Está en la página 1de 9

+

October 1996
Volume 85, Number 10

REVIEW ARTICLE

Pharmaceutical Applications of Cyclodextrins. 1. Drug Solubilization and Stabilization


THORSTEINN LOFTSSON*X
AND

MARCUS E. BREWSTER

Received December 29, 1995, from the *Department of Pharmacy, University of Iceland, P.O. Box 7210, IS-127 Reykjavik, Iceland, and Pharmos Corporation, Two Innovation Drive, Alachua, FL 32615 . Final revised manuscript received March 1, 1996 . Accepted for publication March 19, 1996X.
Abstract 0 Cyclodextrins are cyclic oligosaccharides which have recently been recognized as useful pharmaceutical excipients. The molecular structure of these glucose derivatives, which approximates a truncated cone or torus, generates a hydrophilic exterior surface and a nonpolar cavity interior. As such, cyclodextrins can interact with appropriately sized molecules to result in the formation of inclusion complexes. These noncovalent complexes offer a variety of physicochemical advantages over the unmanipulated drugs including the possibility for increased water solubility and solution stability. Further, chemical modification to the parent cyclodextrin can result in an increase in the extent of drug complexation and interaction. In this short review, the effects of substitution on various cyclodextrin properties and the forces involved in the drugcyclodextrin complex formation are discussed. Some general observations are made predicting drug solubilization by cyclodextrins. In addition, methods which are useful in the optimization of complexation efficacy are reviewed. Finally, the stabilizing/destabilizing effects of cyclodextrins on chemically labile drugs are evaluated.
ents. These carbohydrates are mainly used to increase the aqueous solubility, stability, and bioavailability of drugs, but they can also, for example, be used to convert liquid drugs into microcrystalline powders, prevent drug-drug or drugadditive interactions, reduce gastrointestinal or ocular irritation, and reduce or eliminate unpleasant taste and smell. The following is a short review of the effects of cyclodextrins on the solubility and stability of drugs in aqueous solutions with emphasis on the more recent developments. For further information on cyclodextrins and their physicochemical properties the reader is referred to several excellent books and reviews published in recent years.5-13

Structure and Physicochemical Properties


Cyclodextrins are cyclic (R-1,4)-linked oligosaccharides of R-D-glucopyranose containing a relatively hydrophobic central cavity and hydrophilic outer surface. Owing to lack of free rotation about the bonds connecting the glucopyranose units, the cyclodextrins are not perfectly cylindrical molecules but are toroidal or cone shaped. Based on this architecture, the primary hydroxyl groups are located on the narrow side of the torus while the secondary hydroxyl groups are located on the wider edge (Figure 1). The most common cyclodextrins are R-cyclodextrin, -cyclodextrin, and -cyclodextrin, which consist of six, seven, and eight glucopyranose units, respectively. While it is thought that, due to steric factors, cyclodextrins having fewer than six glucopyranose units cannot exist, cyclodextrins containing nine, ten, eleven, twelve, and thirteen glucopyranose units, which are designated -, -, -, -, and -cyclodextrin, respectively, have been reported.14,15 Of these large-ring cyclodextrins only -cyclodextrin has been well characterized.16,17 Chemical and physical properties of the four most common cyclodextrins are given in Table 1. The melting points of R-, -, and -cyclodextrin are between 240 and 265 C, consistent with their stable crystal lattice structure.18 The parent cyclodextrins, in particular -cyclodextrin, have limited aqueous solubility, and their complex formation with

Introduction
Although cyclodextrins are frequently regarded as a new group of pharmaceutical excipients, they have been known for over 100 years.1 The foundations of cyclodextrin chemistry were laid down in the first part of this century2,3 and the first patent on cyclodextrins and their complexes was registered in 1953.4 However, until 1970 only small amounts of cyclodextrins could be produced and high production costs prevented their widespread usage in pharmaceutical formulations. Recent biotechnological advancements have resulted in dramatic improvements in cyclodextrin production, which has lowered their production costs. This has led to the availability of highly purified cyclodextrins and cyclodextrin derivatives which are well suited as pharmaceutical excipiX

Abstract published in Advance ACS Abstracts, May 1, 1996.

1996, American Chemical Society and American Pharmaceutical Association

S0022-3549(95)00534-X CCC: $12.00

Journal of Pharmaceutical Sciences / 1017 Vol. 85, No. 10, October 1996

Figure 1s(a) The chemical structure and (b) the toroidal shape of the -cyclodextrin molecule. Table 1sSome Characteristics of r-, -, -, and -Cyclodextrina R No. of glucopyranose units Molecular weight Central cavity diameter () Water solubility at 25 C (g/100 mL)
a

7 1135 6.06.5 1.85

8 1297 7.58.3 23.2

9 1459 10.311.2 8.19

6 972 4.75.3 14.5

Modified from refs 5 and 17.

lipophilic drugs, and other compounds with limited aqueous solubility, frequently gives rise to B-type phase-solubility diagrams as defined by Higuchi.19 That is, addition of these unmodified cyclodextrins to aqueous drug solutions or drug suspensions often results in precipitation of solid drugcyclodextrin complexes. The aqueous solubility of the parent cyclodextrins is much lower than that of comparable acyclic saccharides, and this could partly be due to relatively strong binding of the cyclodextrin molecules in the crystal state (i.e., relatively high crystal lattice energy). In addition, - and -cyclodextrin form intramolecular hydrogen bonds between secondary OH groups, which detracts from hydrogen bond formation with surrounding water molecules, resulting in less negative heats of hydration.5,17 Thus, intramolecular hydrogen bonding can result in relatively unfavorable enthalpies of solution and low aqueous solubilities. Substitution of any of the hydrogen bond forming hydroxyl groups, even by hydrophobic moieties such as methoxy and ethoxy functions, will result in a dramatic increase in water solubility.5 For example, the aqueous solubility of -cyclodextrin is only 1.85% (w/v) at room temperature but increases with increasing degree of methylation. The highest solubility is obtained when two-thirds of the hydroxyl groups (i.e., 14 of 21) are methylated, but then falls upon more complete alkylation. That is, the permethylated derivative has a solubility that is lower than that of, e.g., heptakis(2,6-O-dimethyl)--cyclodextrin but that is still considerably higher than that of unsubstituted -cyclodextrin.7 Other common cyclodextrin derivatives are formed by other types of alkylation or hydroxyalkylation of the hydroxyl groups.5,20 The main reason for the solubility enhancement in these derivatives is that chemical manipulation frequently transforms the crystalline cyclodextrins into amorphous mixtures of isomeric derivatives. For example, (2-hydroxypropyl)--cyclodextrin is obtained by treating a base-solubilized solution of -cyclodextrin with propylene oxide, resulting in an isomeric system that has an aqueous solubility well in excess of 60% (w/v).21 The number of isomers generated based on random substitution is very large. Statistically, for example, there are about 130 000 possible heptakis(2-O-(hydroxypropyl))--cyclodextrin derivatives, and given that introduction of the 2-hydroxypropyl function also 1018 / Journal of Pharmaceutical Sciences Vol. 85, No. 10, October 1996

introduced an optically active center, the number of total isomers, i.e., geometrical and optical, is even much greater. In reality, the chemical alkylation of cyclodextrins is not totally random, based on relative reactivities of the hydroxy functions in the molecule. The secondary OH groups on the cyclodextrin molecule (i.e., OH-2 and OH-3 on the glucopyranose units) are somewhat more acidic than the primary OH group (i.e., OH-6). Thus, alkylation of OH-6, the least sterically crowded functionality, is favored in strong basic solutions while alkylation of OH-2, the most acidic of the hydroxyl groups but also the most hindered, is favored in a weak basic solution.22 Thus, some degree of regioselectivity is possible. Both the molar substitution, i.e., the average number of alkyl or hydroxyalkyl groups that have been reacted with one glucopyranose unit, and the location of the alkyl or hydroxyalkyl groups on the cyclodextrin molecule will affect the physicochemical properties of the derivatives including their ability to form drug complexes.23 However, theoretical studies have shown that the alkylation and hydroxyalkylation of the cyclodextrins should not introduce significant steric hindrance.24 Some of the commercially available cyclodextrins are listed in Table 2.

