Sei sulla pagina 1di 14

REVIEW

Preparation and Properties of Resorcinol Formaldehyde Organic and Carbon Gels**


By Shaheen A. Al-Muhtaseb and James A. Ritter*
A brief overview on the preparation and properties of resorcinolformaldehyde organic and carbon gels reveals very interesting features about their structural and performance characteristics. The resulting nanostructure was very sensitive to the various synthesis and processing conditions. This leads to a remarkable potential for designing and tailoring these materials to fit specific applications. Based on step-bystep comparisons of the published studies, approximate generalizations on the specific roles the synthesis and processing conditions play on the final properties are provided. Overall, resorcinolformaldehyde organic gels undergo two main stages during synthesis. The first stage is associated with the preparation of the sol mixture, and the subsequent gelation and curing of the gel. The second stage is associated with the drying of the wet gel. The most important factors that affect the properties of the organic gel during the first stage are the catalyst concentration, the initial gel pH, and the concentration of the solids in the sol. The most important factors that affect the properties of the organic gel during the second stage are the drying procedure (e.g., super- or subcritical drying), and the difference between the surface tensions of the solvent before and after drying. The corresponding resorcinolformaldehyde carbon gels are produced from the organic gels during a third stage, which is associated with carbonization or activation. Depending on the conditions, carbonization and activation both impact the structural and performance characteristics significantly.

1. Introduction
Resorcinolformaldehyde (RF) solgels have been receiving considerable attention in the literature over the past decade or so.[163] Pekala and co-workers[10,23,26,49] appear to have been the first to synthesize RF organic solgels according to a hydrolysiscondensation reaction mechanism that is analogous to the solgel synthesis of inorganic oxides.[6466] Since these initial studies, numerous articles have appeared in the literature that describe not only the various synthesis and processing conditions that can be used to produce organic and carbon aerogels and xerogels, but also how these conditions affect the final structure of these gels. These and other articles have also reported on the uniqueness of the physical, chemical and electrochemical properties of the organic aerogels and xerogels. Hence, much of the literature has been concerned with trying to understand how the synthesis and processing condi-

[*] Prof. J. A. Ritter, Dr. S. A. Al-Muhtaseb Department of Chemical Engineering Swearingen Engineering Center University of South Carolina Columbia, SC 29208 (USA) E-mail: ritter@engr.sc.edu,

[**] Support by the U.S. Army Research Office under Grant No. DAAH0496-1-0421 is greatly appreciated.

tions affect the final nanostructure of the RF gel and then to relate it to the macroscopic performance of the material. The particularly important and useful properties include high porosities (>80 %), surface areas (4001200 m2 g1), and pore volumes, the magnitudes of which depend markedly on the synthesis and processing conditions. As an example, it has been reported that certain RF organic aerogels exhibit extremely low thermal conductivities, which was attributed to its ultra-porous structure.[5] As another example, a number of studies have shown how the nanoporous structure of RF carbon aerogels and xerogels controls the performance of the material as an electrode in electrochemical double-layer capacitors (EDLCs).[30] Many other studies have also been done that link the nanostructure of an RF gel to a particular macroscopic property. Therefore, the objective of this Review article is to give an up-to-date and comprehensive overview of the growing literature on RF organic and carbon aerogels and xerogels. This review should serve as a starting point for those interested in these types of materials, and it should provide insight on how to adjust the synthesis and processing conditions to tailor the nanostructure of an RF organic or carbon gel for a specific application. In Section 2, the synthesis and processing conditions and corresponding properties are reviewed. The significant trends are summarized in Tables 1 to 3. In Section 3, a brief summary of the trends is given; and to gain an apprecia-

Adv. Mater. 2003, 15, No. 2, January 16

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

0935-9648/03/0201-0101 $ 17.50+.50/0

101

S. A. Al-Muhtaseb, J. A. Ritter/ResorcinolFormaldehyde Organic and Carbon Gels

REVIEW

tion of the vast range of possible conditions, a list of the synthesis and processing conditions utilized in each study is given in Table 4. Due to space limitations, the authors have restricted this Review to unmodified RF organic and carbon gels; for example, this article does not include an in-depth review on the doping of RF gels with metals or metal oxides; readers interested in this topic are referred elsewhere.[2,6771]

the properties of RF gels are summarized in Tables 1 to 3. The following sections address the most noteworthy variations of the synthesis and processing procedures, along with the resulting impacts from each variation on the properties of the corresponding gels.

