Está en la página 1de 3

2.

2 Fenómenos de adsorción
Investigaciones anteriores han demostrado que todos los carbonos con una estructura grafítica, independientemente de la porosidad,
adsorbían oro hasta cierto punto (Sibrell, 1991). El oro no es absorbido de manera significativa por el diamante (Miller y Sibrell,
1991). Los factores primarios que determinan la adsorción de oro por carbono parecen ser la estructura grafitica y la superficie
(Sibrell, 1991; Miller y Sibrell, 1991). Como se esperaba, la adsorción del oro aumenta con el aumento de la superficie (Sibrell, 1991;
Stenebråten y otros, 2000). En el caso del material carbonoso natural que se preparó a partir del yacimiento de Goldstrike, se midieron
superficies que variaban de 1,5 a 43,8 m2/g y que correspondían a entre el 2,46% y alrededor del 69% de contenido de carbono para
16 muestras (Stenebråten et al., 2000). Abotsi y Osseo-Asare (1986) trabajando en una muestra de mineral de Prestea midieron 30,2
m2/g para un contenido de carbono del 4,56%. Sibrell (1991) correlacionó el contenido de carbono con la superficie y la constante de capacidad de
adsorción de oro (valor K), mostrando que un material que contenía 12e40% de carbono tenía una superficie que variaba de 25 a 42,7
m2/g y una capacidad de adsorción correspondiente de 330e2200 mg Au/kg de carbono. Estos estudios también mostraron que un aumento en el contenido de carbono del mineral
aumenta las pérdidas de oro (
Sibrell, 1991; Stenebråten y otros, 2000). Sin embargo, la superficie no es el único factor que determina la
capacidad de adsorción.
Sibrell y Miller (1992) sugirieron que algún tipo de enlace químico no específico de sitios se forma en la adsorción de oro, similar
a lo que se ha postulado que ocurre durante la adsorción de oro a partir de una solución de cianuro por carbón negro o carbón
activado comercial. Los resultados de esta investigación sugieren que algunos sitios de adsorción son preferidos sobre otros y que
estos sitios están directamente relacionados con la superficie.
Otros estudios fundamentales han sugerido que el carbono natural es una combinación de estructura amorfa y grafítica ( Afenya,
1991; Adams y Burger, 1998b). Entre las estructuras amorfas y grafíticas, hay un rango de madurez, que se indica por un aumento en
el tamaño de los cristalitos de grafito en el material de carbono. Con la maduración, el contenido de grupos funcionales en el material
orgánico se reduce y la estructura cristalina se vuelve más ordenada. El grafito es el producto de maduración final, con estructuras de
células unitarias hexagonales (anillos de carbono) y espaciado intercalar (espaciado en d) de aproximadamente 150 Å ( Adams, 1989;
Stenebråten y otros, 2000).
Algunos autores coinciden en que el grafito, el carbón activado comercial y el material carbonoso natural maduro están
compuestos de láminas apiladas de anillos de carbono aromático entrecruzados de manera aleatoria. Parece que los microporos de los
carbonos activados son en su mayoría espacios en forma de hendidura de menos de 2 nm entre hojas aromáticas retorcidas. Estas
estructuras de láminas se conocen como microcristalitas o cristalitas (Sibrell, 1991; Adams, 1989; Bansal y otros, 1988; Schmitz y
otros, 2001a,b; Stenebråten y otros, 2000). La madurez del carbono está directamente relacionada con la dimensión de la
microcristalita perpendicular a los planos aromáticos (Lc), e inversamente relacionada con la distancia atómica entre las láminas
aromáticas (d-spacing) (Schmitz et al., 2001a). Se ha detectado que la adsorción de complejos aurocianuros por carbón activo
comercial está inversamente relacionada con el parámetro microcristalino Lc y directamente relacionada con el espaciado d
microcristalino (Schmitz et al., 2001a; Adams, 1989). Adams (1989) reported an Lc value of 6.9e12.3Å for microcrystallites in a
commercial activated carbon. Stenebråten et al. (2000) found values of 30e88 Å for a high preg-robbing ore, while higher Lc values
(117e372 Å) comparable to 150 Å for graphite (Adams, 1993) were established for a low preg-robbing ore.
Using X-ray diffractometry, Adams (1989) determined an average d-spacing of 3.7 Å for commercial activated carbon having a
high affinity for the aurocyanide complex. The same behavior occurs for carbon from the preg-robbing Goldstrike ores according to
Schmitz et al. (2001a) and Stenebråten et al. (2000), who reported a value of 3.5 Å for the d-spacing of a high preg-robbing Goldstrike
ore, while a lower d-spacing value (3.36 Å) similar to graphite (3.349 Å) (Mantell, 1968) was determined for a low preg-robbing
Goldstrike ore.
Stenebråten et al. (2000) and Schmitz et al. (2001a) found a close to linear correlation between microcrystallite d-spacing,
percentage of preg-robbing in parent ore and carbon content for the Goldstrike ore, which indicates that the extent of graphitization has
a strong inverse correlation with gold losses in the cyanidation process; that is, the higher the crystalline maturity, the lower are the
preg-robbing properties in the ore. However, it is evident that some graphitization is necessary and, as discussed next, adsorption
depends on the availability of edge defects in the crystal lattices.
The extensive porosity of activated carbons has a great effect on gold adsorption, particularly those highly microporous carbons
with pore diameters of about 20 Å. Pore size vs. area distribution studies for Carlin carbonaceous material (Sibrell, 1991) showed a
higher pore size for the carbonaceous matter (40 Å) than for activated carbon. Stenebråten et al. (2000) found lower total pore volume
for high preg-robbing samples, ranging from 0.13 to 0.17 cm 3/g in comparison to activated carbons that range from 0.48 to 1.76 cm 3/g
(Adams, 1989). However, low preg-robbing samples showed higher pore volume (0.25e0.31 cm3/g) than the high preg-robbing ores,
which indicates that this is not a good parameter for comparison. Miller and Sibrell (1991) suggested an inverse correlation between
gold adsorption and micropore size for different carbons, varying from microporous (less than 20 Å) for activated carbon to
macroporous (over 500 Å) for graphite, consistent with the results of Stenebråten et al. (2000), that showed an average pore diameter
of 257e314 Å for high preg-robbing ores, significantly lower pore diameters than those for low preg-robbing ores (519e1050 Å).
Wet ball-mill grinding of carbonaceous ore (Stenebråten et al., 2000) showed that the microcrystallite dimensions, i.e., thickness
(Lc) and length (La) were not affected, while average pore size, total pore volume, and BET surface area suffered some variation,
probably due to the breakdown of macroporosity sites.
In general, gold adsorption by natural carbonaceous matter present in an ore is believed to be due to the activated carbon type of
compounds and gold adsorption is similar to that which occurs with man-made activated carbon typically used for gold recovery by
industry. In this regard, gold adsorption by activated carbon from alkaline solution has been studied extensively. It has been proposed
that the combined effect of electrostatic and chemical interaction between the goldecyanide anion and the carbon surface results in the
886 PART | III Case Study Flowsheets