Cyclodextrin Complexes
The central cavity of the cyclodextrin molecule is lined with skeletal carbons and ethereal oxygens of the glucose residues. It is therefore lipophilic. The polarity of the cavity has been estimated to be similar to that of aqueous ethanolic solution.5 It provides a lipophilic microenvironment into which suitably sized drug molecules may enter and be included. No covalent bonds are formed or broken during drug-cyclodextrin complex formation, and in aqueous solutions, the complexes are readily dissociated. Free drug molecules are in equilibrium with the molecules bound within the cyclodextrin cavity. Measurements of stability or equilibrium constants (Kc) or the dissociation constants (Kd) of the drug-cyclodextrin complexes are important since this is an index of changes in physicochemical properties of a compound upon inclusion. Most methods for determining the K values are based on titrating changes in the physicochemical properties of the guest molecule, i.e., the drug molecule, with the cyclodextrin and then analyzing the concentration dependencies. Additive properties that can be titrated in this way to provide information on the K values include25 aqueous solubility,19,26-28 chemical reactivity,10,29,30 molar absorptivity and other optical properties (CD, ORD),31-34 phase solubility measurements,35 NMR chemical shifts,23,36 pH-metric methods,37 calorimetric titration,38 freezing point depression,39 and LC chromato-

Table 2sSome Currently Available Cyclodextrins Obtained by Substitution of the OH Groups Located on the Edge of the Cyclodextrin Ringa
Cyclodextrin Derivatives R Methyl Butyl Methyl Ethyl Butyl Hydroxylalkylated: Hydroxyethyl 2-Hydroxypropyl 2-Hydroxybutyl Esterified: Acetyl Propionyl Butyryl Succinyl Benzoyl Palmityl Toluenesulfonyl Esterified and Alkylated: Acetyl methyl Acetyl butyl Branched: Glucosyl Maltosyl Ionic: Carboxymethyl ether Carboxymethyl ethyl Phosphate ester 3-Trimethylammonium-2hydroxypropyl ether Sulfobutyl ether Polymerized: Simple polymers Carboxymethyl Alkylated: Methyl Butyl Pentyl Hydroxyethyl 2-Hydroxypropyl

Table 3sStandard Enthalpy Change (H) and Standard Entropy Change (S) for Several DrugCyclodextrin Complexes Cyclodextrina HP-R-CD -CD -CD -CD -CD -CD Drug Hydrocortisone Phenytoin, un-ionized Phenytoin, ionized Naproxen Adenine arabinoside Adenosine Ibuprofen (pKa 5.2) pH H (kJ/mol) S (J/(mol K)) Ref 7 7 7 7 2 4 5 6 2 3 4 6 5 8 9 1 32 38 21 13 28 21 29 32 29 17 0.2 3.3 17 18 40 39 42 68 18 71 20 55 63 57 20 28 75 73 70 67 21 18 64 53 15 4 3 34 70 69 22 19 62 59 70 166 26 151 6 127 144 134 28 56 194 176 49 46 46 31 32 32 47 47 47 47 47 47 47 47 47 47 47 48 49 49 49 48 48 48 48 48 48 48

2-Hydroxypropyl

-CD

Diazepam (pKa 3.3)

Acetyl Succinyl

Acetyl

-CD
Succinyl

Hydrochlorothiazide (pKa 8.8 and 10.4) Acetylsalicylic acid Acetazolamide 17-Estradiol Hydrocortisone Methyl acetylsalicylate Methyl salicylate Acetylsalicylic acid Methyl acetylsalicylate Acetylsalicylic acid Methyl acetylsalicylate Methyl salicylate

Glucosyl Maltosyl Carboxymethyl ether Phosphate ester

Glucosyl Maltosyl Carboxymethyl ether Phosphate ester

HP--CD HP--CD HP--CD HP--CD HP--CD HP--CD M/DM--CD M/DM--CD HP--CD HP--CD HP--CD

1 1 1 1 1 1 1

a HP-R-CD: (2-hydroxypropyl)-R-cyclodextrin . -CD: -cyclodextrin. HP-CD: (2-hydroxypropyl)--cyclodextrin. M/DM--CD: mixture of maltosyl- and dimaltosyl--cyclodextrin (3:7). HP--CD: (2-hydroxypropyl)--cyclodextrin.

Simple polymers Carboxymethyl

Simple polymers Carboxymethyl

a Since both the number of substitutes and their location will affect the physicochemical properties of the cyclodextrin molecules, such as their aqueous solubility and complexing abilities, each derivative listed should be regarded as a group of closely related cyclodextrin derivatives.

graphic retention times.40 While it is possible to use both guest or host changes to generate equilibrium constants, guest properties are usually most easily assessed. Connors has evaluated the population characteristics of cyclodextrin complex stabilities in aqueous solution.41 The thermodynamic parameters, i.e., the standard free energy change (G), the standard enthalpy change (H), and the standard entropy change (S), can be obtained from the temperature dependence of the stability constant of the cyclodextrin complex.42 The thermodynamic parameters for several series of drugs and other compounds have been determined and analyzed.43-45 The thermodynamic parameters of several other drugs are listed in Table 3. The complex formation is almost always associated with a relatively large negative H and a S that can be either positive or negative. Also, complex formation is largely independent of the chemical properties of the guest (i.e., drug) molecules. The association of binding constants with substrate polarizability suggests that van der Waals forces are important in complex formation.50 Hydrophobic interactions are associated with a slightly positive H and a large positive S; therefore, classical hydrophobic interactions are entropy driven, suggesting that they are not involved with cyclodextrin complexation since, as indicated, these are enthalpically driven processes. Furthermore, for a series of guests there tends to be a linear relationship between enthalpy and entropy, with increasing

enthalpy related to less negative entropy values.43-45,48 This effect, termed compensation, is often correlated with water acting as a driving force in complex formation. The main driving force for complex formation could, therefore, be the release of enthalpy-rich water from the cyclodextrin cavity.47 The water molecules located inside the cavity cannot satisfy their hydrogen-bonding potentials; therefore, they are of higher enthalpy.51 The energy of the system is lowered when these enthalpy-rich water molecules are replaced by suitable guest molecules which are less polar than water. Other mechanisms that are thought to be involved with complex formation have been identified in the case of R-cyclodextrin. In this instance, release of ring strain is thought to be involved with the driving force for compound-cyclodextrin interaction. Hydrated R-cyclodextrin is associated with an internal hydrogen bond to an included water molecule which perturbs the cyclic structure of the macrocycle. Elimination of the included water and the associated hydrogen bond is related to a significant release of steric strain decreasing the system enthalpy.52 In addition, nonclassical hydrophobic effects have been invoked to explain complexation. These nonclassical hydrophobic effects are a composite force in which the classic hydrophobic effects (characterized by large positive S) and van der Waals effects (characterized by negative H and negative S) are operating in the same system. Using adamantanecarboxylates as probes, R-, -, and -cyclodextrins were examined.53 In the case of R-cyclodextrin, experimental data indicated small changes in H and S consistent with little interaction between the bulky probe and the small cavity. In the case of -cyclodextrin, a deep and snug-fitting complex was formed leading to a large negative H and a near zero S. Finally, complexation with -cyclodextrin demonstrated near zero H values and large positive S values consistent with a classical hydrophobic interaction. Evidently, the cavity size of -cyclodextrin was too large to provide for a significant