2.1. Materials and Initial Solution Recipes

2. Synthesis, Processing, and Properties


RF carbon solgels have been produced with different variations of essentially the same recipe. This recipe can be summarized as follows. First, resorcinol (R) and formaldehyde (F) are mixed at the appropriate molar ratio in the presence of a basic[10] (or, in very few cases, an acidic)[9] catalyst. Then the solution is heated in a closed container to a predetermined temperature for a sufficient period of time to form a stable crosslinked gel. The gel may then be washed (or not) with a suitable organic solvent to exchange the aqueous solvent. Next, the wet gel is dried either supercritically with carbon dioxide or subcritically in air or nitrogen, which produces an organic aerogel or xerogel, respectively. To produce a carbon aerogel or xerogel, the dried gel is carbonized usually in nitrogen to form the highly porous carbon network.[10] The dried gel may also be activated, e.g., with CO2, either during or following carbonization. Other activation methods may also be used, but few have been tried with RF carbon aerogels and xerogels. The effects of the different processing conditions on

Resorcinol (1,3-dihydroxybenzene, C6H4(OH)2) is a phenolic tri-functional compound, which is capable of adding formaldehyde (HCHO) in the 2-, 4-, and/or 6- positions in its aromatic ring. Varying the concentration ratios of the different reactants has a profound affect on the resulting properties of the gels, as summarized in Table 1. The stoichiometric R/F molar ratio of 1:2 is the most commonly used ratio in the literature.[10] Nevertheless, using excess F results in a dilution effect, which increases the particle size near the gelation limit.[33] Another major factor that can result in this dilution effect is the reduction of the density of the reactants by increasing the amount of solvent. The solvent can be either distilled (and preferably, deionized) water (W) or an organic solvent (e.g., acetone, methanol, ethanol, n-propanol, or isopropanol).[9,17] The final gels produced with water as the solvent are named hydrogels or aquagels, and those produced with organic solvents are called lyogels (including the alcogels, which are based on alcoholic solvents). Overall, the density of the reactants in the initial solution has a considerable effect on the final density of the RF gel.[33]

James A. Ritter received his Ph.D. in Chemical Engineering from the University of Buffalo in 1989. After spending four years with the Westinghouse Savannah River Technology Center, in 1993 he joined the faculty in the Department of Chemical Engineering at the University of South Carolina. His major research interests lie in three areas: nanostructured materials for hydrogen storage, magnetic nanoparticles and supports, electrodes, and adsorbents; magnetic field-enhanced processes for separations, targeted drug delivery, and manipulation of matter at the nanoscale; and cyclic adsorption processes for gas separation and purification, and hydrogen storage.

Shaheen A. Al-Muhtaseb received his Ph.D. in Chemical Engineering from the University of South Carolina in 2001. He is currently a Postdoctoral Research Associate with Professor Harry J. Ploehn. His major research interests include the design of nanostructured materials for separation and energy transport and storage applications, characterization of nanostructured materials, adsorption processes and vapor-adsorbed phase equilibria, extraction processes and liquidliquid equilibria, absorption processes and vaporliquid equilibria, equations of state and adsorption isotherm models, and prediction of derived thermodynamic properties for multiphase systems.

102

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

0935-9648/03/0201-0102 $ 17.50+.50/0

Adv. Mater. 2003, 15, No. 2, January 16

S. A. Al-Muhtaseb, J. A. Ritter/ResorcinolFormaldehyde Organic and Carbon Gels

REVIEW

Table 1. Stage 1: Preparation of initial solution, gelation, and curing: Effects of various factors on the resulting properties. Factor Decreasing reactant concentrations (equivalent to reducing R/F, R/W, or R/C ratios) Effect Smaller particles and pore sizes Less compaction (less voids) of gel structure Increase surface areas of xerogels Either reduce or increase pore volumes of xerogels, depending on pH Increase electrochemical capacitance Either increases or decreases lithium ion charge and discharge capacities, Depending on pyrolysis temperature and gel pH At low RF concentrations: small, smooth, fractal aggregates of particles with wide PSDs At high RF concentrations: no fractal aggregates, very narrow PSDs may reduce gelation time High concentrations: polymeric gels (small polymer particles interconnected with large necks, high surface areas, high mechanical strengths), reduces gelation time Low concentrations: colloidal gels (large particles interconnected with narrow necks, low surface areas, low mechanical strengths) Increase surface areas and pore volumes of carbon aerogels Increase electrochemical capacitance of carbon aerogels Insignificant effect on surface area of carbon xerogels Increase pore volume of carbon xerogels at high density of reactants Either increase or decrease electrochemical capacitance of carbon xerogels, depending on concentration of reactants Either increases or decreases lithium-ion charge and discharge capacities, depending on pyrolysis temperature and reactants concentrations Required for improving the crosslinking of polymer particles

