang. A Visible‐Light‐Induced Dynamic Mechanical Bond as a Linkage for Dynamic
Materials. Angewandte Chemie International Edition 2019, 58 (36) , 12705-12710.
https://doi.org/10.1002/anie.201906761

417. Martin Drøhse Kilde, Rikke Kristensen, Gunnar Olsen, Jan O. Jeppesen, Mogens Brøndsted Nielsen.
Redox-Active Monopyrrolotetrathiafulvalene-Based Rotaxane Incorporating the
Dihydroazulene/Vinylheptafulvene Photo/Thermoswitch. European Journal of Organic Chemistry 2019, 2019
(31-32) , 5532-5539. https://doi.org/10.1002/ejoc.201900690

418. Mikhail Yu. Belikov, Mikhail Yu. Ievlev, Sergey V. Fedoseev, Oleg V. Ershov. Novel group of negative
photochromes containing a nitrile-rich acceptor: synthesis and photochromic properties. Research on
Chemical Intermediates 2019, 45 (9) , 4625-4636. https://doi.org/10.1007/s11164-019-03853-w

419. Federico Lancia, Alexander Ryabchun, Nathalie Katsonis. Life-like motion driven by arti cial molecular
machines. Nature Reviews Chemistry 2019, 3 (9) , 536-551. https://doi.org/10.1038/s41570-019-0122-2

420. Ohad Shpielberg, Takahiro Nemoto. Imitating nonequilibrium steady states using time-varying
equilibrium force in many-body diffusive systems. Physical Review E 2019, 100 (3)
https://doi.org/10.1103/PhysRevE.100.032104

421. Muraoka, Aoyama, Fujihara, Yamane, Hisaki, Miyata, Murata, Nakatsuji. Template-Free Synthesis of a
Phenanthroline-Containing [2]Rotaxane: A Reversible pH-Controllable Molecular Switch. Symmetry 2019, 11
(9) , 1137. https://doi.org/10.3390/sym11091137

422. Sudhakar Gaikwad, Merve Sinem Özer, Susnata Pramanik, Michael Schmittel. Three-state switching in a
double-pole change-over nanoswitch controlled by redox-dependent self-sorting. Organic & Biomolecular
Chemistry 2019, 17 (34) , 7956-7963. https://doi.org/10.1039/C9OB01456J

  
423. Munendra Pal Singh, Arup Tarai, Jubaraj B. Baruah. Photo-physical properties of salts of a di-topic
imidazole-tethered anthracene derivative in solid and solution. CrystEngComm 2019, 21 (33) , 4898-4909.
https://doi.org/10.1039/C9CE00791A

424. Harry A. Klein, Heike Kuhn, Paul D. Beer. Anion and pH dependent molecular motion by a halogen
bonding [2]rotaxane. Chemical Communications 2019, 55 (67) , 9975-9978.
https://doi.org/10.1039/C9CC04752B

425. Xiaochuan Li, Yaxin Wang, Caixia Jia, Hyorim Kim, Young-A Son. Photochromic reactivity induced by
electron distribution: active or inactive. Molecular Crystals and Liquid Crystals 2019, 689 (1) , 83-91.
https://doi.org/10.1080/15421406.2019.1597556

426. Elvira Pantuso, Giovanni Filpo, Fiore Pasquale Nicoletta. Light‐Responsive Polymer Membranes.
Advanced Optical Materials 2019, 7 (16) , 1900252. https://doi.org/10.1002/adom.201900252

427. Xiaofei Chen, Miriam Baumert, Roland Fröhlich, Markus Albrecht. Cation triggered spring-like helicates
based on ketone-substituted bis-catechol ligands. Journal of Inclusion Phenomena and Macrocyclic
Chemistry 2019, 94 (3-4) , 133-140. https://doi.org/10.1007/s10847-019-00888-9

428. Anjani Kumar Pandey, Sarnali Sanfui, Firoz Shah Tuglak Khan, Sankar Prasad Rath. Axial ligand
mediated switchable rotary motions in a ferrocene-bridged diiron(III) porphyrin dimer. Journal of
Organometallic Chemistry 2019, 894 , 28-38. https://doi.org/10.1016/j.jorganchem.2019.05.002

429. Zheng Li, Nan Song, Ying-Wei Yang. Stimuli-Responsive Drug-Delivery Systems Based on
Supramolecular Nanovalves. Matter 2019, 1 (2) , 345-368. https://doi.org/10.1016/j.matt.2019.05.019

430. A C Barato, R Chetrite, A Faggionato, D Gabrielli. A unifying picture of generalized thermodynamic


uncertainty relations. Journal of Statistical Mechanics: Theory and Experiment 2019, 2019 (8) , 084017.
https://doi.org/10.1088/1742-5468/ab3457

431. Wataru Setaka, Yusuke Inagaki, Kentaro Yamaguchi. Chemistry of Macrocage Molecules with a Bridged
π-Electron System as Crystalline Molecular Gyrotops. Journal of Synthetic Organic Chemistry, Japan 2019, 77
(8) , 813-822. https://doi.org/10.5059/yukigoseikyokaishi.77.813

432. Jeff M. Van Raden, Ramesh Jasti. How to make interlocked nanocarbons. Science 2019, 365 (6450) ,
216-217. https://doi.org/10.1126/science.aay2861

433. Chiara Biagini, Stephen D. P. Fielden, David A. Leigh, Fredrik Schaufelberger, Stefano Di Stefano, Dean
Thomas. Dissipative Catalysis with a Molecular Machine. Angewandte Chemie 2019, 131 (29) , 9981-9985.
https://doi.org/10.1002/ange.201905250

434. Chiara Biagini, Stephen D. P. Fielden, David A. Leigh, Fredrik Schaufelberger, Stefano Di Stefano, Dean
Thomas. Dissipative Catalysis with a Molecular Machine. Angewandte Chemie International Edition 2019, 58
(29) , 9876-9880. https://doi.org/10.1002/anie.201905250

435. Benjamin Riss-Yaw, Thomas-Xavier Métro, Frédéric Lamaty, Frédéric Coutrot. Association of liquid-
assisted grinding with aging accelerates the inherently slow slipping-on of a dibenzo-24-crown-8 over the N -
hydroxysuccinimide ester of an ammonium-containing thread. RSC Advances 2019, 9 (37) , 21587-21590.
https://doi.org/10.1039/C9RA05045K
  
436. Yuan Chen, Cheng Qian, Qian Zhao, Ming Cheng, Xinran Dong, Yue Zhao, Juli Jiang, Leyong Wang.
Adjustable chiral self-sorting and self-discriminating behaviour between diamond-like Tröger's base-linked
cryptands. Chemical Communications 2019, 55 (56) , 8072-8075. https://doi.org/10.1039/C9CC03577J

437. Danyu Xia, Xiaoqing Lv, Kexian Chen, Pi Wang. A [2]pseudorotaxane based on a pillar[6]arene and its
application in the construction of a metallosupramolecular polymer. Dalton Transactions 2019, 48 (27) , 9954-
9958. https://doi.org/10.1039/C9DT01713E

438. Jovana V. Milić, François Diederich. The Quest for Molecular Grippers: Photo‐Electric Control of
Molecular Gripping Machinery. Chemistry – A European Journal 2019, 25 (36) , 8440-8452.
https://doi.org/10.1002/chem.201900852

439. Merve S. Özer, Indrajit Paul, Abir Goswami, Michael Schmittel. Cation exchange reversibly switches rotor
speed and is monitored by a networked uorescent reporter. Dalton Transactions 2019, 48 (25) , 9043-9047.
https://doi.org/10.1039/C9DT01633C

440. Miku Oi, Ryo Takita, Junichiro Kanazawa, Atsuya Muranaka, Chao Wang, Masanobu Uchiyama.
Organocopper cross-coupling reaction for C–C bond formation on highly sterically hindered structures.
Chemical Science 2019, 10 (24) , 6107-6112. https://doi.org/10.1039/C9SC00891H

441. Liyun Wang, Danyu Xia, Jianbin Chao, Junjie Zhang, Xuehong Wei, Pi Wang. A
dimethoxypillar[5]arene/azastilbene host–guest recognition motif and its applications in the fabrication of
polypseudorotaxanes. Organic & Biomolecular Chemistry 2019, 17 (24) , 6038-6042.
https://doi.org/10.1039/C9OB00862D

442. David Abbasi-Pérez, Hongqian Sang, Lluïsa Pérez-García, Andrea Floris, David B. Amabilino, Rasmita
Raval, J. Manuel Recio, Lev Kantorovich. Controlling the preferential motion of chiral molecular walkers on a
surface. Chemical Science 2019, 10 (23) , 5864-5874. https://doi.org/10.1039/C9SC01135H

443. Changliang Ren, Feng Chen, Ruijuan Ye, Yong Siang Ong, Hongfang Lu, Su Seong Lee, Jackie Y. Ying,
Huaqiang Zeng. Molecular Swings as Highly Active Ion Transporters. Angewandte Chemie 2019, 131 (24) ,
8118-8122. https://doi.org/10.1002/ange.201901833

444. Kenji Caprice, Marion Pupier, Antonio Bauzá, Antonio Frontera, Fabien B. L. Cougnon. Synchronized
On/Off Switching of Four Binding Sites for Water in a Molecular Solomon Link. Angewandte Chemie 2019, 131
(24) , 8137-8141. https://doi.org/10.1002/ange.201902278

445. Changliang Ren, Feng Chen, Ruijuan Ye, Yong Siang Ong, Hongfang Lu, Su Seong Lee, Jackie Y. Ying,
Huaqiang Zeng. Molecular Swings as Highly Active Ion Transporters. Angewandte Chemie International
Edition 2019, 58 (24) , 8034-8038. https://doi.org/10.1002/anie.201901833

446. Kenji Caprice, Marion Pupier, Antonio Bauzá, Antonio Frontera, Fabien B. L. Cougnon. Synchronized
On/Off Switching of Four Binding Sites for Water in a Molecular Solomon Link. Angewandte Chemie
International Edition 2019, 58 (24) , 8053-8057. https://doi.org/10.1002/anie.201902278

447. Jatinder Singh, Dong Hwan Kim, Eun-Hee Kim, Nem Singh, Hyunuk Kim, Rizky Hadiputra, Jaehoon Jung,
Ki-Whan Chi. Selective and quantitative synthesis of a linear [3]catenane by two component coordination-
driven self-assembly. Chemical Communications 2019, 55 (48) , 6866-6869.
https://doi.org/10.1039/C9CC03336J
  
448. He-Ye Zhou, Ying Han, Qiang Shi, Chuan-Feng Chen. A Triply Operable Molecular Switch: Anion-,
Acid/Base- and Solvent-Responsive [2]Rotaxane. European Journal of Organic Chemistry 2019, 2019 (21) ,
3406-3411. https://doi.org/10.1002/ejoc.201801785

449. Yves Aeschi, Laurent Jucker, Daniel Häussinger, Marcel Mayor. Slow Formation of Pseudorotaxanes in
Water. European Journal of Organic Chemistry 2019, 2019 (21) , 3384-3390.
https://doi.org/10.1002/ejoc.201801864

450. Tainára Orlando, Paulo R. S. Salbego, Fellipe F. S. Farias, Gustavo H. Weimer, João P. P. Copetti, Helio G.
Bonacorso, Nilo Zanatta, Manfredo Hoerner, José Berná, Marcos A. P. Martins. Crystallization Mechanisms
Applied to Understand the Crystal Formation of Rotaxanes. European Journal of Organic Chemistry 2019,
2019 (21) , 3451-3463. https://doi.org/10.1002/ejoc.201801870

451. Maxime Gauthier, Frédéric Coutrot. Weinreb Amide as Secondary Station for the Dibenzo-24-crown-8 in
a Molecular Shuttle. European Journal of Organic Chemistry 2019, 2019 (21) , 3391-3395.
https://doi.org/10.1002/ejoc.201900046

452. Adrian Wolf, Juan-José Cid, Emilie Moulin, Frédéric Niess, Guangyan Du, Antoine Goujon, Eric Busseron,
Adrian Ruff, Sabine Ludwigs, Nicolas Giuseppone. Unsymmetric Bistable [ c 2]Daisy Chain Rotaxanes which
Combine Two Types of Electroactive Stoppers. European Journal of Organic Chemistry 2019, 2019 (21) ,
3421-3432. https://doi.org/10.1002/ejoc.201900179

453. Michael Franz, Johanna A. Januszewski, Frank Hampel, Rik R. Tykwinski. [3]Rotaxanes with Mixed Axles:
Polyynes and Cumulenes. European Journal of Organic Chemistry 2019, 2019 (21) , 3503-3512.
https://doi.org/10.1002/ejoc.201900188

454. Paulo R. S. Salbego, Tainára Orlando, Fellipe F. S. Farias, Helio G. Bonacorso, Marcos A. P. Martins.
[2]Rotaxanes Bearing a Tetralactam Macrocycle: The Role of a Trifurcated Hydrogen Bond in the Crystalline
State. European Journal of Organic Chemistry 2019, 2019 (21) , 3464-3471.
https://doi.org/10.1002/ejoc.201900216

455. Melis Özkan, Yağmur Keser, Seyed Ehsan Hadi, Dönüs Tuncel. A [5]Rotaxane-Based Photosensitizer for
Photodynamic Therapy. European Journal of Organic Chemistry 2019, 2019 (21) , 3534-3541.
https://doi.org/10.1002/ejoc.201900278

456. Ruth Dorel, Ben L. Feringa. Photoswitchable catalysis based on the isomerisation of double bonds.
Chemical Communications 2019, 55 (46) , 6477-6486. https://doi.org/10.1039/C9CC01891C

457. Monika Schildhauer, Florian Rott, Stefan Thumser, Peter Mayer, Regina de Vivie‐Riedle, Henry Dube. A
Prospective Ultrafast Hemithioindigo Molecular Motor. ChemPhotoChem 2019, 3 (6) , 365-371.
https://doi.org/10.1002/cptc.201900074

458. Moisey I. Belinsky. The in-plane spin helicity of coplanar helical spin con gurations of frustrated single
trimer V3 and Cu3 nanomagnets, inversion (switching) of spin helicity. Chemical Physics 2019, 522 , 267-278.
https://doi.org/10.1016/j.chemphys.2019.01.014

459. Mathieu Denis, James E.M. Lewis, Florian Modicom, Stephen M. Goldup. An Auxiliary Approach for the
Stereoselective Synthesis of Topologically Chiral Catenanes. Chem 2019, 5 (6) , 1512-1520.
https://doi.org/10.1016/j.chempr.2019.03.008

  
460. Arshia Zarrin, David A. Sivak, Aidan I. Brown. Breaking time-reversal symmetry for ratchet models of
molecular machines. Physical Review E 2019, 99 (6) https://doi.org/10.1103/PhysRevE.99.062127

461. Timur Koyuk, Udo Seifert. Operationally Accessible Bounds on Fluctuations and Entropy Production in
Periodically Driven Systems. Physical Review Letters 2019, 122 (23)
https://doi.org/10.1103/PhysRevLett.122.230601

462. Yuan Chen, Ming Cheng, Benkun Hong, Qian Zhao, Cheng Qian, Juli Jiang, Shuhua Li, Chen Lin, Leyong
Wang. N-Centered Chiral Self-Sorting and Supramolecular Helix of Tröger's Base-Based Dimeric Macrocycles
in Crystalline State. Frontiers in Chemistry 2019, 7 https://doi.org/10.3389/fchem.2019.00383

463. Nadia Manganaro, Ilenia Pisagatti, Anna Notti, Melchiorre F. Parisi, Giuseppe Gattuso. Self-sorting
assembly of a calixarene/crown ether polypseudorotaxane gated by ion-pairing. New Journal of Chemistry
2019, 43 (21) , 7936-7940. https://doi.org/10.1039/C9NJ01583C

464. Zhisong Wang, Ruizheng Hou, Iong Ying Loh. Track-walking molecular motors: a new generation beyond
bridge-burning designs. Nanoscale 2019, 11 (19) , 9240-9263. https://doi.org/10.1039/C9NR00033J

465. Sheng‐Ying Chou, Hiroshi Masai, Susumu Tsuda, Jun Terao. Synthetic Methodology for Structurally
De ned and Insulated Molecular Wires Bearing Non‐centrosymmetric Conjugated Axle Components via
Iterative Intramolecular Slippage. Chemistry – An Asian Journal 2019, 14 (10) , 1667-1671.
https://doi.org/10.1002/asia.201801706

466. Maria Richter, Jesús González-Vázquez, Zdeněk Mašín, Danilo S. Brambila, Alex G. Harvey, Felipe
Morales, Fernando Martín. Ultrafast imaging of laser-controlled non-adiabatic dynamics in NO 2 from time-
resolved photoelectron emission. Physical Chemistry Chemical Physics 2019, 21 (19) , 10038-10051.
https://doi.org/10.1039/C9CP00649D

467. Mateusz Woźny, Karolina M. Tomczyk, Agnieszka Więckowska, Szymon Sutuła, Damian Trzybiński,
Krzysztof Woźniak, Bohdan Korybut-Daszkiewicz. The in uence of metal-complexing macrocycle size on
intramolecular movement in rotaxanes. Dalton Transactions 2019, 48 (19) , 6546-6557.
https://doi.org/10.1039/C9DT00083F

468. Xu-Qing Wang, Wei Wang, Wei-Jian Li, Yi Qin, Guang-Qiang Yin, Wei-Ling Jiang, Xiaopeng Li, Shuai Wu,
Hai-Bo Yang. Rotaxane-branched dendrimers with aggregation-induced emission behavior. Organic Chemistry
Frontiers 2019, 6 (10) , 1686-1691. https://doi.org/10.1039/C9QO00308H

469. Giorgio Baggi, Lorenzo Casimiro, Massimo Baroncini, Serena Silvi, Alberto Credi, Stephen J. Loeb.
Threading-gated photochromism in [2]pseudorotaxanes. Chemical Science 2019, 10 (19) , 5104-5113.
https://doi.org/10.1039/C9SC00913B

470. Alexander G. Martynov, Evgeniya A. Safonova, Aslan Yu. Tsivadze, Yulia G. Gorbunova. Functional
molecular switches involving tetrapyrrolic macrocycles. Coordination Chemistry Reviews 2019, 387 , 325-347.
https://doi.org/10.1016/j.ccr.2019.02.004

471. Stefan Borsley, Marius M. Haugland, Samuel Oldknow, James A. Cooper, Michael J. Burke, Aaron Scott,
William Grantham, Julia Vallejo, Euan K. Brechin, Paul J. Lusby, Scott L. Cockroft. Electrostatic Forces in Field-
Perturbed Equilibria: Nanopore Analysis of Cage Complexes. Chem 2019, 5 (5) , 1275-1292.
https://doi.org/10.1016/j.chempr.2019.03.004
472. Wojciech Danowski, Thomas van Leeuwen, Shaghayegh Abdolahzadeh, Diederik Roke, Wesley R.
 
Browne, Sander J. Wezenberg, Ben L. Feringa. Unidirectional rotary motion in a metal–organic framework.