gold adsorption. Another mechanism is the ion-pair adsorption reaction, where gold is adsorbed onto the carbon surface as the neutral
ion pair [Mnþ][Au(CN)2]n.
From a fundamental point of view, activated carbon has a disordered microcrystalline structure, similar to graphite. The random
organization of graphite microcrystals and oxidation during activation creates a porous structure, which accounts for the high surface
area of activated carbon and its notable adsorptive capacity. A significant amount of research has been carried out to explain the
mechanism of gold adsorption by activated carbon, the nature of the adsorption sites, and the role of specific surface functional groups
in gold adsorption (Adams, 1989; McDougall et al., 1980; McDougall and Hancock, 1981; Jones et al., 1989a,b,c; Ibrado and
Fuerstenau, 1992; Sibrell and Miller, 1991, 1992; Lagerge et al., 1997, 1999).
Currently, the most accepted theories regarding the mechanism of gold adsorption include the adsorption of ion pairs
Mnþ[Au(CN)2]n onto active sites of activated carbon. This ion-pair adsorption process is somewhat selective due to the structure of the
aurocyanide anion, which is considered to be less hydrated than other cyanoanions. On this basis, the selective solvent extraction of
gold from alkaline cyanide solutions has been previously explained (Mooiman and Miller, 1984, 1986; Miller et al., 1987). In addition,
some adsorption of the unpaired anion AuðCNÞ2 occurs through electrostatic interactions at highly active sites of activated carbons
that have appropriate polarity (Lagerge et al., 1997, 1999). In terms of the adsorption sites, the most accepted theory suggests that gold
is adsorbed at the edge or defect sites rather than at sites on the basal planes ( Sibrell and Miller, 1991, 1992; Lagerge et al., 1999).
Regarding the role of surface functional groups on adsorption by activated carbon, researchers suggest that the presence of basic
surface functional groups, specifically the pyrone type, favors the adsorption of gold onto activated carbon (Papirer et al., 1987, 1991,
1995; Polonia-León y otros, 1993).
In 1992, Sibrell and Miller clearly showed that adsorption occurs at the edges (defects) of graphite crystals and that adsorption on
the basal plane face of the crystal was insignificant by comparison. This was demonstrated from autoradiographs as shown in Figure
49.2 and Table 49.2.
These results were later confirmed by in situ scanning tunneling micro-scopy in 1998 (Poinen et al., 1998). It is evident that site-
specific adsorption is prevalent in the adsorption of gold by graphitic carbons (graphite, carbon black, and activated carbon), and most
of the favored sites are at edge defects in the graphite crystal structure. Other researchers ( McDougall et al., 1980) have hypothesized
that the adsorption of the goldecyanide complex is due to ion exchange under conditions of low ionic strength and to ion-pair
adsorption at high ionic strengths.
It is likely that both of these mechanisms take place at edge defects in the graphite structure. Ion exchange would take place with
functional groups that would be found at graphitic carbon edges. For adsorption of ion pairs, van der Waals forces would be involved,
and the unsymmetrical distribution of charge at edge defects in the graphite structure could well play a role in the adsorption process.
It may also be that the adsorbed complex is stabilized by attachment to more than one plane on the carbon. Although this would not
be possible on the flat basal-plane surface, it could be achieved at three-dimensional edge defects in the crystal structure. These
cavities created by edges would accommodate or, in effect, solvate the ion pair. This important