Journal of Pharmaceutical Sciences / 1019 Vol. 85, No. 10, October 1996

Table 4sEffect of Poly(vinylpyrrolidone) Concentration on the Value of the Apparent Stability Constant (Kc) of Some Drug(2-Hydroxypropyl)--cyclodextrin (1:1) Complexes at Room Temperature (2023 C)a

Kc (M-1)
PVP (% w/v) 0.00 0.10 0.25 0.50
a

Acetazolamide 86.2 95.4 97.0 96.2

Hydrocortisone 1010 1450 c 1190

17-Estradiol 52900 58800 78200 80400

water-soluble polymers to an aqueous complexation medium, followed by heating of the medium in an autoclave, can significantly increase the apparent stability constant of the drug-cyclodextrin complex (Table 4).49,64,65 A somewhat similar effect has been obtained through formation of drughydroxy acid-cyclodextrin ternary complexes or salts with basic drugs.66-68

Drug Solubilization
The most common pharmaceutical application of cyclodextrins is to enhance drug solubility in aqueous solutions. Some of the reports generated on this topic have been reviewed,5-9 and additional data is available from the individual cyclodextrin manufacturers. The solubilizing effects of various cyclodextrins on three different drugs are listed in Table 5. Although prediction of compound solubilization by cyclodextrins continues to be highly empirical, various historical observations permit several general statements. First, the lower the aqueous solubility of the pure drug, the greater the relative solubility enhancement obtained through cyclodextrin complexation. Drugs that possess aqueous solubility in the micromole/liter range generally demonstrate much greater enhancement than drugs possessing solubility in the micromole/liter range or higher. In Table 5, the enhancement factor, i.e., the solubility in the aqueous cyclodextrin solution divided by the solubility in pure water, for paclitaxel, for example, is much larger than the enhancement factors for hydrocortisone and pancratistatin. A similar observation was made when the solubilizing effect of (2-hydroxypropyl)--cyclodextrin on 53 different drugs was investigated.9 Second, cyclodextrin derivatives of lower molar substitution are better solubilizers than the same type of derivatives of higher molar substitution. In Table 5, both randomly methylated - and -cyclodextrins with molar substitution 0.6 provide for better solubilization than the same type of randomly methylated cyclodextrins with molar substitution 1.8. With the exception of R-cyclodextrin, permethylated derivatives (of - and -cyclodextrin) possess a lower complexing potential (lower Kc value) than the parent cyclodextrins.23 Of the commercially available materials, the methylated cyclodextrins with relatively low molar substitution appear to be the most powerful solubilizers. The chain length of the alkyl group, on the other hand, appears to be of less importance.24,70 Third, charged cyclodextrins can be powerful solubilizers, but their solubilizing effect appears to depend on the relative proximity of the charge to the cyclodextrin cavity. The farther away the charge is located, the better the complexing abilities. For example, (2-hydroxy-3(trimethylammonio)propyl)-- and --cyclodextrin possess excellent solubilizing effects while -cyclodextrin sulfate has a relatively low complexation potential (Table 5). Sulfobutyl ether -cyclodextrin, where the anion has been moved away from the cavity by a butyl ether spacer group, is an excellent solubilizer.71 (Carboxymethyl)--cyclodextrin is another interesting anionic cyclodextrin derivative.72 Compared to neutral cyclodextrins, enhanced complexation is frequently observed when the drug and cyclodextrin molecules have opposite charge but decreased complexation is observed if they carry same type of charge. For example, (2-hydroxy-3(trimethylammonio)propyl)--cyclodextrin is an excellent solubilizer for many acidic drugs capable of forming anions. Another finding is that while many ionizable drugs are able to form cyclodextrin complexes, the stability constant of the complex is much larger for the un-ionized than for the ionized form. For example, both the un-ionized and the cationic (i.e., the protonated) form of chlorpromazine give rise to 1:1 complexes with -cyclodextrin but the stability constant for the un-ionized form is 4 times larger than for the cationic

From ref 49. b Poly(vinylpyrrolidone). c Not determined.

contribution by van der Waals-type interactions. These various explanations show that there is no simple construct to describe the driving force for complexation. Although release of enthalpy-rich water molecules from the cyclodextrin cavity is probably an important driving force for drugcyclodextrin complex formation, other forces may be important. These forces include van der Waals interactions,34,54 hydrogen bonding,55,56 hydrophobic interactions,34,57 release of ring strain in the cyclodextrin molecule,56 and changes in solvent-surface tensions.58 Methods of preparing drug-cyclodextrin complexes have been reviewed.25 In the solution phase, the procedure is generally as follows: an excess amount of the drug is added to an aqueous cyclodextrin solution, and the suspension is agitated for up to 1 week at the desired temperature. The suspension is then filtered or centrifuged to form a clear drugcyclodextrin complex solution. For preparation of solid formulations of the drug-cyclodextrin complex, the water is removed from the aqueous drug-cyclodextrin complex solution by evaporation or sublimation. It is sometimes possible to shorten this process by formation of supersaturated solutions through sonication followed by precipitation at the desired temperature. In some cases, the efficiency of complexation is not very high, and therefore, relatively large amounts of cyclodextrins must be used to complex small amounts of drug. To add to this difficulty, vehicle additives, osmolality modifiers, and pH adjustments commonly used in drug formulations, such as sodium chloride, buffer salts, surfactants, preservatives, and organic solvents, very often reduce the efficiency. For example, in aqueous solutions, ethanol and propylene glycol at low concentrations have been shown to reduce the cyclodextrin complexation of testosterone and ibuprofen by acting as competing guest molecules while at higher concentrations they can reduce complexation through a manipulation of solvent dielectric constant.48,59 Likewise, non-ionic surfactants have been shown to reduce cyclodextrin complexation of diazepam60 and preservatives to reduce the cyclodextrin complexation of various steroids.61 On the other hand, additives such as ethanol can promote complex formation in the solid or semisolid state.62 Un-ionized drugs usually form a more stable cyclodextrin complex than their ionic counterparts; thus, the complexation efficiency of basic drugs can be enhanced by addition of ammonia to the aqueous complexation media. For example, solubilization of pancratistatin with (hydroxypropyl)-cyclodextrins was optimized upon addition of ammonium hydroxide.63 Freeze-drying of the solutions removed ammonia, resulting in ammonia-free solid complex preparations which dissolved rapidly to form clear supersaturated pancratistatin solutions. The resulting solutions were stable for a few hours, time sufficient for potential use in parenteral preparations. Finally, enhanced complexation can be obtained by formation of ternary complexes (or cocomplexes) between a drug molecule, a cyclodextrin molecule, and a third component. For instance, addition of a small amount of various 1020 / Journal of Pharmaceutical Sciences Vol. 85, No. 10, October 1996