Nature Nanotechnology 2019, 14 (5) , 488-494. https://doi.org/10.1038/s41565-019-0401-6

473. Arthur H. G. David, Pablo García‐Cerezo, Araceli G. Campaña, Francisco Santoyo‐González, Victor
Blanco. [2]Rotaxane End‐Capping Synthesis by Click Michael‐Type Addition to the Vinyl Sulfonyl Group.
Chemistry – A European Journal 2019, 25 (24) , 6170-6179. https://doi.org/10.1002/chem.201900156

474. Jason Y. C. Lim, Nattawut Yuntawattana, Paul D. Beer, Charlotte K. Williams. Isoselective Lactide Ring
Opening Polymerisation using [2]Rotaxane Catalysts. Angewandte Chemie 2019, 131 (18) , 6068-6072.
https://doi.org/10.1002/ange.201901592

475. Jason Y. C. Lim, Nattawut Yuntawattana, Paul D. Beer, Charlotte K. Williams. Isoselective Lactide Ring
Opening Polymerisation using [2]Rotaxane Catalysts. Angewandte Chemie International Edition 2019, 58 (18) ,
6007-6011. https://doi.org/10.1002/anie.201901592

476. Yangyang Yang, Shiwei Zhang, Shengtao Yao, Rizhao Pan, Kumi Hidaka, Tomoko Emura, Chunhai Fan,
Hiroshi Sugiyama, Yufang Xu, Masayuki Endo, Xuhong Qian. Programming Rotary Motions with a Hexagonal
DNA Nanomachine. Chemistry – A European Journal 2019, 25 (20) , 5158-5162.
https://doi.org/10.1002/chem.201900221

477. Bo Li, Shichao Jiang, Jingshuo Gao, Xueting Wu, Jiajie Deng, Linmeng Zhang, Zhipeng Yu. Dual
Colorimetric/Fluorometric Double‐Throw pH‐Switches: The Dimroth Rearrangement of N , 9 ‐Diaryl 8‐
Azaadenines. ChemPlusChem 2019, 84 (4) , 427-431. https://doi.org/10.1002/cplu.201900117

478. Lifei Zheng, Hui Zhao, Yanxiao Han, Haibin Qian, Lela Vukovic, Jasmin Mecinović, Petr Král, Wilhelm T. S.
Huck. Catalytic transport of molecular cargo using diffusive binding along a polymer track. Nature Chemistry
2019, 11 (4) , 359-366. https://doi.org/10.1038/s41557-018-0204-7

479. Suehyun Park, Jeongeun Song, Jun Soo Kim. In silico construction of a exibility-based DNA Brownian
ratchet for directional nanoparticle delivery. Science Advances 2019, 5 (4) , eaav4943.
https://doi.org/10.1126/sciadv.aav4943

480. Antoine Antoine, Emilie Moulin, Gad Fuks, Nicolas Giuseppone. [c2]Daisy Chain Rotaxanes as Molecular
Muscles. CCS Chemistry 2019, , 83-96. https://doi.org/10.31635/ccschem.019.20180023

481. Kirill Nikitin, Ryan O'Gara. Mechanisms and Beyond: Elucidation of Fluxional Dynamics by Exchange
NMR Spectroscopy. Chemistry – A European Journal 2019, 25 (18) , 4551-4589.
https://doi.org/10.1002/chem.201804123

482. Parvej Alam, Nelson L. C. Leung, Yanhua Cheng, Haoke Zhang, Junkai Liu, Wenjie Wu, Ryan T. K. Kwok,
Jacky W. Y. Lam, Herman H. Y. Sung, Ian D. Williams, Ben Zhong Tang. Spontaneous and Fast Molecular
Motion at Room Temperature in the Solid State. Angewandte Chemie 2019, 131 (14) , 4584-4588.
https://doi.org/10.1002/ange.201813554

483. Parvej Alam, Nelson L. C. Leung, Yanhua Cheng, Haoke Zhang, Junkai Liu, Wenjie Wu, Ryan T. K. Kwok,
Jacky W. Y. Lam, Herman H. Y. Sung, Ian D. Williams, Ben Zhong Tang. Spontaneous and Fast Molecular
Motion at Room Temperature in the Solid State. Angewandte Chemie International Edition 2019, 58 (14) ,
4536-4540. https://doi.org/10.1002/anie.201813554
484. Yukari Okuma, Toshihiro Tsukamoto, Takayuki Inagaki, Shinobu Miyagawa, Masaki Kimura, Masaya
Naito, Hikaru Takaya, Tsuneomi Kawasaki, Yuji Tokunaga. Rotational isomerism of the amide units in
  
rotaxanes based on a cyclic tetraamide and secondary ammonium ions. Organic Chemistry Frontiers 2019, 6
(7) , 1002-1009. https://doi.org/10.1039/C9QO00096H

485. Xiang Wang, Quan Gan, Barbara Wicher, Yann Ferrand, Ivan Huc. Directional Threading and Sliding of a
Dissymmetrical Foldamer Helix on Dissymmetrical Axles. Angewandte Chemie 2019, 131 (13) , 4249-4253.
https://doi.org/10.1002/ange.201813125

486. Katsuya Ichihashi, Daisuke Konno, Kseniya Yu. Maryunina, Katsuya Inoue, Kazuhiro Toyoda, Shogo
Kawaguchi, Yoshiki Kubota, Yoko Tatewaki, Tomoyuki Akutagawa, Takayoshi Nakamura, Sadafumi Nishihara.
Selective Ion Exchange in Supramolecular Channels in the Crystalline State. Angewandte Chemie 2019, 131
(13) , 4213-4216. https://doi.org/10.1002/ange.201813709

487. Alberto Credi. Eine molekulare Seilbahn für den transmembranären Ionentransport. Angewandte Chemie
2019, 131 (13) , 4152-4155. https://doi.org/10.1002/ange.201814333

488. Xiang Wang, Quan Gan, Barbara Wicher, Yann Ferrand, Ivan Huc. Directional Threading and Sliding of a
Dissymmetrical Foldamer Helix on Dissymmetrical Axles. Angewandte Chemie International Edition 2019, 58
(13) , 4205-4209. https://doi.org/10.1002/anie.201813125

489. Katsuya Ichihashi, Daisuke Konno, Kseniya Yu. Maryunina, Katsuya Inoue, Kazuhiro Toyoda, Shogo
Kawaguchi, Yoshiki Kubota, Yoko Tatewaki, Tomoyuki Akutagawa, Takayoshi Nakamura, Sadafumi Nishihara.
Selective Ion Exchange in Supramolecular Channels in the Crystalline State. Angewandte Chemie International
Edition 2019, 58 (13) , 4169-4172. https://doi.org/10.1002/anie.201813709

490. Alberto Credi. A Molecular Cable Car for Transmembrane Ion Transport. Angewandte Chemie
International Edition 2019, 58 (13) , 4108-4110. https://doi.org/10.1002/anie.201814333

491. Francesco Colizzi, Adam Hospital, Sanja Zivanovic, Modesto Orozco. Predicting the Limit of
Intramolecular Hydrogen Bonding with Classical Molecular Dynamics. Angewandte Chemie 2019, 131 (12) ,
3799-3803. https://doi.org/10.1002/ange.201810922

492. Bin Li, Bin Wang, Xiayang Huang, Lu Dai, Lei Cui, Jian Li, Xueshun Jia, Chunju Li. Terphen[ n ]arenes and
Quaterphen[ n ]arenes ( n =3–6): One‐Pot Synthesis, Self‐Assembly into Supramolecular Gels, and Iodine
Capture. Angewandte Chemie 2019, 131 (12) , 3925-3929. https://doi.org/10.1002/ange.201813972

493. Francesco Colizzi, Adam Hospital, Sanja Zivanovic, Modesto Orozco. Predicting the Limit of
Intramolecular Hydrogen Bonding with Classical Molecular Dynamics. Angewandte Chemie International
Edition 2019, 58 (12) , 3759-3763. https://doi.org/10.1002/anie.201810922

494. Bin Li, Bin Wang, Xiayang Huang, Lu Dai, Lei Cui, Jian Li, Xueshun Jia, Chunju Li. Terphen[ n ]arenes and
Quaterphen[ n ]arenes ( n =3–6): One‐Pot Synthesis, Self‐Assembly into Supramolecular Gels, and Iodine
Capture. Angewandte Chemie International Edition 2019, 58 (12) , 3885-3889.
https://doi.org/10.1002/anie.201813972

495. Arun Richard Chandrasekaran, Ken Halvorsen. Controlled disassembly of a DNA tetrahedron using
strand displacement. Nanoscale Advances 2019, 1 (3) , 969-972. https://doi.org/10.1039/C8NA00340H

496. Xiaoyong Jia, Cancan Shao, Xin Bai, Qiaohui Zhou, Bin Wu, Linjun Wang, Bingbing Yue, Haiming Zhu,
Liangliang Zhu. Photoexcitation-controlled self-recoverable molecular aggregation for icker
phosphorescence. Proceedings of the National Academy of Sciences 2019, 116 (11) , 4816-4821.
https://doi.org/10.1073/pnas.1821991116
  
497. Hao Shi, Kun Zhang, Rui-Lian Lin, Wen-Qi Sun, Xiang-Feng Chu, Xin-Hua Liu, Jing-Xin Liu. pH-Controlled
Multiple Interconversion between Cucurbit[7]uril-Based Molecular Shuttle, [3]Pseudorotaxane and
[2]Pseudorotaxane. Asian Journal of Organic Chemistry 2019, 8 (3) , 339-343.
https://doi.org/10.1002/ajoc.201800708

498. Paul M. Bogie, Tabitha F. Miller, Richard J. Hooley. Synthesis and Applications of Endohedrally
Functionalized Metal‐Ligand Cage Complexes. Israel Journal of Chemistry 2019, 59 (3-4) , 130-139.
https://doi.org/10.1002/ijch.201800067

499. Michael Schmittel. Dynamic Functional Molecular Systems: From Supramolecular Structures to Multi‐
Component Machinery and to Molecular Cybernetics. Israel Journal of Chemistry 2019, 59 (3-4) , 197-208.
https://doi.org/10.1002/ijch.201800124

500. Igor Goychuk. Perfect anomalous transport of subdiffusive cargos by molecular motors in viscoelastic
cytosol. Biosystems 2019, 177 , 56-65. https://doi.org/10.1016/j.biosystems.2018.11.004

501. Yu. A. Makhnovskii. Effect of particle size oscillations on drift and diffusion along a periodically
corrugated channel. Physical Review E 2019, 99 (3) https://doi.org/10.1103/PhysRevE.99.032102

502. Shohei Saito. Flapping Molecules for Photofunctional Materials. 2019,,, 17-51.
https://doi.org/10.1002/9783527823987.vol3_c2

503. Toshikazu Takata. Stimuli-Responsive Molecular and Macromolecular Systems Controlled by Rotaxane
Molecular Switches. Bulletin of the Chemical Society of Japan 2019, 92 (2) , 409-426.
https://doi.org/10.1246/bcsj.20180330

504. Kerstin Hoffmann, Peter Mayer, Henry Dube. A hemithioindigo molecular motor for metal surface
attachment. Organic & Biomolecular Chemistry 2019, 17 (7) , 1979-1983.
https://doi.org/10.1039/C8OB02424C

505. Yong Dou, Kiran Dhatt-Gauthier, Kyle J.M. Bishop. Thermodynamic costs of dynamic function in active
soft matter. Current Opinion in Solid State and Materials Science 2019, 23 (1) , 28-40.
https://doi.org/10.1016/j.cossms.2018.11.002

506. Lidan Li, Na Li, Shengnan Fu, Yingnan Deng, Changyuan Yu, Xin Su. Base excision repair-inspired DNA
motor powered by intracellular apurinic/apyrimidinic endonuclease. Nanoscale 2019, 11 (3) , 1343-1350.
https://doi.org/10.1039/C8NR07813K

507. Elena Pazos, Paula Novo, Carlos Peinador, Angel E. Kaifer, Marcos D. García. Supramolekulare Schalter
auf der Basis von Cucurbit[8]uril (CB[8]). Angewandte Chemie 2019, 131 (2) , 409-422.
https://doi.org/10.1002/ange.201806575

508. Elena Pazos, Paula Novo, Carlos Peinador, Angel E. Kaifer, Marcos D. García. Cucurbit[8]uril (CB[8])-
Based Supramolecular Switches. Angewandte Chemie International Edition 2019, 58 (2) , 403-416.
https://doi.org/10.1002/anie.201806575
509. Yves Aeschi, Sylvie Drayss‐Orth, Michal Valášek, Daniel Häussinger, Marcel Mayor. Aqueous Assembly of
Zwitterionic Daisy Chains. Chemistry – A European Journal 2019, 25 (1) , 285-295.
https://doi.org/10.1002/chem.201803944
  

510. Bianca Maranescu, Aurelia Visa. Metal-Organic Framework Composites IPMC Sensors and Actuators.
2019,,, 1-18. https://doi.org/10.1007/978-3-030-13728-1_1

511. Cai-Xin Zhao, Qi Zhang, Gábor London, Da-Hui Qu. Functional Rotaxanes. 2019,,, 1-34.
https://doi.org/10.1007/978-981-13-1744-6_12-1

512. Bin Li, Yiliang Wang, Chunju Li. Biphen[n]arenes: Synthesis and Host–Guest Properties. 2019,,, 1-29.
https://doi.org/10.1007/978-981-13-1744-6_13-1

513. Wensheng Cai, Haohao Fu. Molecular Simulations of Supramolecular Architectures. 2019,,, 1-27.
https://doi.org/10.1007/978-981-13-1744-6_45-1

514. Philipp Langer, Lixu Yang, Constance R. Pfeiffer, William Lewis, Neil R. Champness. Restricting shuttling
in bis(imidazolium)…pillar[5]arene rotaxanes using metal coordination. Dalton Transactions 2019, 48 (1) , 58-
64. https://doi.org/10.1039/C8DT04096F

515. Abdulselam Adam, Saber Mehrparvar, Gebhard Haberhauer. An azobenzene container showing a
de nite folding – synthesis and structural investigation. Beilstein Journal of Organic Chemistry 2019, 15 ,
1534-1544. https://doi.org/10.3762/bjoc.15.156

516. Chaocao Lu, Bu Htan, Chunmiao Ma, Rong-Zhen Liao, Quan Gan. Acylhydrazone Switches: E/Z Stability
Reversed by Introduction of Hydrogen Bonds. European Journal of Organic Chemistry 2018, 2018 (48) , 7046-
7050. https://doi.org/10.1002/ejoc.201801466

517. Ganhua Xie, Pei Li, Zhiju Zhao, Xiang‐Yu Kong, Zhen Zhang, Kai Xiao, Huanting Wang, Liping Wen, Lei
Jiang. Bacteriorhodopsin‐Inspired Light‐Driven Arti cial Molecule Motors for Transmembrane Mass
Transportation. Angewandte Chemie 2018, 130 (51) , 16950-16954.
https://doi.org/10.1002/ange.201809627

518. Ganhua Xie, Pei Li, Zhiju Zhao, Xiang‐Yu Kong, Zhen Zhang, Kai Xiao, Huanting Wang, Liping Wen, Lei
Jiang. Bacteriorhodopsin‐Inspired Light‐Driven Arti cial Molecule Motors for Transmembrane Mass
Transportation. Angewandte Chemie International Edition 2018, 57 (51) , 16708-16712.
https://doi.org/10.1002/anie.201809627

519. Rakesh Rajegowda, Sarith P. Sathian. Analysing thermophoretic transport of water for designing
nanoscale-pumps. Physical Chemistry Chemical Physics 2018, 20 (48) , 30321-30330.
https://doi.org/10.1039/C8CP05521A

520. Xiaoxiong Li, Jason Y. C. Lim, Paul D. Beer. Acid-Regulated Switching of Metal Cation and Anion Guest
Binding in Halogen-Bonding Rotaxanes. Chemistry - A European Journal 2018, 24 (67) , 17788-17795.
https://doi.org/10.1002/chem.201803902

521. Hao Chen, Jing Cao, Panpan Zhou, Xiang Li, Yujie Xie, Weisheng Liu, Yu Tang. Multiplex recognition and
logic devices for molecular robot prototype based on an europium(iii)–cyclen system. Biosensors and
Bioelectronics 2018, 122 , 1-7. https://doi.org/10.1016/j.bios.2018.09.029
522. Yan Liu, Jian Luo. Nonadiabatic dynamics simulation of photoisomerization mechanism of photoswitch
azodicarboxamide: Hydrogen bonding effects. Journal of Photochemistry and Photobiology A: Chemistry
  
2018, 367 , 236-239. https://doi.org/10.1016/j.jphotochem.2018.08.040

523. Walace D. do Pim, Tatiana A. Ribeiro-Santos, Ingrid S. Jardim, Mateus C.M. de Castro, Adriano H. Braga,
Gustavo M. do Nascimento, Ildefonso Binatti, Humberto O. Stumpf, Eudes Lorençon, Maria H. Araujo, Cynthia
L.M. Pereira. Bistable copper(II) metallosurfactant as molecular machine for the preparation of hybrid silica-
based porous materials. Materials & Design 2018, 160 , 876-885.
https://doi.org/10.1016/j.matdes.2018.10.023

524. Takehiro Hirao, Dong Sub Kim, Xiaodong Chi, Vincent M. Lynch, Kazuaki Ohara, Jung Su Park, Kentaro
Yamaguchi, Jonathan L. Sessler. Control over multiple molecular states with directional changes driven by
molecular recognition. Nature Communications 2018, 9 (1) https://doi.org/10.1038/s41467-018-03220-0

525. Maximilian J. Urban, Steffen Both, Chao Zhou, Anton Kuzyk, Klas Lindfors, Thomas Weiss, Na Liu. Gold
nanocrystal-mediated sliding of doublet DNA origami laments. Nature Communications 2018, 9 (1)
https://doi.org/10.1038/s41467-018-03882-w

526. Yingying Wu, Guangxia Wang, Qiaolian Li, Junfeng Xiang, Hua Jiang, Ying Wang. A multistage rotational
speed changing molecular rotor regulated by pH and metal cations. Nature Communications 2018, 9 (1)
https://doi.org/10.1038/s41467-018-04323-4

527. Taisuke Matsuno, Yusuke Nakai, Sota Sato, Yutaka Maniwa, Hiroyuki Isobe. Ratchet-free solid-state
inertial rotation of a guest ball in a tight tubular host. Nature Communications 2018, 9 (1)
https://doi.org/10.1038/s41467-018-04325-2

528. Aaron Gerwien, Monika Schildhauer, Stefan Thumser, Peter Mayer, Henry Dube. Direct evidence for hula
twist and single-bond rotation photoproducts. Nature Communications 2018, 9 (1)
https://doi.org/10.1038/s41467-018-04928-9

529. R. L. Greenaway, V. Santolini, M. J. Bennison, B. M. Alston, C. J. Pugh, M. A. Little, M. Miklitz, E. G. B.


Eden-Rump, R. Clowes, A. Shakil, H. J. Cuthbertson, H. Armstrong, M. E. Briggs, K. E. Jelfs, A. I. Cooper. High-
throughput discovery of organic cages and catenanes using computational screening fused with robotic
synthesis. Nature Communications 2018, 9 (1) https://doi.org/10.1038/s41467-018-05271-9

530. Xu-Qing Wang, Wei Wang, Wei-Jian Li, Li-Jun Chen, Rui Yao, Guang-Qiang Yin, Yu-Xuan Wang, Ying
Zhang, Junlin Huang, Hongwei Tan, Yihua Yu, Xiaopeng Li, Lin Xu, Hai-Bo Yang. Dual stimuli-responsive
rotaxane-branched dendrimers with reversible dimension modulation. Nature Communications 2018, 9 (1)
https://doi.org/10.1038/s41467-018-05670-y

531. Akihiko Nakamura, Kei-ichi Okazaki, Tadaomi Furuta, Minoru Sakurai, Ryota Iino. Processive chitinase is
Brownian monorail operated by fast catalysis after peeling rail from crystalline chitin. Nature Communications
2018, 9 (1) https://doi.org/10.1038/s41467-018-06362-3

532. Teresa Naranjo, Kateryna M. Lemishko, Sara de Lorenzo, Álvaro Somoza, Felix Ritort, Emilio M. Pérez,
Borja Ibarra. Dynamics of individual molecular shuttles under mechanical force. Nature Communications
2018, 9 (1) https://doi.org/10.1038/s41467-018-06905-8

533. Ye Yuan, Ghaneema N. Abuhaimed, Qingkun Liu, Ivan I. Smalyukh. Self-assembled nematic colloidal
motors powered by light. Nature Communications 2018, 9 (1) https://doi.org/10.1038/s41467-018-07518-x
534. Shunsuke Tanaka, Kingo Takiguchi, Masahito Hayashi. Repetitive stretching of giant liposomes utilizing
the nematic alignment of con ned actin. Communications Physics 2018, 1 (1)
https://doi.org/10.1038/s42005-018-0019-2
  

535. Martin Šiler, Luca Ornigotti, Oto Brzobohatý, Petr Jákl, Artem Ryabov, Viktor Holubec, Pavel Zemánek,
Radim Filip. Diffusing up the Hill: Dynamics and Equipartition in Highly Unstable Systems. Physical Review
Letters 2018, 121 (23) https://doi.org/10.1103/PhysRevLett.121.230601

536. Xu-Qing Wang, Wei-Jian Li, Wei Wang, Hai-Bo Yang. Heterorotaxanes. Chemical Communications 2018,
54 (95) , 13303-13318. https://doi.org/10.1039/C8CC07283C

537. Hendrik V. Schröder, Amel Mekic, Henrik Hupatz, Sebastian Sobottka, Felix Witte, Leonhard H. Urner,
Marius Gaedke, Kevin Pagel, Biprajit Sarkar, Beate Paulus, Christoph A. Schalley. Switchable synchronisation of
pirouetting motions in a redox-active [3]rotaxane. Nanoscale 2018, 10 (45) , 21425-21433.
https://doi.org/10.1039/C8NR05534C

538. Massimo Baroncini, Martina Canton, Lorenzo Casimiro, Stefano Corra, Jessica Groppi, Marcello La Rosa,
Serena Silvi, Alberto Credi. Photoactive Molecular-Based Devices, Machines and Materials: Recent Advances.
European Journal of Inorganic Chemistry 2018, 2018 (42) , 4589-4603.
https://doi.org/10.1002/ejic.201800923