Acidic catalyst solutions

Alkaline catalyst solutions

Increasing gel pH

Gelation and curing

Higher reactant densities result in a denser formation of the RF crosslinked clusters. The RF solution density also has a more pronounced effect on the properties of RF carbon (and, hypothetically, organic) xerogels than aerogels.[3] Increasing the density of the reactants in the initial solution causes a decrease in the surface area of RF carbon xerogels, and either a decrease in their total pore volume at low pH or vise versa.[3] It also decreases the electrochemical double layer capacitance of RF carbon xerogels (especially at low pH) and aerogels.[3] Sodium carbonate (Na2CO3) is the most commonly used alkaline catalyst (C) for the polymerization reaction of R with F. This catalyst activates a small portion of R to act as sites for the growth of the monomer particles.[46] However, when the intention is to incorporate certain transition metals (e.g., Pt, Pd, or Ag) in the final structure of the RF carbon gels, various salts of these metals (e.g., [Pt(NH3)4]Cl2, PdCl2, or AgOOC CH3) are used as the catalyst.[2] The addition of these metals increases the meso- and macropore volumes (and results in the highest possible total pore volumes in the case of small concentrations of Pt).[2] In addition to the alkaline catalysts, dilute acidic catalyst solutions (e.g., HClO4[9,39] or HNO3[38]) can be used. In the case of low RF concentrations, this produces small, smooth, fractal aggregates of gel particles[9] with wide pore size distributions (PSDs).[9,38] In contrast, when using an acidic catalyst solution with high RF concentrations, the fractal aggregates are no longer observed and very narrow PSDs (67 nm) are obtained.[9] An example on these effects

of the catalyst type and ratio is shown in Figure 1.[38] In some cases, the use of an acidic catalyst resulted in a reduction in the gelation time,[39] which may indicate a change in the polymerization mechanism. These findings are summarized in Table 1. In the case of Na2CO3, a molar R/C ratio ranging between 50 and 300 is typical; but, in some cases ratios as high as 1500 are used, in which case crosslinked gel microspheres are formed.[19,20] Overall, the final structure and properties of the polymerized gels are mostly determined by the relative amount of C in the sol.[4,22,38] Low R/C ratios results in small polymer particles (~35 nm) that are interconnected with large necks (giving the gel a fibrous appearance);[22,47] this produces higher density gels.[45] In contrast, high R/C ratios results in large polymer particles (16 to 200 nm in diameter) that are connected by narrow necks in a string-of-pearls fashion.[22,23,33,38,47] These two types of gels are commonly described as polymeric and colloidal RF gels, respectively.[47] Overall, the polymeric RF carbon aerogels have a small particle size, high surface area, high compressive modulus (high mechanical strength), and exhibit substantial shrinkage during supercritical drying,[47] whereas the opposite behavior and characteristics are exhibited by the colloidal RF carbon aerogels.[47] The surface area of RF organic aerogels, polymerized with an alkaline catalyst, increases slightly with an increase in the already low R/C ratio, and then it decreases continuously with

Adv. Mater. 2003, 15, No. 2, January 16

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

0935-9648/03/0201-0103 $ 17.50+.50/0

103

S. A. Al-Muhtaseb, J. A. Ritter/ResorcinolFormaldehyde Organic and Carbon Gels

REVIEW

(a)
Pore size is small

First stage of gelation

Formation of interconnected structure

After gelation

(b)
Pore size is large

First stage of gelation

Formation of interconnected Structure

After gelation

Monomer (substituted resorcinol) Catalyst Reacted resorcinol and formaldehyde

Fig. 2. Model of gelation progress with a) high and b) low C/W ratios [34]. (Reproduced with permission from J. Non-Cryst. Solids, Elsevier, 2001.)

Fig. 1. Scanning electron microscopy (SEM) images of RF carbon aerogels synthesized with a) low and b) high resorcinol to alkaline catalyst ratios and c) low resorcinol to acidic catalyst ratio [38]. (Reproduced with permission from J. Non-Cryst. Solids, Elsevier, 2001.)