Table 5sSolubility of Drugs in Different Cyclodextrin Solutions at Room Temperature Drug Hydrocortisone (MW 362) Cyclodextrina None Glucosyl-R-CD Maltosyl-R-CD HP--CD MS 0.6 HE--CD RM--CD MS 0.6 RM--CD MS 1.8 HTMAP--CD MS 0.5 CM--CD MS 0.6 Glucosyl--CD Maltosyl--CD RM--CD MS 0.6 RM--CD MS 1.8 None -CD Dimaltosyl--CD HE--CD HP--CD DM--CD -CD HP--CD None HTMAP--CD MS 1.4 S--CD Na-salt MS 2.3 CM--CD Na-salt MS 0.6 HP--CD MS 0.5 Maltosyl--CD MS 0.14 DM--CD MS 2.0 HE--CD -CD HTMAP--CD MS 0.3 HP--CD MS 0.7 TM--CD MS 3.0 Concnb (% w/v) 10 10 10 10 10 10 10 10 10 10 10 10 1.5 50 50 50 50 15 50 10 10 10 10 10 10 10 10 10 10 10 Solubility (mM) 0.993 7.45 11.3 33.7 48.3 72.2 50.8 30.3 44.6 46.9 28.7 58.8 38.6 4 10-4 0.005 0.115 0.914 0.856 39.6 0.020 0.080 0.16 0.86 0.28 0.83 1.0 0.95 1.2 0.83 0.80 0.49 0.83 0.49 Enhancementc Factor 7.50 11.4 33.9 48.6 72.7 51.2 30.1 44.9 47.2 28.9 55.2 38.9 13 288 2285 2140 99.000 50 200 5.4 1.8 5.2 6.3 5.9 7.5 5.2 5.0 3.1 5.2 3.1 Ref 49 49 49 49 49 27 27 27 27 49 49 27 27 69 69 69 69 69 69 69 69 63 63 63 63 63 63 63 63 63 63 63 63

Paclitaxel (Taxol, MW 854)d

Pancratistatin (MW 325)

a -CD: -cyclodextrin. HP--CD: (2-hydroxypropyl)--cyclodextrin. HE--CD: (hydroxyethyl)--cyclodextrin. RM--CD: randomly methylated -cyclodextrin. HTMAP--CD: (2-hydroxy-3-(trimethylammonio)propyl)--cyclodextrin. CM--CD: (carboxymethyl)--cyclodextrin. Glucosyl--CD: glucosyl--cyclodextrin. Maltosyl-CD: maltosyl--cyclodextrin. DM--CD: 2,6-O-dimethyl--cyclodextrin. S--CD: -cyclodextrin sulfate. -CD: -cyclodextrin. RM--CD: randomly methylated -cyclodextrin. HP--CD: (2-hydroxypropyl)--cyclodextrin. HTMAP--CD: (2-hydroxy-3-(trimethylammonio)propyl)--cyclodextrin. TM--CD: trimethyl -cyclodextrin. MS: molar substitution (i.e., the average number of OH groups on each glucose repeat unit that have been substituted). Na salt: sodium salt. b Concentration of the aqueous cyclodextrin solution. c The solubility in the aqueous cyclodextrin solution divided by the solubility in water. d pH 7.4

form.37 The Kc for the phenytoin--cyclodextrin complex is over 3 times larger for the un-ionized form than for the anionic form.46 However, it is frequently possible to enhance cyclodextrin solubilization of ionizable drugs by appropriate pH adjustments. Thus, the solubilizing effects of both (2-hydroxypropyl)--cyclodextrin and dimethyl--cyclodextrin on dihydroergotamine mesylate have been found to increase with decreasing pH (i.e., formation of the cationic form). Both the saturation solubility and the slopes of the phase-solubility diagrams increase with decreasing pH.73 Similar results have been reported for the complexation of phenytoin with -cyclodextrin46 and for the complexation of indomethacin,74 prazepam, acetazolamide, and sulfamethoxazole75 with (2hydroxypropyl)--cyclodextrin. As mentioned before, it is also possible to enhance complexation and, thus, the solubilizing effect of cyclodextrins by addition of polymers or hydroxy acids to the cyclodextrin solutions. It has been shown that polymers, such as watersoluble cellulose derivatives and other rheological agents, can form complexes with cyclodextrins and that such complexes possess physicochemical properties different from those of individual cyclodextrin molecules.49,76 In aqueous solutions water-soluble polymers increase the solubilizing effect of cyclodextrins on various hydrophobic drugs by increasing the apparent stability constants of the drug-cyclodextrin complexes. For example, the solubilizing effect of 10% (w/v) (2hydroxypropyl)--cyclodextrin solution on a series of drugs and other compounds was increased from 12 to 129% when 0.25%

(w/v) poly(vinylpyrrolidone) was added to the aqueous cyclodextrin solution.49 Water-soluble polymers are also capable of increasing aqueous solubilities of the parent cyclodextrins without decreasing their complexing abilities, thus making them more feasible as pharmaceutical excipients. Likewise, addition of hydroxy acids, such as citric, malic, or tartaric acid, can enhance the solubilizing effect of cyclodextrins through formation of super complexes or salts.67 It is frequently possible to obtain even larger solubilization enhancement by applying several methods simultaneously. For instance, prazepam is a benzodiazepine with a pKa of about 3. (2Hydroxypropyl)--cyclodextrin has a solubilizing effect on both the un-ionized and the ionized form of the drug, and as expected, hydroxypropyl methylcellulose has a synergistic effect on the solubilization. However, the synergistic effect was more pronounced for the ionized form (Figure 2).75 Finally, pharmaceutical formulations should contain as small an amount of cyclodextrin as possible since excess cyclodextrin can reduce, e.g., drug bioavailability and preservative efficacy. Drug solubility should be determined in the final formulation and under normal production conditions to determine if too much, or too little, cyclodextrin is being used.

Effect on Drug Stability


The effects of cyclodextrins on the chemical stability of drugs is another useful property of these excipients and has been extensively examined in the literature.10 Cyclodextrin

Journal of Pharmaceutical Sciences / 1021 Vol. 85, No. 10, October 1996

observed first-order rate constant (kobs) is the weight average of the two rate constants:

d[D]t ) -kobs[D]t dt

and

kobs ) koff + kc(1 - ff)

where ff is the fraction of free drug in solution, or

ff )

1 1 + Kc[CD]

A Lineweaver-Burk type of equation25 can be obtained by further manipulations of the above equations:

1 1 1 1 ) + ko - kobs Kc(ko - kc) [CD] ko - kc


A plot of 1/(ko - kobs) versus 1/[CD] will give rise to a straight line (i.e. if the assumption of a 1:1 complex is correct) with a y-intercept equal to 1/(ko - kc) and a slope equal to 1/Kc(ko kc), from which the values of kc and Kc can be derived. At low concentration most drug-cyclodextrin complexes are of 1:1 stoichiometry. Even complexes which are of higher order stoichiometry at high cyclodextrin and/or drug concentration form 1:1 complexes at lower concentration. The stoichiometry (i.e., the guest:host molar ratio) will, however, affect the stabilizing/destabilizing effect of the complexation.77-80 Thus, at relatively high concentrations the antiallergic drug, tranilast, forms a 2:1 (guest:host) complex with -cyclodextrin which accelerates the drug degradation (dimerization) by approximately 5500-fold. With increasing -cyclodextrin concentrations, 1:1 and 1:2 complexes are formed resulting in a decreased rate of dimerization.77 The rate of dimerization was, for example, 19 300 times slower within the 1:2 complex than outside it. Similar observations were made when the degradation rate of some pilocarpine prodrugs was studied in aqueous (2-hydroxypropyl)--cyclodextrin solutions.80 As mentioned before, the enthalpy of the system decreases during complex formation, resulting in increased complexation (i.e., increased Kc value) when the temperature is lowered. Thus, better overall stabilization is frequently obtained at low temperatures than at high temperatures. Drug-cyclodextrin complexation can be regarded as molecular encapsulation, i.e., encapsulation of drug at the molecular level. The cyclodextrin molecule shields, at least partly, the drug molecule from attack by various reactive molecules. That is, the cyclodextrin can insulate a labile compound from a potentially corrosive environment and, in this way, reduce or even prevent drug hydrolysis, oxidation, steric rearrangement, racemization, and other forms of isomerization, polymerization, and even enzymatic decomposition of drugs. For example, the anticancer drug doxorubicin is unstable in aqueous solutions undergoing acid-catalyzed glycosidic bond hydrolysis, A-ring aromatization subsequent to cleavage of the 9-hydroxymethyl ketone function, and photodecomposition.81-85 Fluorescence, absorbance, circular dichroism, and NMR studies have all indicated complex formation between doxorubicin and both - and -cyclodextrins.86 The A-ring of the anthraquinonic nucleus of the drug is located inside the cyclodextrin cavity, resulting in a significantly slower rate of degradation (Table 6). It has been shown that doxorubicin forms a stronger complex with -cyclodextrin than with -cyclodextrin86 and that, in general, -cyclodextrins are more effective stabilizers of doxorubicin than, for example, (2-hydroxypropyl)--cyclodextrin.89 It appears, however, that complexation of a closely related drug, daunorubicin, with methylated -cyclodextrin (MCD) offers

Figure 2sThe effect of ionization and hydroxypropyl methylcellulose (HPMC) on the (2-hydroxypropyl)--cyclodextrin (HPCD) solubilization of prazepam (pKa 3) in aqueous buffer solutions.

interaction with labile compounds can result in several outcomes: cyclodextrins can retard degradation, can have no effect on reactivity, or can accelerate drug degradation. In some ways, therefore, cyclodextrins can mimic enzymatic catalysis or inhibition. Similarities include substrate binding prior to reaction, saturation kinetics, competitive inhibition, and stereospecific interactions. Due to saturation kinetics, the observed first-order rate constants for reaction (kobs) asymptotically approach a maximum (catalysis) or minimum (inhibition) value with increasing cyclodextrin concentration. The concentration dependence of the kobs can be used to derive both Kc and kc, the rate constant for the reaction of the included compound, by methods analogous to MichaelisMenten analysis. For the formation of a 1:1 complex, the following equilibrium applies:

where ko represents the observed first-order rate constant for the degradation of the free drug (D), CD is the cyclodextrin, and D-CD is the drug-cyclodextrin 1:1 complex. In the above equilibrium,

Kc )

[D-CD] [D]([CD] - [D-CD]) or, if [CD] . [D], Kc ) [D-CD] [D][CD]

where [CD] is the total cyclodextrin concentration in the solution. The degree of stabilization/destabilization of a drug upon cyclodextrin complexation is dependent on not only the rate of drug degradation within the complex (i.e., the value of kc) but also the fraction of the drug that resides within the complex (which again depends on the value of Kc). The 1022 / Journal of Pharmaceutical Sciences Vol. 85, No. 10, October 1996

Table 6sProposed Structure of the Doxorubicin-Cyclodextrin Complex86 and Stabilization of Doxorubicin and Related Drugs by Cyclodextrin Complexation87-89

Drug Daunorubicin Demethoxydaunorubicin Doxorubicin

pH 1.5 1.5 1.5 1.01 1.84 5.90 7.72 1.5

Temp ( C) 50 50 50 75 75 75 75 50

koa (min-1)
2.16 10-3 2.00 10-3 2.16 10-3 0.17 1.86 10-2 1.23 10-2 5.48 10-2 1.71 10-3

Cyclodextrinb M--CD -CD M--CD HP--CD HP--CD HP--CD HP--CD -CD

kca (min-1)
2.64 10-4 3.72 10-4 5.40 10-4 3.02 10-2 2.10 10-3 4.70 10-3 1.03 10-2 3.36 10-4

ko/kc
8.2 5.4 4.0 5.7 8.9 2.6 5.3 5.1

Kca (M-1)
1960 211 3690 69.8 193 243 132 197

a k represents the observed first-order rate constant for the degradation of the free drug, k represents the observed first-order rate constant for the degradation o c of the drug within the complex, and Kc is the observed stability constant for the complex, assuming 1:1 complex formation. b M--CD: methylated -cyclodextrin. -CD: -cyclodextrin. HP--CD: (2-hydroxypropyl)--cyclodextrin.

better protection, i.e., a larger stability constant (Kc) and a more favorable ko:kc ratio, than -cyclodextrin. Aspirin (acetylsalicylic acid) is a phenolic acetate ester, and thus, it is unstable in aqueous solutions. In acidic buffer solutions (at pH about 1), the ester is hydrolyzed via an AAC2 mechanism whereby it undergoes an acyl-oxy cleavage subsequent to protonation, attack by water molecules, and formation of an unstable tetrahedral intermediate.90 Unionized aspirin forms stable (1:1) inclusion complexes with the various -cyclodextrins. NMR studies have shown that in the complex the benzene ring is located well inside the cavity with the acetyl ester group protruding from cavity. This location of the acetyl ester does not completely prevent its hydrolysis but due to steric hindrance, the hydrolysis was determined to be 4-6 times slower within the complex than outside it (i.e., the ko:kc ratio in Table 7 is between 4 and 6). However, under neutral conditions, where aspirin is in the ionized form, the same cyclodextrins did not affect the observed hydrolytic rate constant. NMR studies indicated that the ionized aspirin does not form complexes with the -cyclodextrins tested. The cyclodextrins did not influence the kinetic behavior (e.g., the order of reaction) or the degradation mechanism, only the rate of reaction.48 Sulfobutyl ether -cyclodextrin, which is an anionic -cyclodextrin derivative, has been shown to be highly effective in improving the chemical stability of the antitumor drug O6benzylguanine. The benzyl moiety of the drug was responsible for the cyclodextrin complex formation resulting in an objective increased shelf-life of an aqueous parenteral O6-benzylguanine formulation.71 The same cyclodextrin derivative has been used to increase the shelf-life (and ocular absorption) of pilocarpine in aqueous eye drop solutions.91 The cyclodextrin stabilization of pilocarpine appeared to be independent of the drug ionization status. Another anionic type cyclodextrin, i.e., O-(carboxymethyl)-O-ethyl--cyclodextrin, has been used to stabilize prostaglandin E1 in a fatty alcohol propylene

Table 7sProposed Structure of the Aspirin-Cyclodextrin Complex and Stabilization of Aspirin by Cyclodextrin Complexation48

Drug

pH

Temp (C) ko (min-1) Cyclodextrina 65 4.76 10-3 H--CD M/DM--CD HP--CD

kc (min-1)
1.11 10-3 8.25 10-4 1.18 10-3

ko/kc Kc (M-1)
4.3 5.8 4.0 76.0 53.3 23.2

Aspirin Ca. 1

a HP--CD: (2-hydroxypropyl)--cyclodextrin. M/DM--CD: mixture of maltosyland dimaltosyl--cyclodextrin (3:7). HP--CD: (2-hydroxypropyl)--cyclodextrin.

glycol ointment.92 Dihydroergotamine nasal spray has been used as an acute treatment of migraine. However, dihydroergotamine, the free base, has both limited aqueous solubility and stability. Cyclodextrins, such as (2-hydroxypropyl)-cyclodextrin, have been used to solubilize the drug in aqueous solutions and to stabilize it during autoclaving.73 Degradation kinetics in the solid state are, in general, more complicated and they progress more slowly than in aqueous solutions. Consequently, there are fewer reports on the effects of cyclodextrins on the solid-state decomposition of drugs.