539. Alena N. Kulakova, Alexey N. Bilyachenko, Alexander A. Korlyukov, Lidia S. Shul'pina, Xavier Bantreil,
Frédéric Lamaty, Elena S. Shubina, Mikhail M. Levitsky, Nikolay S. Ikonnikov, Georgiy B. Shul'pin. A new “bicycle
helmet”-like copper( ii ),sodiumphenylsilsesquioxane. Synthesis, structure and catalytic activity. Dalton
Transactions 2018, 47 (44) , 15666-15669. https://doi.org/10.1039/C8DT03209B

540. Mateusz Woźny, Agnieszka Więckowska, Damian Trzybiński, Szymon Sutuła, Sławomir Domagała,
Krzysztof Woźniak. [3]rotaxanes composed of two dibenzo-24-crown-8 ether wheels and an azamacrocyclic
complex. Dalton Transactions 2018, 47 (44) , 15845-15856. https://doi.org/10.1039/C8DT03225D

541. Michael A. Jinks, Alberto de Juan, Mathieu Denis, Catherine J. Fletcher, Marzia Galli, Ellen M. G.
Jamieson, Florian Modicom, Zhihui Zhang, Stephen M. Goldup. Stereoselective Synthesis of Mechanically
Planar Chiral Rotaxanes. Angewandte Chemie 2018, 130 (45) , 15022-15026.
https://doi.org/10.1002/ange.201808990

542. Michael A. Jinks, Alberto de Juan, Mathieu Denis, Catherine J. Fletcher, Marzia Galli, Ellen M. G.
Jamieson, Florian Modicom, Zhihui Zhang, Stephen M. Goldup. Stereoselective Synthesis of Mechanically
Planar Chiral Rotaxanes. Angewandte Chemie International Edition 2018, 57 (45) , 14806-14810.
https://doi.org/10.1002/anie.201808990

543. Raffaello Papadakis, Ioanna Deligkiozi, Hu Li. Photoconductive Interlocked Molecules and
Macromolecules. 2018,,https://doi.org/10.5772/intechopen.79798

544. Jianfei Huang, Akchheta Karki, Viktor V. Brus, Yuanyuan Hu, Hung Phan, Alexander T. Lill, Ming Wang,
Guillermo C. Bazan, Thuc-Quyen Nguyen. Solution-Processed Ion-Free Organic Ratchets with Asymmetric
Contacts. Advanced Materials 2018, 30 (46) , 1804794. https://doi.org/10.1002/adma.201804794

545. Qi Zhang, Si-Jia Rao, Tao Xie, Xin Li, Tian-Yi Xu, Da-Wei Li, Da-Hui Qu, Yi-Tao Long, He Tian. Muscle-like
Arti cial Molecular Actuators for Nanoparticles. Chem 2018, 4 (11) , 2670-2684.
https://doi.org/10.1016/j.chempr.2018.08.030
546. Bin Qi, Xin Li, Liang Sun, Bo Chen, Hao Chen, Chenchen Wu, Haibo Zhang, Xiaohai Zhou. In situ synthesis
of ultra ne metal clusters triggered by dodecaborate supramolecular organic frameworks. Nanoscale 2018, 10
(42) , 19846-19853. https://doi.org/10.1039/C8NR06917D
  

547. Yingying Chai, Xinglong Zhou, Changfu Li, Beibei Ma, Zhen Shen, Ridong Huang, Hai Chen, Bojiang Chen,
Weimin Li, Yang He. Supermacrocyclic Assemblies by Hydrogen‐Bond Codes of C7‐Phenol Pyrazolo and
Pyrrolo Derivatives of Adenine. Chemistry – A European Journal 2018, 24 (58) , 15495-15501.
https://doi.org/10.1002/chem.201803429

548. Tae Y. Kim, Roan A. S. Vasdev, Dan Preston, James D. Crowley. Strategies for Reversible Guest Uptake
and Release from Metallosupramolecular Architectures. Chemistry - A European Journal 2018, 24 (56) ,
14878-14890. https://doi.org/10.1002/chem.201802081

549. Luca Catalano, Panče Naumov. Exploiting rotational motion in molecular crystals. CrystEngComm 2018,
20 (39) , 5872-5883. https://doi.org/10.1039/C8CE00420J

550. Qiaochun Wang, Dizhi Chen, He Tian. Arti cial molecular machines that can perform work. Science
China Chemistry 2018, 61 (10) , 1261-1273. https://doi.org/10.1007/s11426-018-9267-3

551. Minh T. Nguyen, Daniel P. Ferris, Cristian Pezzato, Yuping Wang, J. Fraser Stoddart. Densely Charged
Dodecacationic [3]- and Tetracosacationic Radial [5]Catenanes. Chem 2018, 4 (10) , 2329-2344.
https://doi.org/10.1016/j.chempr.2018.07.010

552. Michael Kathan, Fabian Eisenreich, Christoph Jurissek, Andre Dallmann, Johannes Gurke, Stefan Hecht.
Light-driven molecular trap enables bidirectional manipulation of dynamic covalent systems. Nature Chemistry
2018, 10 (10) , 1031-1036. https://doi.org/10.1038/s41557-018-0106-8

553. Giulio Ragazzon, Leonard J. Prins. Energy consumption in chemical fuel-driven self-assembly. Nature
Nanotechnology 2018, 13 (10) , 882-889. https://doi.org/10.1038/s41565-018-0250-8

554. Andre C Barato, Raphael Chetrite, Alessandra Faggionato, Davide Gabrielli. Bounds on current
uctuations in periodically driven systems. New Journal of Physics 2018, 20 (10) , 103023.
https://doi.org/10.1088/1367-2630/aae512

555. Kazumitsu Onizuka, Takuya Miyashita, Tomoko Chikuni, Mamiko Ozawa, Hiroshi Abe, Fumi Nagatsugi.
Structural optimization of pseudorotaxane-forming oligonucleotides for e cient and stable complex
formation. Nucleic Acids Research 2018, 46 (17) , 8710-8719. https://doi.org/10.1093/nar/gky744

556. Giulio Ragazzon, Christian Schäfer, Paola Franchi, Serena Silvi, Benoit Colasson, Marco Lucarini, Alberto
Credi. Remote electrochemical modulation of pK a in a rotaxane by co-conformational allostery. Proceedings
of the National Academy of Sciences 2018, 115 (38) , 9385-9390. https://doi.org/10.1073/pnas.1712783115

557. Diederik Roke, Sander J. Wezenberg, Ben L. Feringa. Molecular rotary motors: Unidirectional motion
around double bonds. Proceedings of the National Academy of Sciences 2018, 115 (38) , 9423-9431.
https://doi.org/10.1073/pnas.1712784115

558. Liang Zhang, Vanesa Marcos, David A. Leigh. Molecular machines with bio-inspired mechanisms.
Proceedings of the National Academy of Sciences 2018, 115 (38) , 9397-9404.
https://doi.org/10.1073/pnas.1712788115
559. Damien Sluysmans, Floriane Devaux, Carson J. Bruns, J. Fraser Stoddart, Anne-Sophie Duwez. Dynamic
force spectroscopy of synthetic oligorotaxane foldamers. Proceedings of the National Academy of Sciences
2018, 115 (38) , 9362-9366. https://doi.org/10.1073/pnas.1712790115
  

560. R. Dean Astumian. Stochastically pumped adaptation and directional motion of molecular machines.
Proceedings of the National Academy of Sciences 2018, 115 (38) , 9405-9413.
https://doi.org/10.1073/pnas.1714498115

561. Jared D. Harris, Mark J. Moran, Ivan Aprahamian. New molecular switch architectures. Proceedings of
the National Academy of Sciences 2018, 115 (38) , 9414-9422. https://doi.org/10.1073/pnas.1714499115

562. Guillaume Erbland, Yohan Gisbert, Gwénaël Rapenne, Claire Kammerer. Expedient Synthesis of
Thioether-Functionalized Hydrotris(indazolyl)borate as an Anchoring Platform for Rotary Molecular Machines.
European Journal of Organic Chemistry 2018, 2018 (34) , 4731-4739.
https://doi.org/10.1002/ejoc.201800990

563. Benjamin Riss-Yaw, Caroline Clavel, Philippe Laurent, Philip Waelès, Frédéric Coutrot. The Importance of
Length and Flexibility of Macrocycle-Containing Molecular Translocators for the Synthesis of Improbable
[2]Rotaxanes. Chemistry - A European Journal 2018, 24 (51) , 13659-13666.
https://doi.org/10.1002/chem.201802831

564. Debajyoti Basak, Parichita Saha, Nandita Madhavan. Basic Design Elements for Tunable Cation
Transport Using Picolinic-Acid-Incorporated Tetrapeptides. ChemistrySelect 2018, 3 (33) , 9731-9735.
https://doi.org/10.1002/slct.201801612

565. Yujia Qing, Sandra A. Ionescu, Gökçe Su Pulcu, Hagan Bayley. Directional control of a processive
molecular hopper. Science 2018, 361 (6405) , 908-912. https://doi.org/10.1126/science.aat3872

566. Valeria Zanichelli, Margherita Bazzoni, Arturo Arduini, Paola Franchi, Marco Lucarini, Giulio Ragazzon,
Andrea Secchi, Serena Silvi. Redox-Switchable Calix[6]arene-Based Isomeric Rotaxanes. Chemistry - A
European Journal 2018, 24 (47) , 12370-12382. https://doi.org/10.1002/chem.201800496

567. Edgar Uhl, Stefan Thumser, Peter Mayer, Henry Dube. Übertragung unidirektionaler molekularer
Motorrotation auf eine räumlich getrennte Biarylachse. Angewandte Chemie 2018, 130 (34) , 11231-11235.
https://doi.org/10.1002/ange.201804716

568. Edgar Uhl, Stefan Thumser, Peter Mayer, Henry Dube. Transmission of Unidirectional Molecular Motor
Rotation to a Remote Biaryl Axis. Angewandte Chemie International Edition 2018, 57 (34) , 11064-11068.
https://doi.org/10.1002/anie.201804716

569. Diederik Roke, Constantin Stuckhardt, Wojciech Danowski, Sander J. Wezenberg, Ben L. Feringa. Light‐
Gated Rotation in a Molecular Motor Functionalized with a Dithienylethene Switch. Angewandte Chemie 2018,
130 (33) , 10675-10679. https://doi.org/10.1002/ange.201802392

570. Diederik Roke, Constantin Stuckhardt, Wojciech Danowski, Sander J. Wezenberg, Ben L. Feringa. Light‐
Gated Rotation in a Molecular Motor Functionalized with a Dithienylethene Switch. Angewandte Chemie
International Edition 2018, 57 (33) , 10515-10519. https://doi.org/10.1002/anie.201802392

571. Jiří Václavík, Iveta Klimánková, Alena Budinská, Petr Beier. Advances in the Synthesis and Application of
Tetra uoroethylene- and 1,1,2,2-Tetra uoroethyl-Containing Compounds. European Journal of Organic
Chemistry 2018, 2018 (27-28) , 3554-3593. https://doi.org/10.1002/ejoc.201701590
572. Jun Wang, Bo Durbeej. Toward Fast and E cient Visible-Light-Driven Molecular Motors: A Minimal
Design. ChemistryOpen 2018, 7 (8) , 583-589. https://doi.org/10.1002/open.201800089
  

573. Yong Chen, Feihe Huang, Zhan-Ting Li, Yu Liu. Controllable macrocyclic supramolecular assemblies in
aqueous solution. Science China Chemistry 2018, 61 (8) , 979-992. https://doi.org/10.1007/s11426-018-9337-
4

574. Andre C. Barato, Édgar Roldán, Ignacio A. Martínez, Simone Pigolotti. Arcsine Laws in Stochastic
Thermodynamics. Physical Review Letters 2018, 121 (9) https://doi.org/10.1103/PhysRevLett.121.090601

575. G. A. Dushenko, I. E. Mikhailov, O. I. Mikhailova, R. M. Minyaev, V. I. Minkin. Structure and


Rearrangements of 7-(Heptaphenylcycloheptatrienyl)isochalcogencyanates. Russian Journal of Organic
Chemistry 2018, 54 (8) , 1134-1147. https://doi.org/10.1134/S1070428018080043

576. Tim Raeker, Björn Jansen, Dominik Behrens, Bernd Hartke. Simulations of optically switchable molecular
machines for particle transport. Journal of Computational Chemistry 2018, 39 (20) , 1433-1443.
https://doi.org/10.1002/jcc.25212

577. Fatemah M. Alhassawi, Maria G. Corradini, Michael A. Rogers, Richard D. Ludescher. Potential
applications of luminescent molecular rotors in food science and engineering. Critical Reviews in Food
Science and Nutrition 2018, 58 (11) , 1902-1916. https://doi.org/10.1080/10408398.2017.1278583

578. Cristian Pezzato, Minh T. Nguyen, Dong Jun Kim, Ommid Anamimoghadam, Lorenzo Mosca, J. Fraser
Stoddart. Controlling Dual Molecular Pumps Electrochemically. Angewandte Chemie 2018, 130 (30) , 9469-
9473. https://doi.org/10.1002/ange.201803848

579. Cristian Pezzato, Minh T. Nguyen, Dong Jun Kim, Ommid Anamimoghadam, Lorenzo Mosca, J. Fraser
Stoddart. Controlling Dual Molecular Pumps Electrochemically. Angewandte Chemie International Edition
2018, 57 (30) , 9325-9329. https://doi.org/10.1002/anie.201803848

580. Chiara Biagini, Flaminia Di Pietri, Luigi Mandolini, Osvaldo Lanzalunga, Stefano Di Stefano. Photoinduced
Release of a Chemical Fuel for Acid-Base-Operated Molecular Machines. Chemistry - A European Journal
2018, 24 (40) , 10122-10127. https://doi.org/10.1002/chem.201800474

581. Jung-Ho Hong, Adil S. Aslam, Min-Sung Ko, Jonghoon Choi, Yunho Lee, Dong-Gyu Cho. Bond Rotation in
an Aromatic Carbaporphyrin: Allyliporphyrin. Chemistry - A European Journal 2018, 24 (40) , 10054-10058.
https://doi.org/10.1002/chem.201802176

582. . Introduction. 2018,,, 1-29. https://doi.org/10.1002/9781119126126.ch1

583. . Molecular Gearing Systems. 2018,,, 97-140. https://doi.org/10.1002/9781119126126.ch3

584. Michael M. Lerch, Mariangela Di Donato, Adèle D. Laurent, Miroslav Medved', Alessandro Iagatti, Laura
Bussotti, Andrea Lapini, Wybren Jan Buma, Paolo Foggi, Wiktor Szymański, Ben L. Feringa. Solvent Effects on
the Actinic Step of Donor–Acceptor Stenhouse Adduct Photoswitching. Angewandte Chemie 2018, 130 (27) ,
8195-8200. https://doi.org/10.1002/ange.201803058

585. Michael M. Lerch, Mariangela Di Donato, Adèle D. Laurent, Miroslav Medved', Alessandro Iagatti, Laura
Bussotti, Andrea Lapini, Wybren Jan Buma, Paolo Foggi, Wiktor Szymański, Ben L. Feringa. Solvent Effects on
the Actinic Step of Donor–Acceptor Stenhouse Adduct Photoswitching. Angewandte Chemie International
Edition 2018, 57 (27) , 8063-8068. https://doi.org/10.1002/anie.201803058
  
586. Steven M. Bachrach. Tetrasubstituted bisadamantylidenes-Highly twisted alkenes. Journal of Physical
Organic Chemistry 2018, 31 (7) , e3840. https://doi.org/10.1002/poc.3840

587. Mohammad Zarei, Mohanna Zarei. Self-Propelled Micro/Nanomotors for Sensing and Environmental
Remediation. Small 2018, 14 (30) , 1800912. https://doi.org/10.1002/smll.201800912

588. N. Kh. Petrov, D. A. Ivanov, I. V. Kryukov, A. D. Svirida, Yu. A. Shandarov, M. V. Al mov. The Time-Resolved
Fluorescence Stokes Shift of Cucurbit[6]Uril Complexes with a Pyridinium Styryl Dye. Journal of Fluorescence
2018, 28 (4) , 883-887. https://doi.org/10.1007/s10895-018-2256-x

589. I. B. Sivaev. Design of molecular switches based on transition metal bis(dicarbollide) complexes.
Russian Chemical Bulletin 2018, 67 (7) , 1117-1130. https://doi.org/10.1007/s11172-018-2193-5

590. Jie-Shun Cui, Qian-Kai Ba, Hua Ke, Arto Valkonen, Kari Rissanen, Wei Jiang. Directional Shuttling of a
Stimuli-Responsive Cone-Like Macrocycle on a Single-State Symmetric Dumbbell Axle. Angewandte Chemie
2018, 130 (26) , 7935-7940. https://doi.org/10.1002/ange.201803349

591. Jie-Shun Cui, Qian-Kai Ba, Hua Ke, Arto Valkonen, Kari Rissanen, Wei Jiang. Directional Shuttling of a
Stimuli-Responsive Cone-Like Macrocycle on a Single-State Symmetric Dumbbell Axle. Angewandte Chemie
International Edition 2018, 57 (26) , 7809-7814. https://doi.org/10.1002/anie.201803349

592. Sergey A. Anufriev, Svetlana A. Erokhina, Kyrill Yu. Suponitsky, Aleksei A. Anisimov, Julia N. Laskova, Ivan
A. Godovikov, Fabrizia Fabrizi de Biani, Maddalena Corsini, Igor B. Sivaev, Vladimir I. Bregadze. Synthesis and
structure of bis(methylsulfanyl) derivatives of iron bis(dicarbollide). Journal of Organometallic Chemistry 2018,
865 , 239-246. https://doi.org/10.1016/j.jorganchem.2018.04.019

593. Boris Atenas, Sergio Curilef. Dynamics of the d-HMF model: Sensitive dependence on size and initial
conditions. Journal of Physics: Conference Series 2018, 1043 , 012009. https://doi.org/10.1088/1742-
6596/1043/1/012009

594. Sandra P. Amaral, Maun H. Tawara, Marcos Fernandez-Villamarin, Erea Borrajo, José Martínez-Costas,
Anxo Vidal, Ricardo Riguera, Eduardo Fernandez-Megia. Tuning the Size of Nanoassembles: A Hierarchical
Transfer of Information from Dendrimers to Polyion Complexes. Angewandte Chemie 2018, 130 (19) , 5371-
5375. https://doi.org/10.1002/ange.201712244

595. Mathieu Denis, Jessica Pancholi, Kajally Jobe, Michael Watkinson, Stephen M. Goldup. Chelating
Rotaxane Ligands as Fluorescent Sensors for Metal Ions. Angewandte Chemie 2018, 130 (19) , 5408-5412.
https://doi.org/10.1002/ange.201712931

596. Mathieu Denis, Lei Qin, Peter Turner, Katrina A. Jolliffe, Stephen M. Goldup. A Fluorescent Ditopic
Rotaxane Ion-Pair Host. Angewandte Chemie 2018, 130 (19) , 5413-5417.
https://doi.org/10.1002/ange.201713105

597. Sandra P. Amaral, Maun H. Tawara, Marcos Fernandez-Villamarin, Erea Borrajo, José Martínez-Costas,
Anxo Vidal, Ricardo Riguera, Eduardo Fernandez-Megia. Tuning the Size of Nanoassembles: A Hierarchical
Transfer of Information from Dendrimers to Polyion Complexes. Angewandte Chemie International Edition
2018, 57 (19) , 5273-5277. https://doi.org/10.1002/anie.201712244
598. Mathieu Denis, Jessica Pancholi, Kajally Jobe, Michael Watkinson, Stephen M. Goldup. Chelating
Rotaxane Ligands as Fluorescent Sensors for Metal Ions. Angewandte Chemie International Edition 2018, 57
(19) , 5310-5314. https://doi.org/10.1002/anie.201712931
  

599. Mathieu Denis, Lei Qin, Peter Turner, Katrina A. Jolliffe, Stephen M. Goldup. A Fluorescent Ditopic
Rotaxane Ion-Pair Host. Angewandte Chemie International Edition 2018, 57 (19) , 5315-5319.
https://doi.org/10.1002/anie.201713105

600. Yoshiaki Amatatsu. Computational design of a light-driven imine-based motor with bulky chiral stator.
Chemical Physics 2018, 508 , 51-60. https://doi.org/10.1016/j.chemphys.2018.03.011

601. Steve Goldup. Molecular machines swap rings. Nature 2018, 557 (7703) , 39-40.
https://doi.org/10.1038/d41586-018-02732-5