phenomenon has been identified for activated carbon (Sibrell and Miller, 1991; Adams et al., 1987b), for solvent extraction (Wan and
Miller, 1990; Miller et al., 1987; Mooiman and Miller, 1991; McDougall et al., 1987; Adams et al., 1987a, 1990 ) and for resin
adsorption reactions (Akser et al., 1986; Adams et al., 1987b).
These graphite edge-sites or defects in activated carbons and in natural carbonaceous matter are expected to form at high
temperatures and a suitable oxidizing potential. For example, defect pits, or nanocorrals, form spontaneously on graphite surfaces, an
example of which is shown in Figure 49.3 for highly organized pyrolytic graphite (HOPG) (Zhu et al., 2001; Paruchuri et al., 2003).
These pits vary from 2 nm to several microns in size and it seems that such cages could easily accommodate the aurocyanide ion-pair
adsorption reaction.
More recent research results have demonstrated that the preg-robbing capacities of refractory gold ores depend not just on the
graphitic structure of carbonaceous matter and its content, because the preg-robbing capacity varies widely for a given concentration
of graphitic carbon in the ore (Helm et al., 2009). These results from Raman spectroscopy can be used to predict the extent of preg-
robbing and suggest that the degree of disorder in the graphite structure accounts for the variation in preg-robbing activity and these
results support the suggestion that gold adsorption from alkaline cyanide solution occurs at defect sites in the graphitic structure as
discussed by Sibrell and Miller (1992).
It also appears that preg-robbing by carbonaceous matter (CM) extends beyond adsorption by CM and includes adsorption by
pyrite. Experimental results show that gold adsorption by pyrite in the preg-robbing ore is facilitated by
attached carbon (Tan et al., 2005). Pyrite has a greater capacity to adsorb gold when associated with CM. Based on electrochemical
studies, it seems that the CM reduces the rate of gold oxidation, and dissolution, whereas the rate of pyrite oxidation is increased ( van
Deventer et al., 2005).
Finally, it is interesting to note that Dimov et al. (2003) studied the speciation of 16 different gold compounds involving halogen
elements (Cl, Br, I) as well as cyanide, thiocyanate and thiosulfate using several surface-sensitive microbeam techniques, such as TOF-
SIMS (time-of-flight secondary-ion mass spectrometry), TOF-LIMS (time-of-flight laserionization mass spectrometry), and XPS (X-
ray photoelectron spectroscopy). The following step in this research was to study the behavior of the 16 gold compounds after being
adsorbed by activated carbon. They found that complex gold halides when adsorbed on carbon are reduced to Auo, which agrees with
the work of Hiskey and others (Hiskey et al., 1990; Hiskey and Qi, 1991, 1993). Gold cyanide coincided with AuðCNÞ2 speciation,
thiocyanate is reduced to AuðSCNÞ2 and thiosulfate is not adsorbed onto carbon to any significant extent.
Preg-Robbing Gold Ores Chapter | 49 887

Carbonaceous preg-robbing tailings were quantified using positive-ion TOF-LIMS. Results confirmed that the increase in chloride
concentrations in solution from 50 to 250 mg/L increased by five times the concentration of gold adsorbed by the carbonaceous
material, thus increasing gold losses as discussed by Simmons et al. (1998).

También podría gustarte