Journal of Pharmaceutical Sciences / 1023 Vol. 85, No. 10, October 1996

Solid drug-cyclodextrin complexes are more water-soluble than the pure lipophilic drugs, and thus, moisture-promoted solid-state decomposition should be accelerated by formation of water-soluble drug-cyclodextrin complexes. This has been shown to be the case with carmofur, an anticancer agent, where the solid-state rate of degradation was increased upon complexation with -cyclodextrin.93 Under dry conditions cyclodextrin complexation of drugs commonly increases both their chemical (e.g., prostaglandin E194) and physical stability (e.g., nifedipine95,96). Large drug molecules like peptides and proteins can also form cyclodextrin complexes, and frequently the complexation results in both enhanced chemical and physical stability of this type of drug.97 Interestingly, the mechanism of stabilization is qualitatively different than in the case of small molecular weight pharmaceuticals. Thus, maximum benefit is usually obtained at low cyclodextrin concentrations, and the benefits are often only partly concentration dependent. In the case of interleukin-2 (IL-2), for example, (2-hydroxypropyl)--cyclodextrin optimally inhibited aggregation at levels of 0.5% in the dosage form.98 The effects of cyclodextrins on improved conformational stability may be related to excipient interaction with hydrophobic aromatic moieties on the protein molecules. It is also possible that cyclodextrins can stabilize proteins similarly fashion to other polyalcoholic compounds and carbohydrates such as sorbitol.99 On the other hand, cyclodextrins were more effective than their linear saccharide analogues in stabilizing human growth hormone.100 Other examples of interaction of cyclodextrin with peptide systems include the decreased proteolytic degradation of basic fibroblast growth factor by pepsin and R-chymotrypsin by the addition of water-insoluble aluminum salt of -cyclodextrin sulfate,101 and the inhibition of self-association of insulin in aqueous solutions by maltosyl--cyclodextrin complexation.102 Thymopentin, which is a small peptide consisting of five amino acids, has been stabilized in aqueous solutions by complexation with (2-hydroxypropyl)--cyclodextrin.103 Although cyclodextrin complexation of drug molecules usually results in increased drug stability, there are examples of accelerated degradation.10 For example, it has been shown that the specific-base-catalyzed hydrolysis of the -lactam ring is facilitated by simultaneous hydrogen bonding between two adjacent hydroxyl groups on the glucose residue and the amide carbonyl and -lactam carbonyl groups of the -lactam antibiotics.104 The pH-rate profile for the degradation of cephalothin is characterized by a large pH-independent region from pH about 2-8.105 In this region, the (hydroxypropyl)cyclodextrins had a significant stabilizing effect, but at pH 9.7, where the specific-base catalysis dominates, the same cyclodextrins had a destabilizing effect.106 For aztreonam, specific-base-catalyzed degradation was dominant at pH values greater than 6,105 and in this region of the pH-rate profile, cyclodextrins accelerated the degradation.106 Cyclodextrins have also been shown, under some specific conditions, to destabilize other drugs including aspirin,107 the antiallergic drug tranilast,77 the antiulcer agent 2-(carboxymethoxy)-4,4bis(3-methyl-2-butenyloxy)chalcone,79 and the thromboxane synthetase inhibitor (E)-4-(1-imidazoylmethyl)cinnamic acid.78 Finally, -cyclodextrin has been shown to catalyze the hydrolysis of 2-methoxy-2-phenylacetic acid 4-nitrophenyl ester, but the (R)-enantiomer was always catalyzed to a greater extent than the (S)-enantiomer, displaying the possibility of enantiomer-selective cyclodextrin stabilization/destabilization.108

of prostaglandins and nonsteroidal antiinflammatory agents (piroxicam) have already been introduced to the market. In the United States, a monograph for -cyclodextrins is available in the Pharmacopoeia representing the first such citation for an excipient which is not yet available in a marketed (U.S.) product. A monograph for -cyclodextrin is already in the Japanese Pharmacopoeia, and it will soon appear in the European Pharmacopoeia. It is clear from this and other perspectives that the introduction of oral -cyclodextrin formulations is in the offing. Chemically modified cyclodextrins, especially (2-hydroxypropyl)--cyclodextrin, are also receiving attention. A monograph is in preparation for the U.S. Pharmacopoeia, and versions have already appeared in such compendial sources as the Handbook of Pharmaceutical Excipients.18 (2-Hydroxypropyl)--cyclodextrin has been used as a parenteral (iv) excipient for drugs completing both preliminary and advanced human clinical trials, and the first such formulation is expected to garner regulatory approval within the next year or so. Similarly,aqueous eye drop solutions containing (2-hydroxypropyl)--cyclodextrin are currently undergoing clinical testing. Cyclodextrins are, therefore, proving their usefulness as tools to generate aqueous drug solutions without the use of organic cosolvents, surfactants, or lipids, as formulation adjuncts which increase dissolution rates and oral bioavailability of solid drug complexes, and as materials used to generate safe iv dosage forms intended to provide important pharmacokinetic information or act as potential drug products per se. In addition to the role of the currently applied cyclodextrins, newer derivatives are constantly being developed and reported.

References and Notes


1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. Villiers, A. C. R. Hebd. Seances Acad. Sci. 1891, 112, 536-538. Schardinger, F. Z. Unters. Nahf. Genussm. 1903, 6, 865-880. Schardinger, F. Z. Wien. Klin. Wochenschr. 1903, 16, 486-474. Freundenberg, K.; Cramer, F.; Plieninger, H. Inclusion Compounds of Physiologically Active Organic Compounds. German Patent No. 895,769, 1953. Fro mming, K.-H.; Szejtli, J. Cyclodextrins in Pharmacy; Kluwer Acad. Publ.: Dordrecht, 1994. New Trends in Cyclodextrins and Derivatives; Duche ne, D., Ed.; Editions de Sante : Paris, 1991. Szejtli, J. Cyclodextrin Technology; Kluwer Acad. Publ.: Dordrecht, 1988. Cyclodextrins and their Industrial Uses; Duche ne, D., Ed.; Editions de Sante : Paris, 1987. Loftsson, T.; Brewster, M. E.; Derendorf, H.; Bodor, N. Pharm. Ztg. Wiss. 1991, 4/136, 5-10. Loftsson, T. Drug Stab. 1995, 1, 22-33. Szente, L. In Stability Testing in the EC, Japan and the USA. Scientific and Regulatory Requirements; Grimm, W., Krummen, K., Eds.; Wissenschaftliche Verlagsgesellschaft: Stuttgart, 1993; pp 225-244. van Doorne, H. Eur. J. Pharm. Biopharm. 1993, 39, 133-139. Duche ne, D.; Wouessidjewe, D. Chimicaoggi 1993, Jan/Feb, 1724. French, D. Adv. Carbohydr. Chem. 1957, 12, 189-260. French, D.; Pulley, A. O.; Effenberger, J. A.; Rougvie, M. A.; Abdullah, M. Arch. Biochem. Biophys. 1965, 111, 153-160. Fujiwara, T.; Tanaka, N.; Kobayashi, S. Chem. Lett. 1990, 739742. Miyazawa, I.; Ueda, H.; Nagase, H.; Endo, T.; Kobayashi, S.; Nagai, T. Eur. J. Pharm. Sci. 1995, 3, 153-162. Nash, R. A. In Handbook of Pharmaceutical Excipients; Wade, A., Weller, P. J., Eds.; Am. Pharm. Assoc. & Phar. Press: London, 1994; pp 145-148. Higuchi, T.; Connors, K. A. In Advances in Analytical Chemistry and Instrumentation; Reilly, C. N., Ed.; Wiley-Interscience: New York, 1965; Vol. 4, pp 117-212. Hashimoto, H. In New Trends in Cyclodextrins and Derivatives; Duche ne, D., Ed.; Edition de Sante , Paris, 1991; Chapter 3, pp 97-156. Pitha, J.; Milecki, J.; Fales, H.; Pannell, L.; Uekama, K. Int. J. Pharm. 1986, 29, 73-82. Pitha, J.; Rao, C. T.; Lindberg, B.; Seffers, P. Carbohydr. Res. 1990, 200, 429-435.