602. Guillaume De Bo, Malcolm A. Y. Gall, Sonja Kuschel, Julien De Winter, Pascal Gerbaux, David A. Leigh. An
arti cial molecular machine that builds an asymmetric catalyst. Nature Nanotechnology 2018, 13 (5) , 381-
385. https://doi.org/10.1038/s41565-018-0105-3

603. Andre C Barato, Raphael Chetrite. Current uctuations in periodically driven systems. Journal of
Statistical Mechanics: Theory and Experiment 2018, 2018 (5) , 053207. https://doi.org/10.1088/1742-
5468/aabfc5

604. A. D. Svirida, D. A. Ivanov, Yu. A. Shandarov, I. V. Kryukov, N. Kh. Petrov, M. V. Al mov. Effect of Heavy
Water on Ultrafast Dynamics of the Fluorescence Stokes Shift for a Styryl Dye and its Complexes with
Cucurbiturils. High Energy Chemistry 2018, 52 (3) , 269-271. https://doi.org/10.1134/S0018143918030153

605. Jan Gieseler, James Millen. Levitated Nanoparticles for Microscopic Thermodynamics—A Review.
Entropy 2018, 20 (5) , 326. https://doi.org/10.3390/e20050326

606. Valeria Zanichelli, Luca Dallacasagrande, Arturo Arduini, Andrea Secchi, Giulio Ragazzon, Serena Silvi,
Alberto Credi. Electrochemically Triggered Co-Conformational Switching in a [2]catenane Comprising a Non-
Symmetric Calix[6]arene Wheel and a Two-Station Oriented Macrocycle. Molecules 2018, 23 (5) , 1156.
https://doi.org/10.3390/molecules23051156

607. Si-Jia Rao, Qi Zhang, Xu-Hao Ye, Chuan Gao, Da-Hui Qu. Integrative Self-Sorting: One-Pot Synthesis of a
Hetero[4]rotaxane from a Daisy-Chain-Containing Hetero[4]pseudorotaxane. Chemistry - An Asian Journal
2018, 13 (7) , 815-821. https://doi.org/10.1002/asia.201800011

608. G. A. Dushenko, I. E. Mikhailov, O. I. Mikhailova, R. M. Minyaev, V. I. Minkin. Reversible Migrations of Nitro


Group in a Methyltetramethoxycarbonylcyclopentadiene System. Doklady Chemistry 2018, 479 (2) , 53-57.
https://doi.org/10.1134/S0012500818040067

609. Nicholas H. Evans. Chiral Catenanes and Rotaxanes: Fundamentals and Emerging Applications.
Chemistry - A European Journal 2018, 24 (13) , 3101-3112. https://doi.org/10.1002/chem.201704149

610. Shun Yang, Zhoulin Luan, Chuan Gao, Jingjing Yu, Dahui Qu. Triggering a [2]rotaxane molecular shuttle
through hydrogen sul de. Science China Chemistry 2018, 61 (3) , 306-310. https://doi.org/10.1007/s11426-
017-9104-x
611. Damien Sluysmans, Sandrine Hubert, Carson J. Bruns, Zhixue Zhu, J. Fraser Stoddart, Anne-Sophie
Duwez. Synthetic oligorotaxanes exert high forces when folding under mechanical load. Nature
  
Nanotechnology 2018, 13 (3) , 209-213. https://doi.org/10.1038/s41565-017-0033-7

612. Lucy van Dijk, Michael J. Tilby, Robert Szpera, Owen A. Smith, Holly A. P. Bunce, Stephen P. Fletcher.
Molecular machines for catalysis. Nature Reviews Chemistry 2018, 2 (3) https://doi.org/10.1038/s41570-018-
0117

613. Saeed Amirjalayer, Alberto Martinez-Cuezva, Jose Berna, Sander Woutersen, Wybren Jan Buma.
Photoinduced Pedalo-Type Motion in an Azodicarboxamide-Based Molecular Switch. Angewandte Chemie
2018, 130 (7) , 1810-1814. https://doi.org/10.1002/ange.201709666

614. Saeed Amirjalayer, Alberto Martinez-Cuezva, Jose Berna, Sander Woutersen, Wybren Jan Buma.
Photoinduced Pedalo-Type Motion in an Azodicarboxamide-Based Molecular Switch. Angewandte Chemie
International Edition 2018, 57 (7) , 1792-1796. https://doi.org/10.1002/anie.201709666

615. Xiaochuan Li, Yujie Han, Myeong Jin Kim, Young-A Son. Reversed photochromism reactivity of
malononitrile attached bisthienylthene. Molecular Crystals and Liquid Crystals 2018, 662 (1) , 147-156.
https://doi.org/10.1080/15421406.2018.1466534

616. Jian Sun, Ruochen Lan, Yanzi Gao, Meng Wang, Wanshu Zhang, Ling Wang, Lanying Zhang, Zhou Yang,
Huai Yang. Stimuli-Directed Dynamic Recon guration in Self-Organized Helical Superstructures Enabled by
Chemical Kinetics of Chiral Molecular Motors. Advanced Science 2018, 5 (2) , 1700613.
https://doi.org/10.1002/advs.201700613

617. Yu Heng Lau, Jack K. Clegg, Jason R. Price, Rene B. Macquart, Matthew H. Todd, Peter J. Rutledge.
Molecular Switches for any pH: A Systematic Study of the Versatile Coordination Behaviour of Cyclam
Scorpionands. Chemistry - A European Journal 2018, 24 (7) , 1573-1585.
https://doi.org/10.1002/chem.201703488

618. Massimo Baroncini, Lorenzo Casimiro, Christiaan de Vet, Jessica Groppi, Serena Silvi, Alberto Credi.
Making and Operating Molecular Machines: A Multidisciplinary Challenge. ChemistryOpen 2018, 7 (2) , 169-
179. https://doi.org/10.1002/open.201700181

619. Xiaoyan He, Corinne Lagrost, Lucie Norel, Stéphane Rigaut. Ruthenium(II) σ-arylacetylide complexes as
redox active units for (multi-)functional molecular devices. Polyhedron 2018, 140 , 169-180.
https://doi.org/10.1016/j.poly.2017.08.025

620. Joyanta Choudhury. Recent developments on arti cial switchable catalysis. Tetrahedron Letters 2018,
59 (6) , 487-495. https://doi.org/10.1016/j.tetlet.2017.12.070

621. Cailing Ni, Daijun Zha, Hebo Ye, Yu Hai, Yuntao Zhou, Eric V. Anslyn, Lei You. Dynamic Covalent
Chemistry within Biphenyl Scaffolds: Reversible Covalent Bonding, Control of Selectivity, and Chirality Sensing
with a Single System. Angewandte Chemie 2018, 130 (5) , 1314-1319.
https://doi.org/10.1002/ange.201711602

622. Cailing Ni, Daijun Zha, Hebo Ye, Yu Hai, Yuntao Zhou, Eric V. Anslyn, Lei You. Dynamic Covalent
Chemistry within Biphenyl Scaffolds: Reversible Covalent Bonding, Control of Selectivity, and Chirality Sensing
with a Single System. Angewandte Chemie International Edition 2018, 57 (5) , 1300-1305.
https://doi.org/10.1002/anie.201711602
623. Jovana Milić, Michal Zalibera, Darius Talaat, Julia Nomrowski, Nils Trapp, Laurent Ruhlmann, Corinne
Boudon, Oliver S. Wenger, Anton Savitsky, Wolfgang Lubitz, François Diederich. Photoredox-Switchable
  
Resorcin[4]arene Cavitands: Radical Control of Molecular Gripping Machinery via Hydrogen Bonding.
Chemistry - A European Journal 2018, 24 (6) , 1431-1440. https://doi.org/10.1002/chem.201704788

624. Volker Adam, Deepak K. Prusty, Mathias Centola, Marko Škugor, Jeffrey S. Hannam, Julián Valero,
Bernhard Klöckner, Michael Famulok. Expanding the Toolbox of Photoswitches for DNA Nanotechnology Using
Arylazopyrazoles. Chemistry - A European Journal 2018, 24 (5) , 1062-1066.
https://doi.org/10.1002/chem.201705500

625. Baswanth Oruganti, Jun Wang, Bo Durbeej. Quantum chemical design of rotary molecular motors.
International Journal of Quantum Chemistry 2018, 118 (1) , e25405. https://doi.org/10.1002/qua.25405

626. Indrajit Paul, Abir Goswami, Nikita Mittal, Michael Schmittel. Katalytische Drei-Komponenten-Maschinen:
Steuerung der katalytischen Aktivität mittels Maschinengeschwindigkeit. Angewandte Chemie 2018, 130 (1) ,
360-364. https://doi.org/10.1002/ange.201709644

627. Indrajit Paul, Abir Goswami, Nikita Mittal, Michael Schmittel. Catalytic Three-Component Machinery:
Control of Catalytic Activity by Machine Speed. Angewandte Chemie International Edition 2018, 57 (1) , 354-
358. https://doi.org/10.1002/anie.201709644

628. Iwona Nierengarten, Eric Meichsner, Michel Holler, Pauline Pieper, Robert Deschenaux, Béatrice
Delavaux-Nicot, Jean-François Nierengarten. Preparation of Pillar[5]arene-Based [2]Rotaxanes by a Stopper-
Exchange Strategy. Chemistry - A European Journal 2018, 24 (1) , 169-177.
https://doi.org/10.1002/chem.201703997

629. Béatrice Delavaux-Nicot, Haifa Ben Aziza, Iwona Nierengarten, Thi Minh Nguyet Trinh, Eric Meichsner,
Matthieu Chessé, Michel Holler, Rym Abidi, Emmanuel Maisonhaute, Jean-François Nierengarten. A Rotaxane
Scaffold for the Construction of Multiporphyrinic Light-Harvesting Devices. Chemistry - A European Journal
2018, 24 (1) , 133-140. https://doi.org/10.1002/chem.201704124

630. Thomas van Leeuwen, Wojciech Danowski, Stefano F. Pizzolato, Peter Štacko, Sander J. Wezenberg,
Ben L. Feringa. Braking of a Light-Driven Molecular Rotary Motor by Chemical Stimuli. Chemistry - A European
Journal 2018, 24 (1) , 81-84. https://doi.org/10.1002/chem.201704747

631. Gloria Bazargan, Karl Sohlberg. Advances in modelling switchable mechanically interlocked molecular
architectures. International Reviews in Physical Chemistry 2018, 37 (1) , 1-82.
https://doi.org/10.1080/0144235X.2018.1419042

632. Hao Xing, Zhengtao Li, Zi Liang Wu, Feihe Huang. Catenane Crosslinked Mechanically Adaptive Polymer
Gel. Macromolecular Rapid Communications 2018, 39 (1) , 1700361.
https://doi.org/10.1002/marc.201700361

633. Jian Zhang, Jing Xu, Yan Wang, Huaiguo Xue, Huan Pang. Our Contributions in Nanochemistry for
Antibiosis, Electrocatalyst and Energy Storage Materials. The Chemical Record 2018, 18 (1) , 91-104.
https://doi.org/10.1002/tcr.201700026

634. Mohammad Ashrafuzzaman. Molecular Machines of the Cell. 2018,,, 183-235.


https://doi.org/10.1007/978-3-319-77465-7_5
635. Alan Cooper. Thermodynamic Fluctuations. 2018,,, 1-4. https://doi.org/10.1007/978-3-642-35943-
9_10070-1
  
636. Lianjie Xu, Wen-Bin Zhang. Topology: a unique dimension in protein engineering. Science China
Chemistry 2018, 61 (1) , 3-16. https://doi.org/10.1007/s11426-017-9155-2

637. Michael Schmittel, Suchismita Saha. From Self-Sorting of Dynamic Metal–Ligand Motifs to
(Supra)Molecular Machinery in Action. 2018,,, 135-175. https://doi.org/10.1016/bs.adioch.2017.11.006

638. Jean-François Ayme, Jean-Marie Lehn. From Coordination Chemistry to Adaptive Chemistry. 2018,,, 3-
78. https://doi.org/10.1016/bs.adioch.2017.11.009

639. James A. Findlay, James D. Crowley. Functional nanomachines: Recent advances in synthetic molecular
machinery. Tetrahedron Letters 2018, 59 (4) , 334-346. https://doi.org/10.1016/j.tetlet.2017.12.036

640. Toshikazu Takata, Daisuke Aoki. Topology-transformable polymers: linear–branched polymer structural
transformation via the mechanical linking of polymer chains. Polymer Journal 2018, 50 (1) , 127-147.
https://doi.org/10.1038/pj.2017.60

641. Derrick A. Roberts, Ben S. Pilgrim, Jonathan R. Nitschke. Covalent post-assembly modi cation in
metallosupramolecular chemistry. Chemical Society Reviews 2018, 47 (2) , 626-644.
https://doi.org/10.1039/C6CS00907G

642. R. D. Astumian. Stochastic pumping of non-equilibrium steady-states: how molecules adapt to a


uctuating environment. Chemical Communications 2018, 54 (5) , 427-444.
https://doi.org/10.1039/C7CC06683J

643. Yu-Xuan Wang, Qi-Feng Zhou, Li-Jun Chen, Lin Xu, Cui-Hong Wang, Xiaopeng Li, Hai-Bo Yang. Facile
construction of organometallic rotaxane-terminated dendrimers using neutral platinum–acetylides as the
main scaffold. Chemical Communications 2018, 54 (18) , 2224-2227. https://doi.org/10.1039/C7CC08729B

644. Elizabeth Ellis, Suresh Moorthy, Weng-I Katherine Chio, Tung-Chun Lee. Arti cial molecular and
nanostructures for advanced nanomachinery. Chemical Communications 2018, 54 (33) , 4075-4090.
https://doi.org/10.1039/C7CC09133H

645. Sudhakar Gaikwad, Susnata Pramanik, Soumen De, Michael Schmittel. A high-speed network of
nanoswitches for on/off control of catalysis. Dalton Transactions 2018, 47 (6) , 1786-1790.
https://doi.org/10.1039/C7DT04695B

646. Jinjia Xu, Atsuro Takai, Alisa Bannaron, Takafumi Nakagawa, Yutaka Matsuo, Manabu Sugimoto,
Yoshitaka Matsushita, Masayuki Takeuchi. A helically-twisted ladder based on 9,9′-bi uorenylidene: synthesis,
characterization, and carrier-transport properties. Materials Chemistry Frontiers 2018, 2 (4) , 780-784.
https://doi.org/10.1039/C7QM00583K

647. T. Akutagawa. Dynamic molecular assemblies toward a new frontier in materials chemistry. Materials
Chemistry Frontiers 2018, 2 (6) , 1064-1073. https://doi.org/10.1039/C7QM00603A

648. Gulshan Kumar, Richa Goel, Kamaldeep Paul, Vijay Luxami. Investigation of rotameric conformations of
substituted imidazo-[1,2- a ]pyrazine: experimental and theoretical approaches. RSC Advances 2018, 8 (18) ,
9707-9717. https://doi.org/10.1039/C7RA13617J
649. Chiara Biagini, Simone Albano, Rachele Caruso, Luigi Mandolini, José Augusto Berrocal, Stefano Di
 
 fuelled
Stefano. Variations in the fuel structure control the rate of the back and forth motions of a chemically
molecular switch. Chemical Science 2018, 9 (1) , 181-188. https://doi.org/10.1039/C7SC04123C

650. Guillaume De Bo. Mechanochemistry of the mechanical bond. Chemical Science 2018, 9 (1) , 15-21.
https://doi.org/10.1039/C7SC04200K

651. Raiko Hahn, Fabian Bohle, Stefan Kotte, Tristan J. Keller, Stefan-S. Jester, Andreas Hansen, Stefan
Grimme, Birgit Esser. Donor–acceptor interactions between cyclic trinuclear pyridinate gold( i )-complexes and
electron-poor guests: nature and energetics of guest-binding and templating on graphite. Chemical Science
2018, 9 (14) , 3477-3483. https://doi.org/10.1039/C7SC05355J

652. Suehyun Park, Heesun Joo, Jun Soo Kim. Directional rolling of positively charged nanoparticles along a
exibility gradient on long DNA molecules. Soft Matter 2018, 14 (5) , 817-825.
https://doi.org/10.1039/C7SM02016C

653. A. Zubillaga, P. Ferreira, A. J. Parola, S. Gago, N. Basílio. pH-Gated photoresponsive shuttling in a water-
soluble pseudorotaxane. Chemical Communications 2018, 54 (22) , 2743-2746.
https://doi.org/10.1039/C8CC00688A

654. Abir Goswami, Susnata Pramanik, Michael Schmittel. Catalytically active nanorotor reversibly self-
assembled by chemical signaling within an eight-component network. Chemical Communications 2018, 54
(32) , 3955-3958. https://doi.org/10.1039/C8CC01496E

655. Qiang Shi, Zheng Meng, Jun-Feng Xiang, Chuan-Feng Chen. E cient control of movement in non-
photoresponsive molecular machines by a photo-induced proton-transfer strategy. Chemical Communications
2018, 54 (28) , 3536-3539. https://doi.org/10.1039/C8CC01570H

656. Hae-Geun Jeon, Hyun Kyung Lee, Seungwon Lee, Kyu-Sung Jeong. Foldamer-based helicate displaying
reversible switching between two distinct conformers. Chemical Communications 2018, 54 (45) , 5740-5743.
https://doi.org/10.1039/C8CC02758G

657. Yuan Wang, Yancong Tian, Yu-Zhe Chen, Li-Ya Niu, Li-Zhu Wu, Chen-Ho Tung, Qing-Zheng Yang, Roman
Boulatov. A light-driven molecular machine based on stiff stilbene. Chemical Communications 2018, 54 (57) ,
7991-7994. https://doi.org/10.1039/C8CC04542A

658. Sourav Chakraborty, George R. Newkome. Terpyridine-based metallosupramolecular constructs: tailored


monomers to precise 2D-motifs and 3D-metallocages. Chemical Society Reviews 2018, 47 (11) , 3991-4016.
https://doi.org/10.1039/C8CS00030A

659. E. M. G. Jamieson, F. Modicom, S. M. Goldup. Chirality in rotaxanes and catenanes. Chemical Society
Reviews 2018, 47 (14) , 5266-5311. https://doi.org/10.1039/C8CS00097B

660. Hendrik V. Schröder, Felix Witte, Marius Gaedke, Sebastian Sobottka, Lisa Suntrup, Henrik Hupatz, Arto
Valkonen, Beate Paulus, Kari Rissanen, Biprajit Sarkar, Christoph A. Schalley. An aryl-fused redox-active
tetrathiafulvalene with enhanced mixed-valence and radical-cation dimer stabilities. Organic & Biomolecular
Chemistry 2018, 16 (15) , 2741-2747. https://doi.org/10.1039/C8OB00415C

661. Adrian Saura-Sanmartin, Alberto Martinez-Cuezva, Aurelia Pastor, Delia Bautista, Jose Berna. Light-
driven exchange between extended and contracted lasso-like isomers of a bistable [1]rotaxane. Organic &
Biomolecular Chemistry 2018, 16 (38) , 6980-6987. https://doi.org/10.1039/C8OB02234H

  
662. Dechun Huang, Qiao Zhang, Yan Deng, Zheng Luo, Bo Li, Xin Shen, Zhenhui Qi, Shengyi Dong, Yan Ge,
Wei Chen. Polymeric crown ethers: LCST behavior in water and stimuli-responsiveness. Polymer Chemistry
2018, 9 (19) , 2574-2579. https://doi.org/10.1039/C8PY00412A

663. Teng Gong, Peng Li, Qi Sui, Jinquan Chen, Jianhua Xu, En-Qing Gao. A stable electron-de cient metal–
organic framework for colorimetric and luminescence sensing of phenols and anilines. Journal of Materials
Chemistry A 2018, 6 (19) , 9236-9244. https://doi.org/10.1039/C8TA02794C

664. Ihor M. Tkachenko, Yaroslav L. Kobzar, Volodymyr F. Korolovych, Alexandr V. Stryutsky, Liubov K.
Matkovska, Valery V. Shevchenko, Vladimir V. Tsukruk. Novel branched nanostructures based on polyhedral
oligomeric silsesquioxanes and azobenzene dyes containing different spacers and isolation groups. Journal of
Materials Chemistry C 2018, 6 (15) , 4065-4076. https://doi.org/10.1039/C8TC00223A

665. Yoshimitsu Sagara, Atsushi Seki, Yuna Kim, Nobuyuki Tamaoki. Linearly polarized photoluminescence
from an asymmetric cyclophane showing thermo- and mechanoresponsive luminescence. Journal of
Materials Chemistry C 2018, 6 (31) , 8453-8459. https://doi.org/10.1039/C8TC02919A