a further increase in the R/C molar ratio; hence, a maximum surface area results at an R/C of ~50 for the particular conditions.[4] Moreover, the peak radius of the PSD increases when increasing the R/C ratio or decreasing the C/W ratio.[34] A monodispersed pore structure is obtained with either very low R/C or very high R/W ratios, whereas an increase in the R/C ratio or a decrease in the R/W ratio produces polydispersed pore structures.[4] Moreover, the size of the gel particles can be promoted to the micrometer scale when using high R/C ratios, low concentrations of reactants (low R/W ratio), or, equivalently, low C/W ratio as shown in Figure 2,[34] and low gelation temperatures, which necessarily increases the time for gelation.[19] These particles are sometimes referred to as foams or microcellular materials.[19] Moreover, the electrochemical double-layer capacitance increases when increasing the density of the gel[41,47] or reducing the R/C ratio.[41,43] Dilute acids (e.g., HNO3 or HCl) or bases (e.g., NH4OH) are typically used as buffers to control the pH of the initial so-

lution. The initial gelation pH has profound effects on the final properties of RF carbon aerogels[3] and xerogels.[25] The reactants tend to precipitate at very low solution pHs,[49] while the polymerization-condensation reaction is hindered at very high pHs.[25] Therefore, typical pH values are in the approximate range of 5.4 to 7.6.[3,10,25,49] In general, the surface area of RF carbon xerogels has a weak dependence on the initial solution pH in the acidic range,[3,25] but at a pH higher than 7 the surface area diminishes completely.[26] However, the pore volume of the RF carbon xerogels can increase when increasing the pH, but only at high reactants densities in the initial solution.[3] The surface areas and pore volumes of RF carbon aerogels also increase significantly when increasing the pH.[3] The electrochemical double-layer capacitance of RF carbon aerogels is expected to increase when increasing the pH, and that of the RF carbon xerogel to either increase when increasing the pH at high reactants densities or vise versa.[3] When these RF carbon gels are examined for use as the anode in lithium-ion batteries, increasing the gel pH has generally decreased the reversible discharge capacity of the lithium ions, especially from RF carbon aerogels.[57]

2.2. Gelation and Curing The main factor in the gelation step is the catalyzed, endothermic, polycondensation polymerization reaction of the precursors under controlled conditions to form the polymer structure known as the aquagel or alcogel when using water or alcohol, respectively, as the solvent. Overall, the structure and properties of the RF organic and carbon gels depend strongly on this reaction and the conditions at which it proceeds. A summary of the effects of the gelation and curing

104

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

0935-9648/03/0201-0104 $ 17.50+.50/0

Adv. Mater. 2003, 15, No. 2, January 16

S. A. Al-Muhtaseb, J. A. Ritter/ResorcinolFormaldehyde Organic and Carbon Gels

REVIEW

conditions on the final properties of the gels is given in Table 1. The major reactions between R and F include an addition reaction to form hydroxymethyl derivatives (CH2OH), and then a condensation reaction of the hydroxymethyl derivatives to form methylene (CH2)- and methylene ether (CH2OCH2)-bridged compounds,[25,26,47] as illustrated in Figures 3 and 4. The alkaline C is important for the initial formation of the R anions during the addition reaction. These R anions are much more active than the uncharged R towards the addition of F to form the hydroxymethyl derivatives, which are essential for the subsequent condensation reactions. This multi-step mechanism results in highly crosslinked clusters (7 to 10 nm in diameter)[48] of the polymer. After the completion of this step, the colloidal particles begin to aggregate and assemble into a stiff, interconnected structure locally resembling a string of pearls that fills the original volume of the aqueous solution.[46] To prepare for gelation (polymerization), typically R and F are mixed with the polymerization C and the solvent (see the previous section for guidelines on the appropriate proportions), and stirred for a short period (between 5[34] and 30 min[37]) to form a homogeneous mixture that is commonly called the sol. The addition of any additives to the gels, such as carbon cloths for reinforcing the final gels,[30] is usually done during this initial stage. However, the addition of such additives influences the rate of gelation catalytically, causing a negative effect because of the deposition of the RF sol on the walls of the narrow, hydrophilic, solid boundaries. This deposition dilutes the solution and results in an effect similar to the one
1. Addition Reaction
OH Na2 CO3 OH OH O O OH

discussed previously.[33] It is also recommended (but not necessary) to prepare the sol mixture in an inert atmosphere (e.g., in a glove box with a N2 atmosphere) to avoid the possibility of contaminating the gel with components in air such as CO2, which may change the pH of the solution. Finally, the containers (molds) containing the reaction products are sealed before being removed from the environmentally controlled chamber, or before being heated, to minimize solvent evaporation. Heating is required to provide the required energy or the polymerization reaction. However, the heating requirement can be lowered significantly (from ~80 to ~40 C) if acetone is used as the solvent instead of water.[9] It is even possible to carry out the polymerization reaction with water as the solvent at temperatures as low as 30 C,[30] but with far longer gelation times compared to using an elevated temperature, such as 80 C. Typically, gelation may occur by the second day if the initial solution pH is higher than 7.0, and in several hours if the pH is less than 6.8.[25] However, the gelation process can occur more rapidly with low R/C ratios, high reactants densities or high temperatures.[5] The initial cluster formation and particle growth reactions take only about one hour as indicated by the initial increase and then leveling-off of the dynamic viscosity of the solution.[46] The sample at this stage constitutes a colloidal solution of the monomer particles (clusters). Nevertheless, the actual covalent crosslinking of these particles, which leads to the stiffening of the gel, may start to take place only after many hours and then progress very slowly.[46] However, this reaction can be hastened remarkably by lowering the pH at this intermediate stage to compensate for the depletion of