12. 13. 14. 15. 16. 17. 18. 19. 20.

Conclusions and Future Directions


Both parent and chemically modified cyclodextrins are rapidly being assimilated into the formulators armamentarium. In Europe and Japan, oral -cyclodextrin complexes 1024 / Journal of Pharmaceutical Sciences Vol. 85, No. 10, October 1996
21. 22.

23. Botsi, A.; Yannakopoulou, K.; Perly, B.; Hajoudis, E. J. Org. Chem. 1995, 60, 4017-4023. 24. Bodor, N.; Huang, M.-J.; Watts, J. D. J. Pharm. Sci. 1995, 84, 330-336. 25. Hirayama, F.; Uekama, K. In Cyclodextrins and their Industrial Uses; Duche ne, D., Ed.; Editions de Sante : Paris, 1987; Chapter 4, pp 131-172. 26. Brewster, M. E.; Estes, K.; Bodor, N. J. Parenter. Sci. Technol. 1989, 43, 262-265. 27. Sigurdardo ttir, A. M.; Loftsson, T. Int. J. Pharm. 1995, 126, 7378. 28. Hussain, M. A.; DiLuccio, R. C.; Maurin, M. B. J. Pharm. Sci. 1993, 82, 77-79. 29. Adachi, H.; Irie, T.; Hirayama, F.; Uekama, K. Chem. Pharm. Bull. 1992, 40, 1586-1591. 30. Lineweaver, H.; Burk, D. J. Am. Chem. Soc. 1934, 56, 658666. 31. Bettinetti, G. P.; Mura, P.; Liguori, A.; Bramanti, G.; Giordano, F. Farmaco 1989, 44, 195-213. 32. Xiang, T.-X.; Anderson, B. D. Int. J. Pharm. 1990, 59, 45-55. 33. Bekers, O.; Beijnen, J. H.; Otagiri, M.; Bult, A.; Underberg, W. J. M. J. Pharm. Biomed. Anal. 1990, 8, 671-674. 34. Nishijo, J.; Nagai, M. J. Pharm. Sci. 1991, 80, 58-62 35. Liu, F.; Kildsig, D. O.; Mitra, A. K. Pharm. Res. 1992, 9, 16711672. 36. Wiese, M.; Cordes, H.-P.; Chi, H.; Seydel, J. K.; Backensfeld, T.; Mu ller, B.W. J. Pharm. Sci. 1991, 80, 153-156. 37. Boudeville, P.; Burgot, J.-L. J. Pharm. Sci. 1995, 84, 1083-1089. 38. Inoue, Y.; Hakushi, T.; Liu, Y.; Tong, L.-H.; Shen, B.-J.; Jin, D.-S. J. Am. Chem. Soc. 1993, 115, 475-481. 39. Suzuki, M.; Ito, K.; Fushimi, C.; Kondo, T. Chem. Pharm. Bull. 1993, 41, 1444-1447. 40. Uekama, K.; Hirayama, F.; Nasu, S.; Matsuo, N.; Irie, T. Chem. Pharm. Bull. 1978, 26, 3477-3484. 41. Connors, K. A. J. Pharm. Sci. 1995, 84, 843-848. 42. Martin, A. Physical Pharmacy, 4th ed.; Lea & Febiger: Philadelphia, 1993; pp 274-277. 43. Rekharsky, M. V.; Schwarz, F. P.; Tewari, Y. B.; Goldberg, R. N.; Tanaka, M.; Yamashoji, Y. J. Phys. Chem. 1994, 98, 40984103. 44. Rekharsky, M. V.; Schwarz, F. P.; Tewari, Y. B.; Goldberg, R. N. J. Phys. Chem. 1994, 98, 10282-10288. 45. Rekharsky, M. V.; Goldberg, R. N.; Schwarz, F. P.; Tewari, Y. B.; Ross, P. D.; Yamashoji, Y.; Inoue, Y. J. Am. Chem. Soc. 1995, 117, 8830-8840. 46. Menard, F. A.; Dedhiya, M. G.; Rhodes, C. T. Pharm. Acta Helv. 1988, 63, 303-308. 47. Menard, F. A.; Dedhiya, M. G.; Rhodes, C. T. Drug Dev. Ind. Pharm. 1990, 16, 91-113. 48. Loftsson, T.; O lafsdo ttir, B. J.; Fridriksdo ttir, H.; Jo nsdo ttir, S. Eur. J. Pharm. Sci. 1993, 1, 95-101. 49. Loftsson, T.; Fridriksdo ttir, H.; Sigurdardo ttir, A. M.; Ueda, H. Int. J. Pharm. 1994, 110, 169-177. 50. Saenger, W. Angew. Chem., Int. Ed. Engl. 1980, 19, 344-362. 51. Bergeron, R. J. In Inclusion Compounds; Atwood, J. L., Davies, J. E. D., MacNicol, D. D., Eds.; Academic Press: London, 1984; Chapter 12, pp 391-443. 52. Saenger, W. In Inclusion Compounds; Atwood, J. L., Davies, J. E. D., MacNicol, D. D., Eds.; Academic Press, London: 1984; Chapter 8, pp 231-259. 53. Cromwell, W.; Bystrom, K.; Eftink, M. J. Phys. Chem. 1985, 89, 326-332. 54. Cramer, F. Angew. Chem. 1956, 68, 115-120. 55. Tong, W.-Q.; Lach, J. L.; Chin, T.-F.; Guillory, J. K. Pharm. Res. 1991, 8, 951-957. 56. Jones, S. P.; Grant, D. J. W.; Hadgraft, J.; Parr, G. D. Acta Pharm. Technol. 1984, 30, 213-223. 57. Tabushi, I.; Kiyosuke, Y.; Sugiomoto, T.; Yamamura, K. J. Am. Chem. Soc. 1978, 100, 916-919. 58. Orstan, A.; Ross, J. B. A. J. Phys. Chem. 1987, 91, 2735-2745. 59. Pitha, J.; Hoshino, T. Int. J. Pharm. 1992, 80, 243-251. 60. Kraus, C.; Mehnert, W.; Fro mming, H.-H. Pharm. Ztg. Wiss. 1991, 136/4, 11-15. 61. Loftsson, T.; Stefa nsdo ttir, O .; Fridriksdo ttir, H.; Gudmundsson, O . Drug Dev. Ind. Pharm. 1992, 18, 1477-1484. 62. Furuta, T.; Yoshii, H.; Miyamoto, A.; Yasunishi, A.; Hirano, H. Supramol. Chem. 1993, 1, 321-325. 63. Torres-Labndeira, J. J.; Davignon, P.; Pitha, J. J. Pharm. Sci. 1990, 80, 384-386. 64. Loftsson, T.; Fridriksdo ttir, H.; Tho risdo ttir, S.; Stefa nsson, E. Int. J. Pharm. 1994, 104, 181-184. 65. Loftsson, T.; Sigurdardo ttir, A. M. Eur. J. Pharm. Sci. 1994, 2, 297-301. 66. Selva, A.; Redenti, E.; Pasini, M.; Ventura, P.; Casetta, B. J. Mass Spectrom. 1995, 30, 219-220.