666. Ying Han, Li-Ming Xu, Cui-Yun Nie, Shuo Jiang, Jing Sun, Chao-Guo Yan. Synthesis of diamido-bridged
bis-pillar[5]arenes and tris-pillar[5]arenes for construction of unique [1]rotaxanes and bis-[1]rotaxanes. Beilstein
Journal of Organic Chemistry 2018, 14 , 1660-1667. https://doi.org/10.3762/bjoc.14.142

667. Antony Wing Hung Ng, Chi-Chung Yee, Kai Wang, Ho Yu Au-Yeung. E cient catenane synthesis by
cucurbit[6]uril-mediated azide–alkyne cycloaddition. Beilstein Journal of Organic Chemistry 2018, 14 , 1846-
1853. https://doi.org/10.3762/bjoc.14.158

668. Hendrik V Schröder, Christoph A Schalley. Tetrathiafulvalene – a redox-switchable building block to


control motion in mechanically interlocked molecules. Beilstein Journal of Organic Chemistry 2018, 14 , 2163-
2185. https://doi.org/10.3762/bjoc.14.190

669. Г.А. Душенко, И.Е. Михайлов, О.И. Михайлова, Р.М. Миняев, В.И. Минкин. СТРУКТУРА И
ПЕРЕГРУППИРОВКИ 7-(ГЕПТАФЕНИЛЦИКЛОГЕПТАТРИЕНИЛ)ИЗОХАЛЬКОГЕНЦИАНАТОВ, "Журнал
органической химии". Журнал органической химии 2018, (8) , 1127-1139.
https://doi.org/10.7868/S0514749218080046

670. Yenan Shen, Qianwen Zhou, Lei Zou, Qiaochun Wang. Synthesis of a Square [5]Catenane by Simple
Amine-Aldehyde Condensation. ChemistrySelect 2017, 2 (36) , 11977-11980.
https://doi.org/10.1002/slct.201702415

671. Qiong Wu, Phillip M. Rauscher, Xiaolong Lang, Rudy J. Wojtecki, Juan J. de Pablo, Michael J. A. Hore,
Stuart J. Rowan. Poly[ n ]catenanes: Synthesis of molecular interlocked chains. Science 2017, 358 (6369) ,
1434-1439. https://doi.org/10.1126/science.aap7675

672. Marta Estrader, Jorge Salinas Uber, Leoní A. Barrios, Jordi Garcia, Paul Lloyd-Williams, Olivier Roubeau,
Simon J. Teat, Guillem Aromí. A Magneto-optical Molecular Device: Interplay of Spin Crossover, Luminescence,
Photomagnetism, and Photochromism. Angewandte Chemie 2017, 129 (49) , 15828-15833.
https://doi.org/10.1002/ange.201709136

673. Philipp J. Altmann, Alexander Pöthig. Ein pH-abhängiger, mechanisch verzahnter Schalter:
organometallisches [2]Rotaxan und organisches [3]Rotaxan. Angewandte Chemie 2017, 129 (49) , 15939-
15942. https://doi.org/10.1002/ange.201709921

  
674. Marta Estrader, Jorge Salinas Uber, Leoní A. Barrios, Jordi Garcia, Paul Lloyd-Williams, Olivier Roubeau,
Simon J. Teat, Guillem Aromí. A Magneto-optical Molecular Device: Interplay of Spin Crossover, Luminescence,
Photomagnetism, and Photochromism. Angewandte Chemie International Edition 2017, 56 (49) , 15622-
15627. https://doi.org/10.1002/anie.201709136

675. Philipp J. Altmann, Alexander Pöthig. A pH-Dependent, Mechanically Interlocked Switch: Organometallic
[2]Rotaxane vs. Organic [3]Rotaxane. Angewandte Chemie International Edition 2017, 56 (49) , 15733-15736.
https://doi.org/10.1002/anie.201709921

676. Michael Å. Petersen, Brian Rasmussen, Nicolaj N. Andersen, Stephan P. A. Sauer, Mogens Brøndsted
Nielsen, Sophie R. Beeren, Michael Pittelkow. Molecular Switching in Con ned Spaces: Effects of
Encapsulating the DHA/VHF Photo-Switch in Cucurbiturils. Chemistry - A European Journal 2017, 23 (67) ,
17010-17016. https://doi.org/10.1002/chem.201703196

677. Matthijs R. Panman, Chris N. van Dijk, Adriana Huerta-Viga, Hans J. Sanders, Bert H. Bakker, David A.
Leigh, Albert M. Brouwer, Wybren Jan Buma, Sander Woutersen. Transient two-dimensional vibrational
spectroscopy of an operating molecular machine. Nature Communications 2017, 8 (1)
https://doi.org/10.1038/s41467-017-02278-6

678. Martin Šiler, Petr Jákl, Oto Brzobohatý, Artem Ryabov, Radim Filip, Pavel Zemánek. Thermally induced
micro-motion by in ection in optical potential. Scienti c Reports 2017, 7 (1) https://doi.org/10.1038/s41598-
017-01848-4

679. Jamel Ali, U Kei Cheang, James D. Martindale, Mehdi Jabbarzadeh, Henry C. Fu, Min Jun Kim. Bacteria-
inspired nanorobots with agellar polymorphic transformations and bundling. Scienti c Reports 2017, 7 (1)
https://doi.org/10.1038/s41598-017-14457-y

680. Kai-Jen Chen, Pei-Lin Chen, Masaki Horie. Dynamic Pseudorotaxane Crystals Containing Metallocene
Complexes. Scienti c Reports 2017, 7 (1) https://doi.org/10.1038/s41598-017-14505-7

681. Liying Wang, Zhenyu Meng, Felicia Martina, Huilin Shao, Fangwei Shao. Fabrication of circular
assemblies with DNA tetrahedrons: from static structures to a dynamic rotary motor. Nucleic Acids Research
2017, 45 (21) , 12090-12099. https://doi.org/10.1093/nar/gkx1045

682. Igor Sivaev. Ferrocene and Transition Metal Bis(Dicarbollides) as Platform for Design of Rotatory
Molecular Switches. Molecules 2017, 22 (12) , 2201. https://doi.org/10.3390/molecules22122201

683. Ludwig Alexander Huber, Kerstin Hoffmann, Stefan Thumser, Niklas Böcher, Peter Mayer, Henry Dube.
Direct Observation of Hemithioindigo-Motor Unidirectionality. Angewandte Chemie 2017, 129 (46) , 14728-
14731. https://doi.org/10.1002/ange.201708178

684. Ludwig Alexander Huber, Kerstin Hoffmann, Stefan Thumser, Niklas Böcher, Peter Mayer, Henry Dube.
Direct Observation of Hemithioindigo-Motor Unidirectionality. Angewandte Chemie International Edition 2017,
56 (46) , 14536-14539. https://doi.org/10.1002/anie.201708178

685. Yasuyuki Yamada, Ryohei Itoh, Sayaka Ogino, Tatsuhisa Kato, Kentaro Tanaka. Dynamic Molecular
Invasion into a Multiply Interlocked Catenane. Angewandte Chemie 2017, 129 (45) , 14312-14317.
https://doi.org/10.1002/ange.201708248
686. Koki Mitsuno, Takafumi Yoshino, Iti Gupta, Shigeki Mori, Satoru Karasawa, Masatoshi Ishida, Hiroyuki
Furuta. Doubly N-Confused [36]Octaphyrin(1.1.1.1.1.1.1.1): Isomerization, Bis-Metal Coordination, and
  
Topological Chirality. Angewandte Chemie 2017, 129 (45) , 14440-14444.
https://doi.org/10.1002/ange.201708253

687. Yasuyuki Yamada, Ryohei Itoh, Sayaka Ogino, Tatsuhisa Kato, Kentaro Tanaka. Dynamic Molecular
Invasion into a Multiply Interlocked Catenane. Angewandte Chemie International Edition 2017, 56 (45) , 14124-
14129. https://doi.org/10.1002/anie.201708248

688. Koki Mitsuno, Takafumi Yoshino, Iti Gupta, Shigeki Mori, Satoru Karasawa, Masatoshi Ishida, Hiroyuki
Furuta. Doubly N-Confused [36]Octaphyrin(1.1.1.1.1.1.1.1): Isomerization, Bis-Metal Coordination, and
Topological Chirality. Angewandte Chemie International Edition 2017, 56 (45) , 14252-14256.
https://doi.org/10.1002/anie.201708253

689. Dario Mosca, Antoine Stopin, Johan Wouters, Nicola Demitri, Davide Bonifazi. Stereospeci c Winding of
Polycyclic Aromatic Hydrocarbons into Trinacria Propellers. Chemistry - A European Journal 2017, 23 (61) ,
15348-15354. https://doi.org/10.1002/chem.201702032

690. Vyacheslav V. Samoshin, Yu Zheng, Xin Liu. Trans -2-Aminocyclohexanol derivatives as pH-triggered
conformational switches. Journal of Physical Organic Chemistry 2017, 30 (11) , e3689.
https://doi.org/10.1002/poc.3689

691. Bartosz Tylkowski, Anna Trojanowska, Valentina Marturano, Martyna Nowak, Lukasz Marciniak, Marta
Giamberini, Veronica Ambrogi, Pierfrancesco Cerruti. Power of light – Functional complexes based on
azobenzene molecules. Coordination Chemistry Reviews 2017, 351 , 205-217.
https://doi.org/10.1016/j.ccr.2017.05.009

692. Aramballi J. Savyasachi, Oxana Kotova, Sankarasekaran Shanmugaraju, Samuel J. Bradberry, Gearóid M.
Ó’Máille, Thor nnur Gunnlaugsson. Supramolecular Chemistry: A Toolkit for Soft Functional Materials and
Organic Particles. Chem 2017, 3 (5) , 764-811. https://doi.org/10.1016/j.chempr.2017.10.006

693. Sergey A. Anufriev, Igor B. Sivaev, Kyrill Yu. Suponitsky, Vladimir I. Bregadze. Practical synthesis of 9-
methylthio-7,8-nido-carborane [9-MeS-7,8-C2B9H11]-. Some evidences of BH···X hydride-halogen bonds in 9-
XCH2(Me)S-7,8-C2B9H11 (X = Cl, Br, I). Journal of Organometallic Chemistry 2017, 849-850 , 315-323.
https://doi.org/10.1016/j.jorganchem.2017.03.025

694. Yoshiaki Yoshida, Takeshi Endo. Dependence of color change of vinylethylene carbonate copolymers
having N -substituted maleimides on chemical structure by acid-base switching in solution and solid state.
Reactive and Functional Polymers 2017, 120 , 139-146.
https://doi.org/10.1016/j.reactfunctpolym.2017.10.001

695. Zhiqiang Xu, Haojie Ge, Xie Han, Sheng Hua Liu, Xiang-Gao Meng, Jun Yin. Self-recognition behavior of
novel frameworks containing both urea and carboxylate anion motifs. Tetrahedron 2017, 73 (44) , 6386-6391.
https://doi.org/10.1016/j.tet.2017.09.036

696. Somrita Ray, Andre C. Barato. Stochastic thermodynamics of periodically driven systems: Fluctuation
theorem for currents and uni cation of two classes. Physical Review E 2017, 96 (5)
https://doi.org/10.1103/PhysRevE.96.052120

697. Sergey A. Anufriev, Svetlana A. Erokhina, Kyrill Yu. Suponitsky, Ivan A. Godovikov, Oleg A. Filippov,
Fabrizia Fabrizi de Biani, Maddalena Corsini, Alexander O. Chizhov, Igor B. Sivaev. Methylsulfanyl-Stabilized
Rotamers of Cobalt Bis(dicarbollide). European Journal of Inorganic Chemistry 2017, 2017 (38-39) , 4444-
4451. https://doi.org/10.1002/ejic.201700575
  
698. Mónica A. Gordillo, Mónica Soto-Monsalve, Christian C. Carmona-Vargas, Gustavo Gutiérrez, Richard F.
D'vries, Jean-Marie Lehn, Manuel N. Chaur. Photochemical and Electrochemical Triggered Bis(hydrazone)
Switch. Chemistry - A European Journal 2017, 23 (59) , 14872-14882.
https://doi.org/10.1002/chem.201703065

699. Sundus Erbas-Cakmak, Stephen D. P. Fielden, Ulvi Karaca, David A. Leigh, Charlie T. McTernan, Daniel J.
Tetlow, Miriam R. Wilson. Rotary and linear molecular motors driven by pulses of a chemical fuel. Science
2017, 358 (6361) , 340-343. https://doi.org/10.1126/science.aao1377

700. Tony D. James. Specialty Grand Challenges in Supramolecular Chemistry. Frontiers in Chemistry 2017, 5
https://doi.org/10.3389/fchem.2017.00083

701. Giorgio Baggi, Stephen J. Loeb. Rotationally Active Ligands: Dialing-Up Multiple Interlocked Co-
Conformations for Silver(I) Coordination. Chemistry - A European Journal 2017, 23 (57) , 14163-14166.
https://doi.org/10.1002/chem.201703485

702. Shinji Toyota, Kenta Kawahata, Kota Sugahara, Kan Wakamatsu, Tetsuo Iwanaga. Triple and Quadruple
Triptycene Gears in Rigid Macrocyclic Frameworks. European Journal of Organic Chemistry 2017, 2017 (37) ,
5696-5707. https://doi.org/10.1002/ejoc.201701067

703. Wei Zhao, Bo Qiao, Chun-Hsing Chen, Amar H. Flood. High-Fidelity Multistate Switching with Anion-
Anion and Acid-Anion Dimers of Organophosphates in Cyanostar Complexes. Angewandte Chemie 2017, 129
(42) , 13263-13267. https://doi.org/10.1002/ange.201707869

704. Wei Zhao, Bo Qiao, Chun-Hsing Chen, Amar H. Flood. High-Fidelity Multistate Switching with Anion-
Anion and Acid-Anion Dimers of Organophosphates in Cyanostar Complexes. Angewandte Chemie
International Edition 2017, 56 (42) , 13083-13087. https://doi.org/10.1002/anie.201707869

705. Qiang Shi, Ying Han, Chuan-Feng Chen. Complexation Between ( O -Methyl) 6 -2,6-Helic[6]arene and
Tertiary Ammonium Salts: Acid/Base- or Chloride-Ion-Responsive Host-Guest Systems and Synthesis of
[2]Rotaxane. Chemistry - An Asian Journal 2017, 12 (19) , 2576-2582.
https://doi.org/10.1002/asia.201700857

706. Daisuke Aoki, Toshikazu Takata. Mechanically linked supramolecular polymer architectures derived from
macromolecular [2]rotaxanes: Synthesis and topology transformation. Polymer 2017, 128 , 276-296.
https://doi.org/10.1016/j.polymer.2017.08.020

707. Stephen D. P. Fielden, David A. Leigh, Steffen L. Woltering. Molekulare Knoten. Angewandte Chemie
2017, 129 (37) , 11318-11347. https://doi.org/10.1002/ange.201702531

708. J. Fraser Stoddart. Mechanisch verzahnte Moleküle (MIMs) - molekulare Shuttle, Schalter und
Maschinen (Nobel-Aufsatz). Angewandte Chemie 2017, 129 (37) , 11244-11277.
https://doi.org/10.1002/ange.201703216

709. Stephen D. P. Fielden, David A. Leigh, Steffen L. Woltering. Molecular Knots. Angewandte Chemie
International Edition 2017, 56 (37) , 11166-11194. https://doi.org/10.1002/anie.201702531
710. J. Fraser Stoddart. Mechanically Interlocked Molecules (MIMs)-Molecular Shuttles, Switches, and
Machines (Nobel Lecture). Angewandte Chemie International Edition 2017, 56 (37) , 11094-11125.
  
https://doi.org/10.1002/anie.201703216

711. Kuppusamy Senthil Kumar, Mario Ruben. Emerging trends in spin crossover (SCO) based functional
materials and devices. Coordination Chemistry Reviews 2017, 346 , 176-205.
https://doi.org/10.1016/j.ccr.2017.03.024

712. Salma Kassem, Alan T. L. Lee, David A. Leigh, Vanesa Marcos, Leoni I. Palmer, Simone Pisano.
Stereodivergent synthesis with a programmable molecular machine. Nature 2017, 549 (7672) , 374-378.
https://doi.org/10.1038/nature23677

713. Paulraj Mosae Selvakumar. The Art of Converting Molecules into Machines: Supramolecular Chemistry.
MOJ Bioorganic & Organic Chemistry 2017, 1 (4) https://doi.org/10.15406/mojboc.2017.01.00021

714. Silvia Bracco, Fabio Castiglioni, Angiolina Comotti, Simona Galli, Mattia Negroni, Angelo Maspero, Piero
Sozzani. Ultrafast Molecular Rotors and Their CO 2 Tuning in MOFs with Rod-Like Ligands. Chemistry - A
European Journal 2017, 23 (47) , 11210-11215. https://doi.org/10.1002/chem.201702930

715. Ofer Kedem, Bryan Lau, Mark A. Ratner, Emily A. Weiss. Light-responsive organic ashing electron
ratchet. Proceedings of the National Academy of Sciences 2017, 114 (33) , 8698-8703.
https://doi.org/10.1073/pnas.1705973114

716. Susana Ibáñez, Macarena Poyatos, Eduardo Peris. Cation‐Driven Self‐Assembly of a Gold(I)‐Based
Metallo‐Tweezer. Angewandte Chemie 2017, 129 (33) , 9918-9922.
https://doi.org/10.1002/ange.201704359

717. Susana Ibáñez, Macarena Poyatos, Eduardo Peris. Cation‐Driven Self‐Assembly of a Gold(I)‐Based
Metallo‐Tweezer. Angewandte Chemie International Edition 2017, 56 (33) , 9786-9790.
https://doi.org/10.1002/anie.201704359

718. Wei Li, Tairong Kuang, Xiaoping Jiang, Jintao Yang, Ping Fan, Zhengping Zhao, Zhengdong Fei,
Mingqiang Zhong, Lingqian Chang, Feng Chen. Photoresponsive polyelectrolyte/mesoporous silica hybrid
materials with remote-controllable ionic transportation. Chemical Engineering Journal 2017, 322 , 445-453.
https://doi.org/10.1016/j.cej.2017.04.048

719. Cristian Pezzato, Minh T. Nguyen, Chuyang Cheng, Dong Jun Kim, Michael T. Otley, J. Fraser Stoddart.
An e cient arti cial molecular pump. Tetrahedron 2017, 73 (33) , 4849-4857.
https://doi.org/10.1016/j.tet.2017.05.087

720. Luke C. Delmas, Nicholas A. Payne, Franklyn Julien, Avril R. Williams. Modular construction of
pyrido[24]crown-8-based templates in the self-assembly of cross-linked [n]catenanes. Tetrahedron Letters
2017, 58 (33) , 3226-3229. https://doi.org/10.1016/j.tetlet.2017.07.006

721. Grzegorz Markiewicz, Anna Jenczak, Michał Kołodziejski, Julian J. Holstein, Jeremy K. M. Sanders, Artur
R Stefankiewicz. Selective C70 encapsulation by a robust octameric nanospheroid held together by 48
cooperative hydrogen bonds. Nature Communications 2017, 8 (1) https://doi.org/10.1038/ncomms15109

722. Mathieu Denis, Stephen M. Goldup. The active template approach to interlocked molecules. Nature
Reviews Chemistry 2017, 1 (8) https://doi.org/10.1038/s41570-017-0061
723. Andrea Mulas, Xiaoyan He, Yves-Marie Hervault, Lucie Norel, Stéphane Rigaut, Corinne Lagrost. Dual-
Responsive Molecular Switches Based on Dithienylethene-Ru II Organometallics in Self-Assembled
  
Monolayers Operating at Low Voltage. Chemistry - A European Journal 2017, 23 (42) , 10205-10214.
https://doi.org/10.1002/chem.201701903

724. Takahiro Fukino, Hiroshi Yamagishi, Takuzo Aida. Redox-Responsive Molecular Systems and Materials.
Advanced Materials 2017, 29 (25) , 1603888. https://doi.org/10.1002/adma.201603888

725. Wenxin Lu, Chiming Wang, Xinyao Li, Xiaopeng Zhan, Dongdong Qi, Yongzhong Bian. Binaphthol-
strapped chiral bis (porphyrinato) cerium double-decker complexes. Inorganic Chemistry Communications
2017, 81 , 18-21. https://doi.org/10.1016/j.inoche.2017.04.020

726. Ryan Baumgartner, Hailin Fu, Ziyuan Song, Yao Lin, Jianjun Cheng. Cooperative polymerization of α-
helices induced by macromolecular architecture. Nature Chemistry 2017, 9 (7) , 614-622.
https://doi.org/10.1038/nchem.2712