CH2OH H OH CH2OH

2
H

2. Condensation Reaction
OH CH2OH OH CH2OH CH2+ H ,
+

OH CH2+

OH CH2OH OH CH2 OH CH2OH

+
OH

OH CH2OH OH CH2OH HO CH2 O

OH CH2

OH

CH2 CH2OH

H+,

OH CH2 O CH2 O CH2 OH HO

OH

OH CH2OH OH

OH

CH2

HO

CH2 CH2OH

CH2 OH OH CH2 O CH2OH HO CH2 OH OH CH2 OH

Fig. 3. Molecular presentation of the polymerization mechanism of resorcinol with formaldehyde [25]. (Reproduced with permission from Carbon, Elsevier, 1997.)

Adv. Mater. 2003, 15, No. 2, January 16

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

0935-9648/03/0201-0105 $ 17.50+.50/0

105

S. A. Al-Muhtaseb, J. A. Ritter/ResorcinolFormaldehyde Organic and Carbon Gels

REVIEW

OH HOCH2 OH CH2OH OH CH2OH OH CH2OH CH2OH OH HO CH2OH OH CH2OH

OH

Fig. 4. Cluster growth of resorcinol-formaldehyde monomers [46]. (Reproduced with permission from Lawrence Livermore National Laboratory, 1997.)

protons;[46] this facilitates the condensation reaction as illustrated in Figure 3. Nevertheless, if the pH is lowered too early during the particle growth stage, this may result in an incomplete aggregation reaction,[46] and also require much more acid to complete the reaction.[46] This effect is possibly due to hindrance of the R anion formation during the addition reaction, which limits the extent of RF cluster formation. A temperature program is typically applied during the gelation period, where the sample is kept at around room temperature for about 24 h, and then heated gradually (again over 24 h) until it reaches the final temperature (~90 C).[32] Nevertheless, the direct placement of the sealed sample vial in an oven at the desired gelation temperature also provides satisfactory results.[3] The curing step is important because it allows for the previously formed polymer clusters (particles) to crosslink, which forms the final solid shape of the gel. This step takes about a week at elevated temperatures (80 to 90 C). In fact, prolonged curing times may be necessary to make sure that the crosslinking reactions are sufficiently completed to prevent swelling during the subsequent stage of exchanging the aqueous solution with an organic solution.[46] If it is desired to obtain the gel in a certain monolithic shape, it is important that the colloidal solution be poured into a mold (which can be made of glass, metal, or plastic) of the desired shape before the curing step. Also, it is important to consider that the characteristics of the surface of the mold (especially when small molds are used), as well as any surfaces (e.g., fibers) that are added to the gel, may result in noticeable variations in the final structure of the RF gel. The main factors dictating these effects are associated with the hydrophobicity of the solid surface and the comparative ratio of the mean distance between two adjacent surfaces (i.e., volume/ surface ratio) to the characteristic catalytic penetration depth (~1 to 50 lm).[33] In some cases, skin portions of the gel can form around the hydrophilic surfaces (e.g., glass).[33] A general presentation of these effects of the solid surface and surface dimensions is shown in Figure 5.[33] Upon removal of the crosslinked (cured) gels from the containers, they can optionally be placed in a dilute acid (e.g., a

S/V 0 (hydrophilic)

-1 S/V (hydrophilic)

-1 S/V (hydrophilic)

-1 S/V (hydrophobic)

generalized surface

aerogel deposition

aerogel particles

Fig. 5. Effects of solid surface and the comparative ratio of the mean distances between adjacent surfaces (i.e., volume (V)/surface (S) ratio) to the characteristic catalytic penetration depth (k) on the gel structure and skin formation. [33] (Reproduced with permission from Carbon, Elsevier, 2001.)