67. Fenyvesi, E .; Vikmon, M.; Szema n, J.; Szejtli, J.; Ventura, P.; Pasini, M. In The 7th Cyclodextrins Symposium; Osa, T., Ed.; Business Center for Academic Societies Japan: Tokyo, 1994; pp 414-418. 68. Vikmon, M.; Szema n, J.; Szejtli, J.; Pasini, M.; Redenti, E.; Ventura, P. In The 7th Cyclodextrins Symposium; Osa, T., Ed.; Business Center for Academic Societies Japan: Tokyo, 1994; pp 480-483. 69. Sharma, U. S.; Balasubramanian, S. V.; Staubinger, R. M. J. Pharm. Sci. 1995, 84, 1223-1230. 70. Liu, F.; Kildsig, D. O.; Mitra, A. K. Drug Dev. Ind. Pharm. 1992, 18, 1599-1612. 71. Gorecka, B. A.; Sanzgiri, Y. D.; Bindra, D. S.; Stella, V. J. Int. J. Pharm. 1995, 125, 55-61. 72. Cserha ti, T. Int. J. Pharm. 1995, 124, 205-211. 73. Helm, H.; Mu ller, B. W.; Waaler, T. Eur. J. Pharm. Sci. 1995, 3, 195-201. 74. Backensfeld, T.; Mu ller, B. W.; Kolter, K. Int. J. Pharm. 1991, 74, 85-93. 75. Loftsson, T.; Gudmundsdo ttir, T. K.; Fridriksdo ttir, H. Drug Dev. Ind. Pharm. 1996, 22, 403-407. 76. Haldon, T.; Cwiertina, B. Pharmazie 1994, 49, 497-500. 77. Utsuki, T.; Hirayama, F.; Uekama, K. J. Chem. Soc., Perkin Trans. 2 1993, 109-114. 78. Hirayama, F.; Utsuki, T.; Uekama, K.; Yamasaki, M.; Harata, K. J. Pharm. Sci. 1992, 81, 817-822. 79. Utsuki, T.; Imamura, K.; Hirayama, F.; Uekama, K. Eur. J. Pharm. Sci. 1993, 1, 81-87. 80. Jarho, P.; Urtti, A.; Ja rvinen, T. Pharm. Res. 1995, 12, 13711375. 81. Asker, A.; Habib, M. J. Parenter. Sci. Technol. 1988, 42, 153156. 82. Beijnen, J.; van der Houwen, O.; Underberg, W. J. M. Int. J. Pharm. 1986, 32, 123-131. 83. Janssen, M.; Crommelin, D.; Storm, G.; Hulshoff, A. Int. J. Pharm. 1985, 23, 1-11. 84. Beijnen, J.; Wiese, G.; Underberg, W. J. M. Pharm. Weekbl. Sci. 1985, 7, 109-116. 85. Wasserman, K.; Bundgaard, H. Int. J. Pharm. 1983, 14, 7378. 86. Husain, N.; Ndou, T. T.; de la Pen a, A. M.; Warner, I. M. Appl. Spectrosc. 1992, 46, 652-658. 87. Suenaga, A.; Bekers, O.; Beijnen, J. H.; Underberg, W. J. M.; Tanimoto, T.; Koizumi, K.; Otagiri, M. Int. J. Pharm. 1992, 82, 29-37. 88. Bekers, O.; Beijnen, J.; Bramel, E.; Otagiri, M.; Underberg, W. J. M. Pharm. Weekbl. Sci. 1988, 10, 207-212. 89. Brewster, M. E.; Loftsson, T.; Estes, K. S.; Lin, J. L.; Fridriksdo ttir, H.; Bodor, N. Int. J. Pharm. 1992, 79, 289-299. 90. Jones, R. A. Y. Physical and Mechanistic Organic Chemistry; Cambridge Univ. Press: London, 1979; pp 227-235. 91. Ja rvinen, K.; Ja rvinen, T.; Thompson, D. O.; Stella, V. J. Curr. Eye Res. 1994, 13, 897-905. 92. Adachi, H.; Irie, T.; Hirayama, F.; Uekama, K. Chem. Pharm. Bull. 1992, 40, 1586-1591. 93. Kikuchi, M.; Hirayama, F.; Uekama, K. Int. J. Pharm. 1987, 38, 191-198. 94. Yamamoto, M.; Hirayama, F.; Uekama, K. Chem. Pharm. Bull. 1992, 40, 747-751. 95. Uekama, K.; Ikegami, K., Wang, Z.; Horiuchi, Y.; Hirayama, F. J. Pharm. Pharmacol. 1992, 44, 73-78. 96. Hirayama, F.; Wang, Z.; Uekama, K. Pharm. Res. 1994, 11, 1766-1770. 97. Brewster, M. E.; Simpkins, J. W.; Hora, M. S.; Stern, W. C.; Bodor, N. J. Parenter. Sci. Technol., 1989, 43, 231-240. 98. Brewster, M. E.; Simkins, J.; Hora, M.; Bodor, N. Pharm. Res. 1991, 8, 792-795. 99. Manning, M. C.; Patel, K.; Borchardt, R. T. Pharm. Res. 1989, 6, 903-918. 100. Hagenlocher, M.; Pearlman, R. Pharm. Res. 1989, 6, S30. 101. Fukunaga, K.; Hijikata, S.; Ishimura, K.; Sonoda, R.; Irie, T.; Uekama, K. J. Pharm. Pharmacol. 1994, 46, 168-171. 102. Tokihiro, K.; Irie, T.; Uekama, K.; Pitha, J. Pharm. Sci. 1995, 1, 49-53. 103. Brown, N. D.; Butler, D. L.; Chiang, P. K. J. Pharm. Pharmacol. 1993, 45, 666-667. 104. Fujiwara, H.; Kawashima, S.; Yamada, Y. Chem. Pharm. Bull. 1985, 33, 5458-5463. 105. Connors, K. A.; Amidon, K. L.; Stella, V. J. Chemical Stability of Pharmaceuticals; Wiley: New York, 1986; pp 250-256, 309314. 106. Loftsson, T.; Jo hannesson, H. R. Pharmazie 1994, 49, 292-293. 107. Choudhury, S.; Mitra, A. K. Pharm Res. 1993, 10, 156-159. 108. Beyrich, T.; Jira, T.; Beyer, C. Chirality 1995, 7, 560-564.

JS950534B

Journal of Pharmaceutical Sciences / 1025 Vol. 85, No. 10, October 1996

También podría gustarte