727. Andre C Barato, Udo Seifert. Thermodynamic cost of external control. New Journal of Physics 2017, 19
(7) , 073021. https://doi.org/10.1088/1367-2630/aa77d0

728. Adrian Saura-Sanmartin, Juan Martinez-Espin, Alberto Martinez-Cuezva, Mateo Alajarin, Jose Berna.
Effects on Rotational Dynamics of Azo and Hydrazodicarboxamide-Based Rotaxanes. Molecules 2017, 22 (7) ,
1078. https://doi.org/10.3390/molecules22071078

729. Masaki Kimura, Takuma Mizuno, Masahiro Ueda, Shinobu Miyagawa, Tsuneomi Kawasaki, Yuji
Tokunaga. Four-State Molecular Shuttling of [2]Rotaxanes in Response to Acid/Base and Alkali-Metal Cation
Stimuli. Chemistry - An Asian Journal 2017, 12 (12) , 1381-1390. https://doi.org/10.1002/asia.201700493

730. Loïc Le Pleux, Elisabeth Kapatsina, Julia Hildesheim, Daniel Häussinger, Marcel Mayor. A Molecular
Turnstile as an E -Field-Triggered Single-Molecule Switch: Concept and Synthesis. European Journal of Organic
Chemistry 2017, 2017 (22) , 3165-3178. https://doi.org/10.1002/ejoc.201700318

731. Taizo Mori, Daisuke Ishikawa, Yusuke Yonamine, Yoshihisa Fujii, Jonathan P. Hill, Izumi Ichinose,
Katsuhiko Ariga, Waka Nakanishi. Mechanically Induced Opening-Closing Action of Binaphthyl Molecular
Pliers: Digital Phase Transition versus Continuous Conformational Change. ChemPhysChem 2017, 18 (11) ,
1470-1474. https://doi.org/10.1002/cphc.201601144

732. Peter Štacko, Jos C. M. Kistemaker, Thomas van Leeuwen, Mu-Chieh Chang, Edwin Otten, Ben L.
Feringa. Locked synchronous rotor motion in a molecular motor. Science 2017, 356 (6341) , 964-968.
https://doi.org/10.1126/science.aam8808

733. Ke Min Chan, Dominik K. Kölmel, Shenliang Wang, Eric T. Kool. Color-Change Photoswitching of an
Alkynylpyrene Excimer Dye. Angewandte Chemie 2017, 129 (23) , 6597-6601.
https://doi.org/10.1002/ange.201701235

734. Ke Min Chan, Dominik K. Kölmel, Shenliang Wang, Eric T. Kool. Color-Change Photoswitching of an
Alkynylpyrene Excimer Dye. Angewandte Chemie International Edition 2017, 56 (23) , 6497-6501.
https://doi.org/10.1002/anie.201701235

735. Justin T. Foy, Quan Li, Antoine Goujon, Jean-Rémy Colard-Itté, Gad Fuks, Emilie Moulin, Olivier
Schiffmann, Damien Dattler, Daniel P. Funeriu, Nicolas Giuseppone. Dual-light control of nanomachines that
integrate motor and modulator subunits. Nature Nanotechnology 2017, 12 (6) , 540-545.
https://doi.org/10.1038/nnano.2017.28
  
736. V. Nicholas Vukotic, Kelong Zhu, Giorgio Baggi, Stephen J. Loeb. Optical Distinction between “Slow” and
“Fast” Translational Motion in Degenerate Molecular Shuttles. Angewandte Chemie 2017, 129 (22) , 6232-
6237. https://doi.org/10.1002/ange.201612549

737. V. Nicholas Vukotic, Kelong Zhu, Giorgio Baggi, Stephen J. Loeb. Optical Distinction between “Slow” and
“Fast” Translational Motion in Degenerate Molecular Shuttles. Angewandte Chemie International Edition 2017,
56 (22) , 6136-6141. https://doi.org/10.1002/anie.201612549

738. Kazumitsu Onizuka, Tomoko Chikuni, Takuya Amemiya, Takuya Miyashita, Kyoko Onizuka, Hiroshi Abe,
Fumi Nagatsugi. Pseudorotaxane formation via the slippage process with chemically cyclized
oligonucleotides. Nucleic Acids Research 2017, 45 (9) , 5036-5047. https://doi.org/10.1093/nar/gkx265

739. Dawei Zhang, James Robert Cochrane, Sebastiano Di Pietro, Laure Guy, Heinz Gornitzka, Jean-Pierre
Dutasta, Alexandre Martinez. “Breathing” Motion of a Modulable Molecular Cavity. Chemistry - A European
Journal 2017, 23 (27) , 6495-6498. https://doi.org/10.1002/chem.201700395

740. Peter Štacko, Jos C. M. Kistemaker, Ben L. Feringa. Fluorine-Substituted Molecular Motors with a
Quaternary Stereogenic Center. Chemistry - A European Journal 2017, 23 (27) , 6643-6653.
https://doi.org/10.1002/chem.201700581

741. A. Feigel. Dynamics of a mechanical system with multiple degrees of freedom out of thermal
equilibrium. Physical Review E 2017, 95 (5) https://doi.org/10.1103/PhysRevE.95.052106

742. Hiroki Iida, Kenji Ohmura, Ryuta Noda, Soichiro Iwahana, Hiroshi Katagiri, Naoki Ousaka, Taku Hayashi,
Yuh Hijikata, Stephan Irle, Eiji Yashima. Double-Stranded Helical Oligomers Covalently Bridged by Rotary Cyclic
Boronate Esters. Chemistry - An Asian Journal 2017, 12 (8) , 927-935.
https://doi.org/10.1002/asia.201700162

743. James P. Allen. Design of energy-transducing arti cial cells. Proceedings of the National Academy of
Sciences 2017, 114 (15) , 3790-3791. https://doi.org/10.1073/pnas.1703163114

744. Mathilde Berville, Sylvie Choua, Christophe Gourlaouen, Corinne Boudon, Laurent Ruhlmann, Corinne
Bailly, Saioa Cobo, Eric Saint-Aman, Jennifer Wytko, Jean Weiss. Flexible Viologen Cyclophanes: Odd/Even
Effects on Intramolecular Interactions. ChemPhysChem 2017, 18 (7) , 796-803.
https://doi.org/10.1002/cphc.201700011

745. Jason Y. C. Lim, Thanthapatra Bunchuay, Paul D. Beer. Strong and Selective Halide Anion Binding by
Neutral Halogen-Bonding [2]Rotaxanes in Wet Organic Solvents. Chemistry - A European Journal 2017, 23 (19)
, 4700-4707. https://doi.org/10.1002/chem.201700030

746. F. Box, E. Han, C. R. Tipton, T. Mullin. On the motion of linked spheres in a Stokes ow. Experiments in
Fluids 2017, 58 (4) https://doi.org/10.1007/s00348-017-2321-2

747. Ognjen Š. Miljanić. Small-Molecule Systems Chemistry. Chem 2017, 2 (4) , 502-524.
https://doi.org/10.1016/j.chempr.2017.03.002
748. Viktor Holubec, Artem Ryabov, Mohammad Yaghoubi, Martin Varga, Ayub Khodaee, M. Foulaadvand,
Petr Chvosta. Thermal Ratchet Effect in Con ning Geometries. Entropy 2017, 19 (4) , 119.
https://doi.org/10.3390/e19040119
  

749. Saikat Santra, Pradyut Ghosh. Rotamer-Induced Dynamic Nature of a [2]Rotaxane and Control of the
Dynamics by External Stimuli. European Journal of Organic Chemistry 2017, 2017 (12) , 1583-1593.
https://doi.org/10.1002/ejoc.201601525

750. Shuhei Kusano, Sae Konishi, Ryuta Ishikawa, Norihiro Sato, Satoshi Kawata, Fumi Nagatsugi, Osamu
Hayashida. Synthesis of Water-Soluble Triazinophanes and Evaluation of Their Molecular Recognition
Properties. European Journal of Organic Chemistry 2017, 2017 (12) , 1618-1623.
https://doi.org/10.1002/ejoc.201601663

751. Ryohei Yamakado, Yukina Ashida, Ryuma Sato, Yasuteru Shigeta, Nobuhiro Yasuda, Hiromitsu Maeda.
Cooperatively Interlocked [2+1]-Type π-System-Anion Complexes. Chemistry - A European Journal 2017, 23
(17) , 4160-4168. https://doi.org/10.1002/chem.201605765

752. Yangyang Yang, Ryu Tashiro, Yuki Suzuki, Tomoko Emura, Kumi Hidaka, Hiroshi Sugiyama, Masayuki
Endo. A Photoregulated DNA-Based Rotary System and Direct Observation of Its Rotational Movement.
Chemistry - A European Journal 2017, 23 (16) , 3979-3985. https://doi.org/10.1002/chem.201605616

753. Jun Cao. A theoretical study of the excited-state decay of acylhydrazones. International Journal of
Quantum Chemistry 2017, 117 (5) , e25330. https://doi.org/10.1002/qua.25330

754. Huang Wu, Yong Chen, Yu Liu. Reversibly Photoswitchable Supramolecular Assembly and Its
Application as a Photoerasable Fluorescent Ink. Advanced Materials 2017, 29 (10) , 1605271.
https://doi.org/10.1002/adma.201605271

755. Atash V. Gurbanov, Kamran T. Mahmudov, Maximilian N. Kopylovich, Fátima M. Guedes da Silva, Manas
Sutradhar, Firudin I. Guseinov, Fedor I. Zubkov, Abel M. Maharramov, Armando J.L. Pombeiro. Molecular
switching through cooperative ionic interactions and charge assisted hydrogen bonding. Dyes and Pigments
2017, 138 , 107-111. https://doi.org/10.1016/j.dyepig.2016.11.029

756. Grant M. Rotskoff. Mapping current uctuations of stochastic pumps to nonequilibrium steady states.
Physical Review E 2017, 95 (3) https://doi.org/10.1103/PhysRevE.95.030101

757. Malachi W. Gillick-Healy, Elizabeth V. Jennings, Helge Müller-Bunz, Yannick Ortin, Kirill Nikitin, Declan G.
Gilheany. Two Independent Orthogonal Stereomutations at a Single Asymmetric Center: A Narcissistic Couple.
Chemistry - A European Journal 2017, 23 (10) , 2332-2339. https://doi.org/10.1002/chem.201604080

758. Giulio Ragazzon, Alberto Credi, Benoit Colasson. Thermodynamic Insights on a Bistable Acid-Base
Switchable Molecular Shuttle with Strongly Shifted Co-conformational Equilibria. Chemistry - A European
Journal 2017, 23 (9) , 2149-2156. https://doi.org/10.1002/chem.201604783

759. Corina H. Pollok, Tim Riesebeck, Christian Merten. Photoisomerisierung eines Schalters auf Basis eines
chiralen Imins: Verfolgung durch Matrixisolations-VCD-Spektroskopie. Angewandte Chemie 2017, 129 (7) ,
1952-1955. https://doi.org/10.1002/ange.201610918

760. Corina H. Pollok, Tim Riesebeck, Christian Merten. Photoisomerization of a Chiral Imine Molecular
Switch Followed by Matrix-Isolation VCD Spectroscopy. Angewandte Chemie International Edition 2017, 56 (7)
, 1925-1928. https://doi.org/10.1002/anie.201610918
761. Luca Cera, Leonardo Chiappisi, Christoph Böttcher, Andrea Schulz, Stefan Schoder, Michael Gradzielski,
 
Christoph A. Schalley. PolyWhips: Directional Particle Transport by Gradient-Directed Growth and Stiffening of
Supramolecular Assemblies. Advanced Materials 2017, 29 (8) , 1604430.
https://doi.org/10.1002/adma.201604430

762. Boris Atenas, Sergio Curilef. Dynamics and thermodynamics of systems with long-range dipole-type
interactions. Physical Review E 2017, 95 (2) https://doi.org/10.1103/PhysRevE.95.022110

763. Liu-Pan Yang, Fei Jia, Qing-Hai Zhou, Fangfang Pan, Jiao-Nan Sun, Kari Rissanen, Lung Wa Chung, Wei
Jiang. Guest-Induced Folding and Self-Assembly of Conformationally Adaptive Macrocycles into Nanosheets
and Nanotubes. Chemistry - A European Journal 2017, 23 (7) , 1516-1520.
https://doi.org/10.1002/chem.201605701

764. Udaya K. Jayasundara, HyunJong Kim, Krishna P. Sahteli, Joseph I. Cline, Thomas W. Bell. Proton-Gated
Photoisomerization of Amino-Substituted Dibenzofulvene Rotors. ChemPhysChem 2017, 18 (1) , 59-63.
https://doi.org/10.1002/cphc.201601214

765. Jinling Cheng, Peter Štacko, Petra Rudolf, Régis Y. N. Gengler, Ben L. Feringa. Bidirectional
Photomodulation of Surface Tension in Langmuir Films. Angewandte Chemie 2017, 129 (1) , 297-302.
https://doi.org/10.1002/ange.201611187

766. Jinling Cheng, Peter Štacko, Petra Rudolf, Régis Y. N. Gengler, Ben L. Feringa. Bidirectional
Photomodulation of Surface Tension in Langmuir Films. Angewandte Chemie International Edition 2017, 56
(1) , 291-296. https://doi.org/10.1002/anie.201611187

767. K. Kanagaraj, M. Alagesan, Y. Inoue, C. Yang. Solvation Effects in Supramolecular Chemistry. 2017,,, 11-
60. https://doi.org/10.1016/B978-0-12-409547-2.12481-3

768. G. Ragazzon, M. Baroncini, P. Ceroni, A. Credi, M. Venturi. Electrochemically Controlled Supramolecular


Switches and Machines. 2017,,, 343-368. https://doi.org/10.1016/B978-0-12-409547-2.12498-9

769. Y. Zhu, B. Zhiyu, V. Kalavakunda, N.S. Hosmane. Endofullerenes and Carboranes. 2017,,, 479-487.
https://doi.org/10.1016/B978-0-12-409547-2.12524-7

770. S. Dasgupta, J. Wu. Template-Directed Synthesis of Catenanes. 2017,,, 251-268.


https://doi.org/10.1016/B978-0-12-409547-2.12588-0

771. James E. M. Lewis, Marzia Galli, Stephen M. Goldup. Properties and emerging applications of
mechanically interlocked ligands. Chemical Communications 2017, 53 (2) , 298-312.
https://doi.org/10.1039/C6CC07377H

772. Ning-Ning Yang, Wei Sun, Fu-Gui Xi, Qi Sui, Li-Jun Chen, En-Qing Gao. Postsynthetic N-methylation
making a metal–organic framework responsive to alkylamines. Chemical Communications 2017, 53 (10) ,
1747-1750. https://doi.org/10.1039/C6CC10278F

773. Rémi Merindol, Andreas Walther. Materials learning from life: concepts for active, adaptive and
autonomous molecular systems. Chemical Society Reviews 2017, 46 (18) , 5588-5619.
https://doi.org/10.1039/C6CS00738D
774. Nem Singh, Dongwook Kim, Dong Hwan Kim, Eun-Hee Kim, Hyunuk Kim, Myoung Soo Lah, Ki-Whan Chi.
Selective synthesis of iridium( iii )-derived molecular Borromean rings, [2]catenane and ring-in-ring
  
macrocycles via coordination-driven self-assembly. Dalton Transactions 2017, 46 (2) , 571-577.
https://doi.org/10.1039/C6DT04512J

775. Jan Maciejewski, Adam Sobczuk, Alexis Claveau, Adrien Nicolai, Riccardo Petraglia, Luca Cervini, Emilie
Baudat, Pascal Miéville, Daniele Fazzi, Clemence Corminboeuf, Giuseppe Sforazzini. Photochromic Torsional
Switch (PTS): a light-driven actuator for the dynamic tuning of π-conjugation extension. Chemical Science
2017, 8 (1) , 361-365. https://doi.org/10.1039/C6SC03196J

776. Marianna Rossetti, Simona Ranallo, Andrea Idili, Giuseppe Palleschi, Alessandro Porchetta, Francesco
Ricci. Allosteric DNA nanoswitches for controlled release of a molecular cargo triggered by biological inputs.
Chemical Science 2017, 8 (2) , 914-920. https://doi.org/10.1039/C6SC03404G

777. Qi Sui, Xiang-Ting Ren, Yu-Xiang Dai, Kai Wang, Wen-Tao Li, Teng Gong, Jia-Jia Fang, Bo Zou, En-Qing
Gao, Lin Wang. Piezochromism and hydrochromism through electron transfer: new stories for viologen
materials. Chemical Science 2017, 8 (4) , 2758-2768. https://doi.org/10.1039/C6SC04579K

778. Xiaowei Li, Xiangyang Yuan, Pengchi Deng, Lixi Chen, Yi Ren, Chengyuan Wang, Lixin Wu, Wen Feng,
Bing Gong, Lihua Yuan. Macrocyclic shape-persistency of cyclo[6]aramide results in enhanced multipoint
recognition for the highly e cient template-directed synthesis of rotaxanes. Chemical Science 2017, 8 (3) ,
2091-2100. https://doi.org/10.1039/C6SC04714A

779. Fangli Zhao, Lutz Grubert, Stefan Hecht, David Bléger. Orthogonal switching in four-state azobenzene
mixed-dimers. Chemical Communications 2017, 53 (23) , 3323-3326. https://doi.org/10.1039/C7CC00504K

780. Abir Goswami, Indrajit Paul, Michael Schmittel. Three-component nanorotors generated from fusion of
complexes and post-fusion metal–metal exchange. Chemical Communications 2017, 53 (37) , 5186-5189.
https://doi.org/10.1039/C7CC01977G

781. Xu-Sheng Du, Chun-Yu Wang, Qiong Jia, Rong Deng, Hua-Sheng Tian, Hou-Yu Zhang, Kamel Meguellati,
Ying-Wei Yang. Pillar[5]arene-based [1]rotaxane: high-yield synthesis, characterization and application in
Knoevenagel reaction. Chemical Communications 2017, 53 (38) , 5326-5329.
https://doi.org/10.1039/C7CC02364B

782. Guido Orlandini, Giulio Ragazzon, Valeria Zanichelli, Andrea Secchi, Serena Silvi, Margherita Venturi,
Arturo Arduini, Alberto Credi. Covalent capture of oriented calix[6]arene rotaxanes by a metal-free active
template approach. Chemical Communications 2017, 53 (45) , 6172-6174.
https://doi.org/10.1039/C7CC02859H

783. Thomas van Leeuwen, G. Henrieke Heideman, Depeng Zhao, Sander J. Wezenberg, Ben L. Feringa. In
situ control of polymer helicity with a non-covalently bound photoresponsive molecular motor dopant. Chem.
Commun. 2017, 53 (48) , 6393-6396. https://doi.org/10.1039/C7CC03188B

784. Synøve Ø. Scottwell, Jonathan E. Barnsley, C. John McAdam, Keith C. Gordon, James D. Crowley. A
ferrocene based switchable molecular folding ruler. Chemical Communications 2017, 53 (54) , 7628-7631.
https://doi.org/10.1039/C7CC03358C

785. Zhan-Qi Cao, Yi-Chuan Wang, Ai-Hua Zou, Gábor London, Qi Zhang, Chuan Gao, Da-Hui Qu. Reversible
switching of a supramolecular morphology driven by an amphiphilic bistable [2]rotaxane. Chemical
Communications 2017, 53 (62) , 8683-8686. https://doi.org/10.1039/C7CC05008A
786. Debabrata Samanta, Indrajit Paul, Michael Schmittel. Supramolecular ve-component nano-oscillator.
Chem. Commun. 2017, 53 (70) , 9709-9712. https://doi.org/10.1039/C7CC05235A   

787. Hendrik V. Schröder, Jan M. Wollschläger, Christoph A. Schalley. Redox-controlled self-inclusion of a


lasso-type pseudo[1]rotaxane. Chemical Communications 2017, 53 (66) , 9218-9221.
https://doi.org/10.1039/C7CC05259F

788. Gregory T. Rushton, Erik C. Vik, William G. Burns, Roger D. Rasberry, Ken D. Shimizu. Guest control of a
hydrogen bond-catalysed molecular rotor. Chemical Communications 2017, 53 (92) , 12469-12472.
https://doi.org/10.1039/C7CC07672J

789. Atsushi Fujiwara, Yusuke Inagaki, Hiroyuki Momma, Eunsang Kwon, Kentaro Yamaguchi, Manabu
Kanno, Hirohiko Kono, Wataru Setaka. A crystalline molecular gyrotop with a biphenylene dirotor and its
temperature-dependent birefringence. CrystEngComm 2017, 19 (40) , 6049-6056.
https://doi.org/10.1039/C7CE01081H