5 to 15 % acetic acid solution)[27,40] to increase the crosslinking density by promoting the progress of the condensation reaction of the hydroxymethyl groups. This step is commonly referred to as aging of the gel sample and, surprisingly, it occurs with no shrinkage of the sample.[40] This aging step results in a significant increase in the double layer capacitance of RF carbon aerogels;[41] it also causes a slight decrease in the surface area.[41] This step also increases the mechanical strength of the gel when carried out in an acidic medium as opposed to water (pH 7.0);[40] and when the gel is cured in an acidic medium, the significant changes to the mechanical properties of the gel take place mostly during the first week. Overall, this aging step may be a very useful addition to the synthesis procedure, especially when it comes to handling the fragile aerogel monoliths without breaking them.

2.3. Solvent Exchange Once the final crosslinked gel is formed, it becomes necessary to remove the aqueous solvent that is possibly used as the reaction medium. The different methods used to remove the solvent have dramatic effects on the properties of the RF organic gels, as outlined in Table 2. Typically, the aqueous sol-

106

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

0935-9648/03/0201-0106 $ 17.50+.50/0

Adv. Mater. 2003, 15, No. 2, January 16

S. A. Al-Muhtaseb, J. A. Ritter/ResorcinolFormaldehyde Organic and Carbon Gels

REVIEW

structure. This results in a dense polymer called a xerogel. However, a comparison of the pore size Solvent exchange Necessary for supercritical drying with CO2 or freeze-drying distributions of RF xerogels shows Facilitates replacement with drying media that the shrinkage is most proReduction of surface tensions upon subcritical evaporation nounced for pore sizes >10 .[53] Subcritical drying Production of dried dense polymers called xerogels Also, when a material is dried with Causes significant shrinkage of especially wide pores hot air, the shrinkage generally deEffects can be insignificant if gels were synthesized with high pends on both the drying rate and mechanical strength Increases lithium-ion charge and discharge capacities the thickness of the sample, where the shrinkage increases with an inProduction of dried light polymers called aerogels Supercritical drying with CO2 crease in the thickness of the samInsignificant shrinkage of pore structure ple or an increase in the rate of High surface areas, pore volumes and, sometimes, electrochemical drying.[32] Subsequently, the gel can capacitances Requires high pressures, long times for exchanging solvent with CO2 be dried subcritically (sometimes at atmospheric pressure) without Supercritical drying with acetone Like supercritical drying with CO2, but with lower pressures major changes to its structure Eliminates necessity for exchanging solvent with CO2, shortens when it is mechanically strong processing time significantly enough to withstand capillary presRequires high temperatures to shift acetone to supercritical conditions May cause partial thermal decomposition of dried gels sures.[43] This includes the case when the gel is produced with high Freeze-drying Production of dried light polymers called cryogels R/C ratios[43] (with particle sizes of based on sublimation of frozen solvents 100 nm or more)[20] or when the Cryogels mostly mesoporous Density of solvents must be invariant with freezing gel is reinforced by curing within individual fiber sheets.[33] Accordingly, air-drying is quicker, simpler vent is replaced with an organic one (e.g., methanol, acetone, and less expensive than the supercritical or subcritical CO2 isopropanol, or amyl acetate)[10,49] through a repetitive washextraction and drying procedures.[33] ing procedure. The solvent can also be heated during these washing steps to accelerate its rate of diffusion and hence exchange.[49] The removal of water becomes essential in the 2.4.2. Supercritical Drying case of supercritical drying with CO2 because of the insolubility between CO2 and water. Overall, no shrinkage of the RF Sometimes it is desirable to retain the formed skeleton gel samples is noticed after solvent exchange;[40] therefore, the structure of the wet gel through the subsequent stages of prochoice of the new solvent depends on either its evaporative cessing. Therefore, the previously exchanged (i.e., CO2-soluproperties or its mutual solubility with water and CO2 for ble) solvent can be further exchanged with another solvent of aerogels. It is not clear whether this solvent exchange step is lower surface tension (e.g., CO2) by slowly bleeding the air needed for the production of RF xerogels; however, the effect from a supercritical drying chamber while filling it with liquid of not using it is unknown, but probably noticeable. This step CO2. Then the liquid CO2 (Tc = 31 C, Pc = 7.4 MPa) is shifted may be beneficial before the subcritical drying of RF xerogels to its supercritical state condition (e.g., ~45 C and ~11 MPa) because it reduces the required time and temperature without going through the vaporliquid interface, which minirequired for evaporation, and it may reduce the surface tenmizes the mechanical stresses against the walls of the pores. It sion on the pore walls of the wet gel thereby minimizing is then held at the supercritical condition for four hours or shrinkage. more.[3,45] The gels dried supercritically are named aerogels, and they are known to have outstanding characteristics (e.g., higher surface areas, total pore volumes and sometimes elec2.4. Drying Conditions trochemical double-layer capacitances) and even higher lithium-ion charge and discharge capacities and efficiencies com2.4.1. Subcritical Drying pared to carbon xerogels.[57] However, they are highly sensitive to the synthesis conditions for being tailored within Conventional evaporation of the solvent at atmospheric very broad ranges of properties.[3] Hybrid RF xerogelaeroconditions may cause drastic changes in the surface tension gels can also be formed by a partial evaporation followed by of the solvent upon the formation of the vaporliquid intersupercritical drying.[11] Although, ideally it should not occur, face. This huge difference between the surface tensions of some shrinkage of the gel also occurs during supercritical drythe coexisting vapor and liquid phases results in dramatic ing; however, this shrinkage is minimal for gels with large mechanical stresses that lead to the collapse of the pore pore and particle sizes and vice versa.[19] As another example,
Table 2. Stage 2: Solvent-exchange and drying: Effects of various factors on the resulting properties Factor Effect