790. Katsuhiko Ariga, Taizo Mori, Waka Nakanishi, Jonathan P. Hill. Solid surface vs. liquid surface:
nanoarchitectonics, molecular machines, and DNA origami. Physical Chemistry Chemical Physics 2017, 19
(35) , 23658-23676. https://doi.org/10.1039/C7CP02280H

791. Ekaterina Y. Chernikova, Daria V. Berdnikova, Yuri V. Fedorov, Olga A. Fedorova, François Maurel,
Gediminas Jonusauskas. Light-induced piston nanoengines: ultrafast shuttling of a styryl dye inside
cucurbit[7]uril. Phys. Chem. Chem. Phys. 2017, 19 (38) , 25834-25839. https://doi.org/10.1039/C7CP04283C

792. Cristian Pezzato, Chuyang Cheng, J. Fraser Stoddart, R. Dean Astumian. Mastering the non-equilibrium
assembly and operation of molecular machines. Chemical Society Reviews 2017, 46 (18) , 5491-5507.
https://doi.org/10.1039/C7CS00068E

793. James E. M. Lewis, Paul D. Beer, Stephen J. Loeb, Stephen M. Goldup. Metal ions in the synthesis of
interlocked molecules and materials. Chemical Society Reviews 2017, 46 (9) , 2577-2591.
https://doi.org/10.1039/C7CS00199A

794. Salma Kassem, Thomas van Leeuwen, Anouk S. Lubbe, Miriam R. Wilson, Ben L. Feringa, David A. Leigh.
Arti cial molecular motors. Chemical Society Reviews 2017, 46 (9) , 2592-2621.
https://doi.org/10.1039/C7CS00245A

795. Leilei Xu, Fangzhi Mou, Haotian Gong, Ming Luo, Jianguo Guan. Light-driven micro/nanomotors: from
fundamentals to applications. Chemical Society Reviews 2017, 46 (22) , 6905-6926.
https://doi.org/10.1039/C7CS00516D

796. Merve S. Özer, Anup Rana, Pronay K. Biswas, Michael Schmittel. Four-component zinc-porphyrin/zinc-
salphen nanorotor. Dalton Transactions 2017, 46 (29) , 9491-9497. https://doi.org/10.1039/C7DT01323J

797. Lauren R. Holloway, Paul M. Bogie, Richard J. Hooley. Controlled self-sorting in self-assembled cage
complexes. Dalton Transactions 2017, 46 (43) , 14719-14723. https://doi.org/10.1039/C7DT03399K

798. Bryan Lau, Ofer Kedem, James Schwabacher, Daniel Kwasnieski, Emily A. Weiss. An introduction to
ratchets in chemistry and biology. Materials Horizons 2017, 4 (3) , 310-318.
https://doi.org/10.1039/C7MH00062F
799. Karolina M. Tomczyk, Mateusz Woźny, Sławomir Domagała, Agnieszka Wiȩckowska, Joanna
Pawłowska, Krzysztof Woźniak, Bohdan Korybut-Daszkiewicz. Rotaxanes composed of dibenzo-24-crown-8
  
and macrocyclic transition metal complexing tetraimine units. New Journal of Chemistry 2017, 41 (13) , 6004-
6013. https://doi.org/10.1039/C7NJ00909G

800. Marcos A. P. Martins, Geórgia C. Zimmer, Leticia V. Rodrigues, Tainára Orlando, Lilian Buriol, Mateo
Alajarin, Jose Berna, Clarissa P. Frizzo, Helio G. Bonacorso, Nilo Zanatta. Competition between the donor and
acceptor hydrogen bonds of the threads in the formation of [2]rotaxanes by clipping reaction. New Journal of
Chemistry 2017, 41 (22) , 13303-13318. https://doi.org/10.1039/C7NJ02443F

801. Jijun Yan, Chuanqing Kang, Zheng Bian, Rizhe Jin, Xiaoye Ma, Lianxun Gao. Supramolecular self-
assembly of chiral polyimides driven by repeat units and end groups. New Journal of Chemistry 2017, 41 (23) ,
14723-14729. https://doi.org/10.1039/C7NJ02451G

802. Giacomo Mariani, Antoine Goujon, Emilie Moulin, Michel Rawiso, Nicolas Giuseppone, Eric Buhler.
Integration of molecular machines into supramolecular materials: actuation between equilibrium polymers and
crystal-like gels. Nanoscale 2017, 9 (46) , 18456-18466. https://doi.org/10.1039/C7NR04251E

803. Beth E. Fletcher, Michael J. G. Peach, Nicholas H. Evans. Rapidly accessible “click” rotaxanes utilizing a
single amide hydrogen bond templating motif. Organic & Biomolecular Chemistry 2017, 15 (13) , 2797-2803.
https://doi.org/10.1039/C7OB00284J

804. Asha Brown, Thomas Lang, Kathleen M. Mullen, Paul D. Beer. Active metal template synthesis of a
neutral indolocarbazole-containing [2]rotaxane host system for selective oxoanion recognition. Organic &
Biomolecular Chemistry 2017, 15 (21) , 4587-4594. https://doi.org/10.1039/C7OB01040K

805. Tomoyuki Ikai, Yuya Wada, Seiya Awata, Changsik Yun, Katsuhiro Maeda, Motohiro Mizuno, Timothy M.
Swager. Chiral triptycene-pyrene π-conjugated chromophores with circularly polarized luminescence. Org.
Biomol. Chem. 2017, 15 (39) , 8440-8447. https://doi.org/10.1039/C7OB02046E

806. Ashot Gevorgyan, Satenik Mkrtchyan, Tatevik Grigoryan, Viktor O. Iaroshenko. Disilanes as oxygen
scavengers and surrogates of hydrosilanes suitable for selective reduction of nitroarenes, phosphine oxides
and other valuable substrates. Organic Chemistry Frontiers 2017, 4 (12) , 2437-2444.
https://doi.org/10.1039/C7QO00566K

807. J. Śniechowska, P. Paluch, M. J. Potrzebowski. Structure and dynamics processes in free-base chlorins
controlled by chemical modi cations of macroring and aryl groups in meso-positions. RSC Advances 2017, 7
(40) , 24795-24805. https://doi.org/10.1039/C7RA02217D

808. Alberto Martinez-Cuezva, Adrian Saura-Sanmartin, Tomas Nicolas-Garcia, Cristian Navarro, Raul-Angel
Orenes, Mateo Alajarin, Jose Berna. Photoswitchable interlocked thiodiglycolamide as a cocatalyst of a
chalcogeno-Baylis–Hillman reaction. Chemical Science 2017, 8 (5) , 3775-3780.
https://doi.org/10.1039/C7SC00724H

809. Kelong Zhu, Giorgio Baggi, V. Nicholas Vukotic, Stephen J. Loeb. Reversible mechanical protection:
building a 3D “suit” around a T-shaped benzimidazole axle. Chemical Science 2017, 8 (5) , 3898-3904.
https://doi.org/10.1039/C7SC00790F

810. Haohao Fu, Xueguang Shao, Christophe Chipot, Wensheng Cai. The lubricating role of water in the
shuttling of rotaxanes. Chemical Science 2017, 8 (7) , 5087-5094. https://doi.org/10.1039/C7SC01593C
811. Reece Beekmeyer, Michael A. Parkes, Luke Ridgwell, Jamie W. Riley, Jiawen Chen, Ben L. Feringa,
Andrew Kerridge, Helen H. Fielding. Unravelling the electronic structure and dynamics of an isolated molecular
rotary motor in the gas-phase. Chemical Science 2017, 8 (9) , 6141-6148.
  
https://doi.org/10.1039/C7SC01997A

812. Guillaume De Bo, David A. Leigh, Charlie T. McTernan, Shoufeng Wang. A complementary pair of
enantioselective switchable organocatalysts. Chemical Science 2017, 8 (10) , 7077-7081.
https://doi.org/10.1039/C7SC02462B

813. Hendrik V. Schröder, Sebastian Sobottka, Maite Nößler, Henrik Hupatz, Marius Gaedke, Biprajit Sarkar,
Christoph A. Schalley. Impact of mechanical bonding on the redox-switching of tetrathiafulvalene in crown
ether–ammonium [2]rotaxanes. Chemical Science 2017, 8 (9) , 6300-6306.
https://doi.org/10.1039/C7SC02694C

814. T. H. Ngo, J. Labuta, G. N. Lim, W. A. Webre, F. D'Souza, P. A. Karr, J. E. M. Lewis, J. P. Hill, K. Ariga, S. M.
Goldup. Porphyrinoid rotaxanes: building a mechanical picket fence. Chemical Science 2017, 8 (9) , 6679-6685.
https://doi.org/10.1039/C7SC03165C

815. Benjamin Tam, Ozgur Yazaydin. Design of electric eld controlled molecular gates mounted on metal–
organic frameworks. Journal of Materials Chemistry A 2017, 5 (18) , 8690-8696.
https://doi.org/10.1039/C7TA00101K

816. Zhong-Liang Gong, Yu-Wu Zhong, Jiannian Yao. Regulation of intra- and intermolecular Pt–Pt and π–π
interactions of a U-shaped diplatinum complex to achieve pseudo-polymorphic emissions in solution and
crystalline states. Journal of Materials Chemistry C 2017, 5 (29) , 7222-7229.
https://doi.org/10.1039/C7TC02282D

817. Qi Sui, Ye Yuan, Ning-Ning Yang, Xin Li, Teng Gong, En-Qing Gao, Lin Wang. Coordination-modulated
piezochromism in metal–viologen materials. Journal of Materials Chemistry C 2017, 5 (47) , 12400-12408.
https://doi.org/10.1039/C7TC04208F

818. Pawel Kulakowski, Kamil Solarczyk, Krzysztof Wojcik. Routing in FRET-Based Nanonetworks. IEEE
Communications Magazine 2017, 55 (9) , 218-224. https://doi.org/10.1109/MCOM.2017.1600760

819. Daisuke AOKI, Toshikazu TAKATA. Branched/Linear Polymer Topology Transformation Facilitated by
Mechanical Linking of Polymer Chains. NIPPON GOMU KYOKAISHI 2017, 90 (6) , 283-289.
https://doi.org/10.2324/gomu.90.283

820. James Lewis, Joby Winn, Stephen Goldup. Stepwise, Protecting Group Free Synthesis of [4]Rotaxanes.
Molecules 2017, 22 (1) , 89. https://doi.org/10.3390/molecules22010089

821. Naoki Ousaka, Eiji Yashima. Development of Helical Oligomers and Polymers Capable of Extension and
Contraction Motions. Journal of Synthetic Organic Chemistry, Japan 2017, 75 (5) , 466-475.
https://doi.org/10.5059/yukigoseikyokaishi.75.466

822. Danijel Boskovic, Sivakumar Balakrishnan, Shaoming Huang, Gerhard F. Swiegers. Kinematic molecular
manufacturing machines. Coordination Chemistry Reviews 2016, 329 , 163-190.
https://doi.org/10.1016/j.ccr.2016.09.007

823. Zhikai Xu, Mujuan Chen, Jidong Liang, Suhui Zhang, Manli Guo, Ying Li, Lasheng Jiang. High-yield
preparation of new crown ether-based cryptands and improving complexation with paraquat, paraquat
derivatives and diquat. Tetrahedron 2016, 72 (49) , 8009-8014. https://doi.org/10.1016/j.tet.2016.10.030

  
824. Kelong Zhu, V. Nicholas Vukotic, Stephen J. Loeb. Acid-Base Switchable [2]- and [3]Rotaxane Molecular
Shuttles with Benzimidazolium and Bis(pyridinium) Recognition Sites. Chemistry - An Asian Journal 2016, 11
(22) , 3258-3266. https://doi.org/10.1002/asia.201601179

825. . Rotaxanes and Polyrotaxanes. 2016,,, 231-280. https://doi.org/10.1002/9781118994405.ch6

826. Mirco Zerbetto, Antonino Polimeno. Multiscale modeling for interpreting nuclear magnetic resonance
relaxation in exible molecules. International Journal of Quantum Chemistry 2016, 116 (22) , 1706-1722.
https://doi.org/10.1002/qua.25215

827. David A. Leigh. Die Genese der Nanomaschinen: der Chemie-Nobelpreis 2016. Angewandte Chemie
2016, 128 (47) , 14722-14724. https://doi.org/10.1002/ange.201609841

828. David A. Leigh. Genesis of the Nanomachines: The 2016 Nobel Prize in Chemistry. Angewandte Chemie
International Edition 2016, 55 (47) , 14506-14508. https://doi.org/10.1002/anie.201609841

829. Baswanth Oruganti, Jun Wang, Bo Durbeej. Computational Insight to Improve the Thermal Isomerisation
Performance of Overcrowded Alkene-Based Molecular Motors through Structural Redesign. ChemPhysChem
2016, 17 (21) , 3399-3408. https://doi.org/10.1002/cphc.201600766

830. Qiuhua Liu, Min Tang, Wennan Zeng, Xi Zhang, Jianxiu Wang, Zaichun Zhou. Optimal Size Matching and
Minimal Distortion Energy: Implications for Natural Selection by the Macrocycle of the Iron Species in Heme.
European Journal of Inorganic Chemistry 2016, 2016 (33) , 5222-5229.
https://doi.org/10.1002/ejic.201600883

831. Xuan Wu, Lei Gao, Junzhao Sun, Xiao-Yu Hu, Leyong Wang. Stable pillar[5]arene-based
pseudo[1]rotaxanes formed in polar solution. Chinese Chemical Letters 2016, 27 (11) , 1655-1660.
https://doi.org/10.1016/j.cclet.2016.05.004

832. Anne S. Meeussen, Jayson Paulose, Vincenzo Vitelli. Geared Topological Metamaterials with Tunable
Mechanical Stability. Physical Review X 2016, 6 (4) https://doi.org/10.1103/PhysRevX.6.041029

833. Yasuhiro Shirai, Kosuke Minami, Waka Nakanishi, Yusuke Yonamine, Christian Joachim, Katsuhiko Ariga.
Driving nanocars and nanomachines at interfaces: From concept of nanoarchitectonics to actual use in world
wide race and hand operation. Japanese Journal of Applied Physics 2016, 55 (11) , 1102A2.
https://doi.org/10.7567/JJAP.55.1102A2

834. . An Introduction to the Mechanical Bond. 2016,,, 1-54. https://doi.org/10.1002/9781119044123.ch1

835. . Molecular Switches and Machines with Mechanical Bonds. 2016,,, 555-733.
https://doi.org/10.1002/9781119044123.ch6

836. Yoshiaki Amatatsu. Computational Design of a Fluorene-Based Ethylenoid Bridged by Trimethylene


Chain. Bulletin of the Chemical Society of Japan 2016, 89 (10) , 1245-1259.
https://doi.org/10.1246/bcsj.20160161

837. Zheng Meng, Bo-Yang Wang, Jun-Feng Xiang, Qiang Shi, Chuan-Feng Chen. Self-Assembly of a
[2]Pseudorotaxane by an Inchworm-Motion Mechanism. Chemistry - A European Journal 2016, 22 (42) ,
15075-15084. https://doi.org/10.1002/chem.201602785

838. Benoit Colasson, Alberto Credi, Giulio Ragazzon. Light-driven molecular machines based on
  
ruthenium(II) polypyridine complexes: Strategies and recent advances. Coordination Chemistry Reviews 2016,
325 , 125-134. https://doi.org/10.1016/j.ccr.2016.02.012

839. Ai Ito, Hiroto Fujino, Keiko Ushiyama, Eriko Yamanaka, Ryu Yamasaki, Iwao Okamoto. Acid-controlled
switching of conformational preference of N,N-diarylamides bearing pyridine. Tetrahedron Letters 2016, 57
(42) , 4737-4741. https://doi.org/10.1016/j.tetlet.2016.09.035

840. Yuping Wang, Junling Sun, Zhichang Liu, Majed S. Nassar, Youssry Y. Botros, J. Fraser Stoddart.
Symbiotic Control in Mechanical Bond Formation. Angewandte Chemie 2016, 128 (40) , 12575-12580.
https://doi.org/10.1002/ange.201605454

841. Yuping Wang, Junling Sun, Zhichang Liu, Majed S. Nassar, Youssry Y. Botros, J. Fraser Stoddart.
Symbiotic Control in Mechanical Bond Formation. Angewandte Chemie International Edition 2016, 55 (40) ,
12387-12392. https://doi.org/10.1002/anie.201605454

842. Felix B. Schwarz, Thomas Heinrich, J. Ole Kaufmann, Andreas Lippitz, Rakesh Puttreddy, Kari Rissanen,
Wolfgang E. S. Unger, Christoph A. Schalley. Photocontrolled On-Surface Pseudorotaxane Formation with Well-
Ordered Macrocycle Multilayers. Chemistry - A European Journal 2016, 22 (40) , 14383-14389.
https://doi.org/10.1002/chem.201603156

843. Yoshinao Shinozaki, Yuta Mizumura, Yuki Kida, Kosuke Sugawa, Kaori Wada, Sachiko Kishiro, Joe
Otsuki. Crystal Structures and Side-arm Dynamics of Cerium meso -Tetrathienylporphyrin Double-decker
Complexes. Chemistry Letters 2016, 45 (9) , 1123-1125. https://doi.org/10.1246/cl.160549

844. Sudhakar Gaikwad, Abir Goswami, Soumen De, Michael Schmittel. Ein metallregulierter vierstu ger
Nanoschalter zur Steuerung einer zweistu gen sequenziellen Katalyse in einem Elf-Komponenten-System.
Angewandte Chemie 2016, 128 (35) , 10668-10673. https://doi.org/10.1002/ange.201604658

845. Sudhakar Gaikwad, Abir Goswami, Soumen De, Michael Schmittel. A Metalloregulated Four-State
Nanoswitch Controls Two-Step Sequential Catalysis in an Eleven-Component System. Angewandte Chemie
International Edition 2016, 55 (35) , 10512-10517. https://doi.org/10.1002/anie.201604658

846. Takuo Minato, Kosuke Suzuki, Kazuya Yamaguchi, Noritaka Mizuno. Synthesis and
Disassembly/Reassembly of Giant Ring-Shaped Polyoxotungstate Oligomers. Angewandte Chemie 2016, 128
(33) , 9782-9785. https://doi.org/10.1002/ange.201604094

847. Takuo Minato, Kosuke Suzuki, Kazuya Yamaguchi, Noritaka Mizuno. Synthesis and
Disassembly/Reassembly of Giant Ring-Shaped Polyoxotungstate Oligomers. Angewandte Chemie
International Edition 2016, 55 (33) , 9630-9633. https://doi.org/10.1002/anie.201604094

848. Stephanie C. Everhart, Udaya K. Jayasundara, HyunJong Kim, Rolando Procúpez-Schtirbu, Wayne A.
Stanbery, Clay H. Mishler, Brian J. Frost, Joseph I. Cline, Thomas W. Bell. Synthesis and Photoisomerization of
Substituted Dibenzofulvene Molecular Rotors. Chemistry - A European Journal 2016, 22 (32) , 11291-11302.
https://doi.org/10.1002/chem.201600854

849. Nasim Farahani, Kelong Zhu, Christopher A. O'Keefe, Robert W. Schurko, Stephen J. Loeb. Thermally
Driven Dynamics of a Rotaxane Wheel about an Imidazolium Axle inside a Metal-Organic Framework.
ChemPlusChem 2016, 81 (8) , 836-841. https://doi.org/10.1002/cplu.201600176
850. Fei Jiang, Mujuan Chen, Jidong Liang, Zhihong Gao, Mingfei Tang, Zhikai Xu, Bin Peng, Shizheng Zhu,
Lasheng Jiang. Sailboat-Shaped Self-Complexes that Function as Controllable Rotary Switches. 
European 
Journal of Organic Chemistry 2016, 2016 (20) , 3310-3315. https://doi.org/10.1002/ejoc.201600383

851. Flory Wong, Krishna Kanti Dey, Ayusman Sen. Synthetic Micro/Nanomotors and Pumps: Fabrication and
Applications. Annual Review of Materials Research 2016, 46 (1) , 407-432. https://doi.org/10.1146/annurev-
matsci-070115-032047

852. Zhao Gao, Yifei Han, Jiangjun Chen, Xiao Wang, Feng Wang. Trimethylammonium-Derived Molecular
Tweezers and Their Host-Guest Complexation Behaviours in Polar Media. Chemistry - An Asian Journal 2016,
11 (12) , 1775-1779. https://doi.org/10.1002/asia.201600364