Adv. Mater. 2003, 15, No. 2, January 16

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

0935-9648/03/0201-0107 $ 17.50+.50/0

107

S. A. Al-Muhtaseb, J. A. Ritter/ResorcinolFormaldehyde Organic and Carbon Gels

REVIEW

RF aerogels polymerized with acetone as the solvent and an acidic C do not fully preserve their structures during supercritical drying, possibly due to the polymer gel shrinking slightly during the exchange of acetone with CO2.[9] Overall, the main disadvantages of using supercritical CO2 drying are the high pressures and long times (3 to 4 days) required for solvent exchange and drying.[27] To partially circumvent this problem, an alternative procedure to supercritical CO2 drying is the supercritical drying with an organic solvent that is initially used to exchange the aqueous solution in the aquagel (e.g., acetone,[27,51] Tc = 235 C, Pc = 4.7 MPa). This eliminates the necessity of exchanging the organic solvent with CO2. The corresponding process time is shortened significantly.[51] However, although the gel textures obtained after supercritical drying with acetone and CO2 are similar, the shrinkage and density of the RF aerogels from the supercritical drying with acetone are larger than those with CO2.[27] The shrinkage ratio is also very sensitive to the rate of depressurization, where a very high depressurization rate results in very large shrinkage ratios.[51] These deleterious effects of supercritical acetone drying can be moderated significantly by applying N2 at a pressure of ~5 MPa prior to heating the chamber to the supercritical temperature for acetone.[27] The colors of the organic aerogels prepared through the supercritical drying with acetone are also darker than those with CO2 (and sometimes black).[51] This is attributed to the possibility of partial decomposition (or pyrolysis) and condensation of the gel and solvation of the solvent at the high temperatures required for shifting acetone to its supercritical state.[27]

change of the cryogels freeze-dried with tert-butanol is much smaller than that with water,[34] and it causes the formation of more mesopores.[32] The reactant concentrations and time of curing also have significant effects on the micro- and mesoporosities of the cryogels.[31]

2.5. Pyrolysis (Carbonization) The use of a pyrolysis (carbonization) step transforms the organic gel into a relatively pure carbon structure by removing any remaining oxide and hydrogen groups at an elevated temperature. These carbonized RF structures are often called carbon gels. Pyrolysis of RF gels is most commonly carried out in a tube furnace under a constant, moderate (~200 cm3 min1) flow of inert gas, such as N2, argon or He at room temperature for ~ one hour (to replace all the air in the furnace),[34] and eventually at a fixed temperature ranging from 600 to 2100 C.[11] The desired pyrolysis temperature is generally approached gradually using temperature programming. Interestingly, the electrical conductivity of the sample increases significantly during this transformation, as indicated by the common appearance of a broadband, semi-metal-like infrared absorbance.[42] Variations in the pyrolysis conditions cause significant changes in the properties of RF carbon gels. A summary of these changes is shown in Table 3. Higher pyrolysis temperatures tend to reduce the surface areas of both RF carbon aerogels and xerogels; they also reduce the electrochemical double layer capacitance of RF carbon gels.[3,36,47] However, the minor decrease in the surface area with increasing pyrolysis temperature is limited to temperatures above ~600 C, whereas increasing the pyrolysis temperatures while under 600 C increases the surface area.[36] However, RF carbon gels may not be electrically conductive unless carbonized above 750 C.[54] The pyrolysis temperature also has significant effects on the chargedischarge behavior of lithium ions from RF carbon aerogels and xerogels; but the effect also depends on the gel pH and the concentration of reactants.[57] The specific volume (reciprocal density) of RF carbon aerogels also decreases with an increase in the pyrolysis temperature up to ~800[36] or 900 C;[47] it is unaffected at higher temperatures. Figure 6, for example, shows that RF carbon xerogels carbonized at 1200 C are significantly denser than those carbonized at 600 C.[37] The electrochemical double layer capacitance also exhibits a maximum between 800 and 900 C.[48] The specific volumes of RF carbon aerogels are also always higher than these of RF organic aerogels.[36] In agreement with the last trend, the pore volumes of RF carbon xerogels decrease slightly with an increase in the pyrolysis temperature,[37] and the corresponding skeletal density also increases and levels off at just below the skeletal density of graphite after 1050 C.[37] This indicates that these particular RF carbon xerogels have very few closed pores within the skeletal structure. Moreover, the weight loss of RF carbon xerogels due to pyrolysis is practically limited to the period before