853. Thibaut Legigan, Benjamin Riss-Yaw, Caroline Clavel, Frédéric Coutrot. Active Esters as Pseudostoppers
for Slippage Synthesis of [2]Pseudorotaxane Building Blocks: A Straightforward Route to Multi-Interlocked
Molecular Machines. Chemistry - A European Journal 2016, 22 (26) , 8835-8847.
https://doi.org/10.1002/chem.201601286

854. Paola Franchi, Valentina Bleve, Elisabetta Mezzina, Christian Schäfer, Giulio Ragazzon, Marco Albertini,
Donatella Carbonera, Alberto Credi, Marilena Di Valentin, Marco Lucarini. Structural Changes of a Doubly Spin-
Labeled Chemically Driven Molecular Shuttle Probed by PELDOR Spectroscopy. Chemistry - A European
Journal 2016, 22 (26) , 8745-8750. https://doi.org/10.1002/chem.201601407

855. Ming Cheng, Li Liu, Yihan Cao, Juli Jiang, Leyong Wang. A Phosphine Oxide Functional Group Based
[2]Rotaxane That Operates as a Multistable Molecular Shuttle. ChemPhysChem 2016, 17 (12) , 1835-1839.
https://doi.org/10.1002/cphc.201501016

856. Qi Zhang, Da-Hui Qu. Arti cial Molecular Machine Immobilized Surfaces: A New Platform To Construct
Functional Materials. ChemPhysChem 2016, 17 (12) , 1759-1768. https://doi.org/10.1002/cphc.201501048

857. Alberto Martinez-Cuezva, Fernando Carro-Guillen, Aurelia Pastor, Marta Marin-Luna, Raul-Angel Orenes,
Mateo Alajarin, Jose Berna. Co-conformational Exchange Triggered by Molecular Recognition in a
Di(acylamino)pyridine-Based Molecular Shuttle Containing Two Pyridine Rings at the Macrocycle.
ChemPhysChem 2016, 17 (12) , 1920-1926. https://doi.org/10.1002/cphc.201501060

858. Chuyang Cheng, J. Fraser Stoddart. Wholly Synthetic Molecular Machines. ChemPhysChem 2016, 17
(12) , 1780-1793. https://doi.org/10.1002/cphc.201501155

859. Gloria Tabacchi, Serena Silvi, Margherita Venturi, Alberto Credi, Ettore Fois. Dethreading of a Photoactive
Azobenzene-Containing Molecular Axle from a Crown Ether Ring: A Computational Investigation.
ChemPhysChem 2016, 17 (12) , 1913-1919. https://doi.org/10.1002/cphc.201501160

860. Dhiredj C. Jagesar, Piet G. Wiering, Euan R. Kay, David A. Leigh, Albert M. Brouwer. Successive
Translocation of the Rings in a [3]Rotaxane. ChemPhysChem 2016, 17 (12) , 1902-1912.
https://doi.org/10.1002/cphc.201501162

861. Nasim Farahani, Kelong Zhu, Stephen J. Loeb. Rigid, Bistable Molecular Shuttles Combining T-shaped
Benzimidazolium and Y-shaped Imidazolium Recognition Sites. ChemPhysChem 2016, 17 (12) , 1875-1880.
https://doi.org/10.1002/cphc.201600128
862. R. Dean Astumian, Shayantani Mukherjee, Arieh Warshel. The Physics and Physical Chemistry of
Molecular Machines. ChemPhysChem 2016, 17 (12) , 1719-1741. https://doi.org/10.1002/cphc.201600184
  
863. Benjamin Riss-Yaw, Philip Waelès, Frédéric Coutrot. Reverse Anomeric Effect in Large-Amplitude
Pyridinium Amide-Containing Mannosyl [2]Rotaxane Molecular Shuttles. ChemPhysChem 2016, 17 (12) , 1860-
1869. https://doi.org/10.1002/cphc.201600253

864. José Augusto Berrocal, Chiara Biagini, Luigi Mandolini, Stefano Di Stefano. Coupling of the
Decarboxylation of 2-Cyano-2-phenylpropanoic Acid to Large-Amplitude Motions: A Convenient Fuel for an
Acid-Base-Operated Molecular Switch. Angewandte Chemie 2016, 128 (24) , 7111-7115.
https://doi.org/10.1002/ange.201602594

865. José Augusto Berrocal, Chiara Biagini, Luigi Mandolini, Stefano Di Stefano. Coupling of the
Decarboxylation of 2-Cyano-2-phenylpropanoic Acid to Large-Amplitude Motions: A Convenient Fuel for an
Acid-Base-Operated Molecular Switch. Angewandte Chemie International Edition 2016, 55 (24) , 6997-7001.
https://doi.org/10.1002/anie.201602594

866. Miriam R. Wilson, Jordi Solà, Armando Carlone, Stephen M. Goldup, Nathalie Lebrasseur, David A. Leigh.
An autonomous chemically fuelled small-molecule motor. Nature 2016, 534 (7606) , 235-240.
https://doi.org/10.1038/nature18013

867. Kazuhisa Iwaso, Yoshinori Takashima, Akira Harada. Fast response dry-type arti cial molecular muscles
with [c2]daisy chains. Nature Chemistry 2016, 8 (6) , 625-632. https://doi.org/10.1038/nchem.2513

868. Thomas van Leeuwen, Jefri Gan, Jos C. M. Kistemaker, Stefano F. Pizzolato, Mu-Chieh Chang, Ben L.
Feringa. Enantiopure Functional Molecular Motors Obtained by a Switchable Chiral-Resolution Process.
Chemistry - A European Journal 2016, 22 (21) , 7054-7058. https://doi.org/10.1002/chem.201600628

869. O. Raz, Y. Subaşı, C. Jarzynski. Mimicking Nonequilibrium Steady States with Time-Periodic Driving.
Physical Review X 2016, 6 (2) https://doi.org/10.1103/PhysRevX.6.021022

870. Nan Zhu, Kazuko Nakazono, Toshikazu Takata. Solid-state Rotaxane Switch: Synthesis of
Thermoresponsive Rotaxane Shuttle Utilizing a Thermally Decomposable Acid. Chemistry Letters 2016, 45 (4)
, 445-447. https://doi.org/10.1246/cl.151190

871. Alessandra Meli, Eleonora Macedi, Francesco De Riccardis, Vincent J. Smith, Leonard J. Barbour, Irene
Izzo, Consiglia Tedesco. Solid-State Conformational Flexibility at Work: Zipping and Unzipping within a Cyclic
Peptoid Single Crystal. Angewandte Chemie 2016, 128 (15) , 4757-4760.
https://doi.org/10.1002/ange.201511053

872. Alessandra Meli, Eleonora Macedi, Francesco De Riccardis, Vincent J. Smith, Leonard J. Barbour, Irene
Izzo, Consiglia Tedesco. Solid-State Conformational Flexibility at Work: Zipping and Unzipping within a Cyclic
Peptoid Single Crystal. Angewandte Chemie International Edition 2016, 55 (15) , 4679-4682.
https://doi.org/10.1002/anie.201511053

873. Xing-Dong Xu, Chang-Jiang Yao, Li-Jun Chen, Guang-Qiang Yin, Yu-Wu Zhong, Hai-Bo Yang. Facile
Construction of Structurally De ned Porous Membranes from Supramolecular Hexakistriphenylamine
Metallacycles through Electropolymerization. Chemistry - A European Journal 2016, 22 (15) , 5211-5218.
https://doi.org/10.1002/chem.201504480
874. Christian Schäfer, Giulio Ragazzon, Benoit Colasson, Marcello La Rosa, Serena Silvi, Alberto Credi. An
Arti cial Molecular Transporter. ChemistryOpen 2016, 5 (2) , 120-124.
https://doi.org/10.1002/open.201500217
  

875. Mihail Barboiu, Adrian-Mihail Stadler, Jean-Marie Lehn. Kontrollierte Faltungs-, Bewegungs- und
konstitutionelle Dynamik in polyheterocyclischen molekularen Strängen. Angewandte Chemie 2016, 128 (13) ,
4200-4225. https://doi.org/10.1002/ange.201505394

876. Mihail Barboiu, Adrian-Mihail Stadler, Jean-Marie Lehn. Controlled Folding, Motional, and Constitutional
Dynamic Processes of Polyheterocyclic Molecular Strands. Angewandte Chemie International Edition 2016, 55
(13) , 4130-4154. https://doi.org/10.1002/anie.201505394

877. Christopher Schweez, Philip Shushkov, Stefan Grimme, Sigurd Höger. Synthesis and Dynamics of
Nanosized Phenylene-Ethynylene-Butadiynylene Rotaxanes and the Role of Shape Persistence. Angewandte
Chemie 2016, 128 (10) , 3389-3394. https://doi.org/10.1002/ange.201509702

878. Christopher Schweez, Philip Shushkov, Stefan Grimme, Sigurd Höger. Synthesis and Dynamics of
Nanosized Phenylene-Ethynylene-Butadiynylene Rotaxanes and the Role of Shape Persistence. Angewandte
Chemie International Edition 2016, 55 (10) , 3328-3333. https://doi.org/10.1002/anie.201509702

879. Lutz Greb, Andreas Eichhöfer, Jean-Marie Lehn. Internal C-C Bond Rotation in Photoisomers of α-
Bisimines: a Light-Responsive Two-Step Molecular Speed Regulator Based on Double Imine Photoswitching.
European Journal of Organic Chemistry 2016, 2016 (7) , 1243-1246.
https://doi.org/10.1002/ejoc.201600113

880. Petko Ivanov. Performance of some DFT functionals with dispersion on modeling of the translational
isomers of a solvent-switchable [2]rotaxane. Journal of Molecular Structure 2016, 1107 , 31-38.
https://doi.org/10.1016/j.molstruc.2015.11.015

881. Salma Kassem, Alan T. L. Lee, David A. Leigh, Augustinas Markevicius, Jordi Solà. Pick-up, transport and
release of a molecular cargo using a small-molecule robotic arm. Nature Chemistry 2016, 8 (2) , 138-143.
https://doi.org/10.1038/nchem.2410

882. Matthew A. Watson, Scott L. Cockroft. An Autonomously Reciprocating Transmembrane Nanoactuator.


Angewandte Chemie 2016, 128 (4) , 1367-1371. https://doi.org/10.1002/ange.201508845

883. Matthew A. Watson, Scott L. Cockroft. An Autonomously Reciprocating Transmembrane Nanoactuator.


Angewandte Chemie International Edition 2016, 55 (4) , 1345-1349. https://doi.org/10.1002/anie.201508845

884. David Bléger. Orchestrating Molecular Motion with Light - From Single (macro)Molecules to Materials.
Macromolecular Chemistry and Physics 2016, 217 (2) , 189-198. https://doi.org/10.1002/macp.201500330

885. George R. Newkome. Eight-Membered and Larger Rings. 2016,,, 623-644. https://doi.org/10.1016/B978-
0-08-100755-6.00017-X

886. Russell N. Grimes. Carboranes in Other Applications. 2016,,, 985-1019. https://doi.org/10.1016/B978-0-


12-801894-1.00017-2

887. Ai Ito, Masanori Sato, Ryu Yamasaki, Iwao Okamoto. Acid- and base-induced conformational alterations
of N-aryl-N-troponyl amides. Tetrahedron Letters 2016, 57 (3) , 438-441.
https://doi.org/10.1016/j.tetlet.2015.12.048

  
888. Synøve Ø. Scottwell, James D. Crowley. Ferrocene-containing non-interlocked molecular machines.
Chemical Communications 2016, 52 (12) , 2451-2464. https://doi.org/10.1039/C5CC09569G

889. B. Nisar Ahamed, Roland Duchêne, Koen Robeyns, Charles-André Fustin. Catenane-based mechanically-
linked block copolymers. Chemical Communications 2016, 52 (10) , 2149-2152.
https://doi.org/10.1039/C5CC09775D

890. Matthew A. Watson, Scott L. Cockroft. Man-made molecular machines: membrane bound. Chemical
Society Reviews 2016, 45 (22) , 6118-6129. https://doi.org/10.1039/C5CS00874C

891. Zhan-Qi Cao, Zhou-Lin Luan, Qi Zhang, Rui-Rui Gu, Jun Ren, Da-Hui Qu. An acid/base responsive side-
chain polyrotaxane system with a uorescent signal. Polymer Chemistry 2016, 7 (10) , 1866-1870.
https://doi.org/10.1039/C5PY01944C

892. Xin Fu, Qi Zhang, Si-Jia Rao, Da-Hui Qu, He Tian. One-pot synthesis of a [c2]daisy-chain-containing
hetero[4]rotaxane via a self-sorting strategy. Chemical Science 2016, 7 (3) , 1696-1701.
https://doi.org/10.1039/C5SC04844C

893. Jia Shang, Brett M. Rambo, Xiang Hao, Jun-Feng Xiang, Han-Yuan Gong, Jonathan L. Sessler. Post-
synthetic modi cation of a macrocyclic receptor via regioselective imidazolium ring-opening. Chemical
Science 2016, 7 (7) , 4148-4157. https://doi.org/10.1039/C5SC04860E

894. Richard M. Parker, Dominic J. Wales, James C. Gates, Peter G. R. Smith, Martin C. Grossel. Tracking a
photo-switchable surface-localised supramolecular interaction via refractive index. Journal of Materials
Chemistry C 2016, 4 (6) , 1178-1185. https://doi.org/10.1039/C5TC03774C

895. Jiawen Chen, Sander J. Wezenberg, Ben L. Feringa. Intramolecular transport of small-molecule cargo in
a nanoscale device operated by light. Chemical Communications 2016, 52 (41) , 6765-6768.
https://doi.org/10.1039/C6CC02382G

896. Aldo C. Catalán, Jorge Tiburcio. Self-assembly of pseudo-rotaxane and rotaxane complexes using an
electrostatic slippage approach. Chemical Communications 2016, 52 (61) , 9526-9529.
https://doi.org/10.1039/C6CC04619C

897. Héctor Barbero, Sergio Ferrero, Lucía Álvarez-Miguel, Patricia Gómez-Iglesias, Daniel Miguel, Celedonio
M. Álvarez. A nity modulation of photoresponsive hosts for fullerenes: light-gated corannulene tweezers.
Chemical Communications 2016, 52 (88) , 12964-12967. https://doi.org/10.1039/C6CC06445K

898. Bo Qiao, Yun Liu, Semin Lee, Maren Pink, Amar H. Flood. A high-yield synthesis and acid–base response
of phosphate-templated [3]rotaxanes. Chemical Communications 2016, 52 (94) , 13675-13678.
https://doi.org/10.1039/C6CC08113D

899. Tian-Guang Zhan, Meng-Yan Yun, Jia-Le Lin, Xin-Yao Yu, Kang-Da Zhang. Dual absorption spectral
changes by light-triggered shuttling in bistable [2]rotaxanes with non-destructive readout. Chemical
Communications 2016, 52 (98) , 14085-14088. https://doi.org/10.1039/C6CC08314E

900. Felix B. Schwarz, Thomas Heinrich, Andreas Lippitz, Wolfgang E. S. Unger, Christoph A. Schalley. A
photoswitchable rotaxane operating in monolayers on solid support. Chemical Communications 2016, 52
(100) , 14458-14461. https://doi.org/10.1039/C6CC08586E

901. Yu Ohshima, Kazuya Kubo, Takashi Matsumoto, Heng-Yun Ye, Shin-ichiro Noro, Tomoyuki Akutagawa,
  
Takayoshi Nakamura. One-dimensional supramolecular columnar structure of trans-syn-trans-
dicyclohexano[18]crown-6 and organic ammonium cations. CrystEngComm 2016, 18 (41) , 7959-7964.
https://doi.org/10.1039/C6CE00980H

902. Laura R. Blackholly, Helena J. Shepherd, Jennifer R. Hiscock. ‘Frustrated’ hydrogen bond mediated
amphiphile self-assembly – a solid state study. CrystEngComm 2016, 18 (37) , 7021-7028.
https://doi.org/10.1039/C6CE01493C

903. Selvam Amudhan Senthan, Vedamanickam Alexander. Synthesis, luminescence, and electrochemical
studies of tetra- and octanuclear ruthenium( ii ) complexes of tolylterpyridine appended calixarenes. New
Journal of Chemistry 2016, 40 (12) , 10064-10070. https://doi.org/10.1039/C6NJ02564A

904. Xie Han, Guotao Liu, Sheng Hua Liu, Jun Yin. Synthesis of rotaxanes and catenanes using an imine
clipping reaction. Organic & Biomolecular Chemistry 2016, 14 (44) , 10331-10351.
https://doi.org/10.1039/C6OB01581F

905. Guocan Yu, Run Zhao, Dan Wu, Fuwu Zhang, Li Shao, Jiong Zhou, Jie Yang, Guping Tang, Xiaoyuan
Chen, Feihe Huang. Pillar[5]arene-based amphiphilic supramolecular brush copolymers: fabrication,
controllable self-assembly and application in self-imaging targeted drug delivery. Polymer Chemistry 2016, 7
(40) , 6178-6188. https://doi.org/10.1039/C6PY01402J

906. Hao Xing, Bingbing Shi. Supramolecular main-chain polycatenanes formed by orthogonal metal ion
coordination and pillar[5]arene-based host–guest interaction. Polymer Chemistry 2016, 7 (40) , 6159-6163.
https://doi.org/10.1039/C6PY01617K

907. Li-Jun Chen, Bo Jiang, Hai-Bo Yang. Transformable nanostructures of cholesteryl-containing rhomboidal
metallacycles through hierarchical self-assembly. Organic Chemistry Frontiers 2016, 3 (5) , 579-587.
https://doi.org/10.1039/C6QO00017G

908. Gilbert Yu, Yuji Suzaki, Kohtaro Osakada. Ferrocene-containing [1]-, [2]-, [3]- and [4]rotaxanes synthesized
from a common precursor. RSC Advances 2016, 6 (47) , 41369-41375.
https://doi.org/10.1039/C6RA05688A

909. J. E. M. Lewis, R. J. Bordoli, M. Denis, C. J. Fletcher, M. Galli, E. A. Neal, E. M. Rochette, S. M. Goldup.


High yielding synthesis of 2,2′-bipyridine macrocycles, versatile intermediates in the synthesis of rotaxanes.
Chemical Science 2016, 7 (5) , 3154-3161. https://doi.org/10.1039/C6SC00011H

910. Guocan Yu, Dan Wu, Yang Li, Zhihua Zhang, Li Shao, Jiong Zhou, Qinglian Hu, Guping Tang, Feihe
Huang. A pillar[5]arene-based [2]rotaxane lights up mitochondria. Chemical Science 2016, 7 (5) , 3017-3024.
https://doi.org/10.1039/C6SC00036C

911. Timothy A. Barendt, Sean W. Robinson, Paul D. Beer. Superior anion induced shuttling behaviour
exhibited by a halogen bonding two station rotaxane. Chemical Science 2016, 7 (8) , 5171-5180.
https://doi.org/10.1039/C6SC00783J

912. Lauren R. Holloway, Hannah H. McGarraugh, Michael C. Young, Watit Sontising, Gregory J. O. Beran,
Richard J. Hooley. Structural switching in self-assembled metal–ligand helicate complexes via ligand-centered
reactions. Chemical Science 2016, 7 (7) , 4423-4427. https://doi.org/10.1039/C6SC01038E
913. Igor Goychuk. Molecular machines operating on the nanoscale: from classical to quantum. Beilstein
Journal of Nanotechnology 2016, 7 , 328-350. https://doi.org/10.3762/bjnano.7.31   

914. Marzia Galli, James E. M. Lewis, Stephen M. Goldup. A Stimuli-Responsive Rotaxane-Gold Catalyst:
Regulation of Activity and Diastereoselectivity. Angewandte Chemie 2015, 127 (46) , 13749-13753.
https://doi.org/10.1002/ange.201505464

915. Marzia Galli, James E. M. Lewis, Stephen M. Goldup. A Stimuli-Responsive Rotaxane-Gold Catalyst:
Regulation of Activity and Diastereoselectivity. Angewandte Chemie International Edition 2015, 54 (46) ,
13545-13549. https://doi.org/10.1002/anie.201505464

916. Bradley D Smith. Smart molecules for imaging, sensing and health (SMITH). Beilstein Journal of Organic
Chemistry 2015, 11 , 2540-2548. https://doi.org/10.3762/bjoc.11.274

Partners

1155 Sixteenth Street N.W.


Washington, DC 20036
Copyright © 2020
American Chemical Society

About Resources and Support & Contact


Information
About ACS Publications Help
ACS & Open Access Journals A-Z Live Chat
ACS Membership Books and Reference FAQ
Advertising Media Kit
Institutional Sales
ACS Publishing Center
Privacy Policy
  
Terms of Use

Connect with ACS Publications

    

También podría gustarte