2.4.3. Freeze-Drying Cryogels are produced when the liquid solvent is removed by freeze-drying. With this method, the solvent is frozen and then removed by sublimation thus avoiding the formation of a vaporliquid interface.[24,28] Although the freeze-drying method is not generally expected to exhibit vaporliquid interfaces, it is still believed that this method results in some shrinkage of the gel.[32] Therefore, freeze-drying of smaller samples is more practical in producing mesoporous cryogels.[32] Overall, RF carbon cryogels are mostly mesoporous with surface areas >800 m2 g1 and pore volumes >0.55 cm3 g1.[24,28] However, the surface areas and mesopore volumes of the RF carbon cryogels smaller than those of RF carbon aerogels, but micropore formation is easier upon pyrolysis.[24] Before freeze-drying the wet gel, it is very important to exchange the aqueous solution with an alternative liquid (e.g., tert-butanol) that does not exhibit considerable changes in density upon freezing.[24,28] Otherwise, the expansion of the aqueous solution upon freezing not only may result in the destruction of the gel structure, but it may also result in very large pores (macropores) due to ice crystal growth.[31] Rinsing the aquagel twice with tert-butanol, for example, is sufficient for exchanging the aqueous solution.[32] Moreover, the density

108

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

0935-9648/03/0201-0108 $ 17.50+.50/0

Adv. Mater. 2003, 15, No. 2, January 16

S. A. Al-Muhtaseb, J. A. Ritter/ResorcinolFormaldehyde Organic and Carbon Gels

REVIEW

Table 3. Stage 3: Pyrolysis and activation: effects of various factors on the resulting properties. Factor Increasing pyrolysis temperature

Increasing thermal activation time

rises.[42] The first endothermic reaction, which occurs during the first Effect period of heating from room temReduces oxygen content perature to ~250 C, is attributed to Reduces surface area of carbon aerogels and xerogels the release of adsorbed water (as Reduces pore volumes of carbon aerogels and xerogels Increases macropore size distributions indicated by the disappearance of Increases micropore size distributions when very low R/C ratios are used the OH-group infrared bands); this Increases electrochemical capacitance up to ~850 C, thereafter reduces it corresponds to a mass loss of ~3 %. either increases or decreases lithium ion charge and discharge capacities, The other endothermic reactions, depending on gel pH and reactants concentrations which occur during the second periIncreases pore widths, volumes and surface areas od of heating, are attributed to the Increases electrochemical capacitance up to ~3 h, thereafter reduces it release of organic compounds.[42] However, very slight endothermic and exothermic reactions may also occur after these periods during the constant temperature carbonization step.[42] The pyrolysis step reduces the number of macropores (possibly due to shrinkage) and increases the number of micropores and mesopores, which leads to an increase in the surface area[4] of RF carbon aerogels,[2] especially at low pyrolysis temperatures (in accordance with the previously discussed trends). This effect is a result of the burnout of the organic groups, which leads to the creation of new pores or voids in the gel. Overall, shrinkage and mass reduction of 20 and 50 %, respectively, can be expected due to pyrolysis of subcritically dried RF xerogels.[20] Nevertheless, unlike the pore sizes and pore volumes, which both decrease upon pyrolysis, the surface area of the carbon gel may increase after pyrolysis especially with dilute sols, i.e., when low C/W (or, alternatively, R/W) ratios, are used.[34] Shrinkage and mass reduction due to pyrolysis decrease with increasing R/C ratio.[19] Moreover, in addition to the effect on the gelation rate and also pore structure of carbon xerogels, the addition of organic fibers to the gel solution may also minimize shrinkage during pyrolysis.[7]

2.6. Activation
Fig. 6. Transmission electron microscopy images of RF carbon xerogels carbonized at A) 600 C, and B) 1200 C [37]. (Reproduced with permission from Carbon, Elsevier, 2000.)

reaching ~750 C, after which it is almost constant at ~50 % which indicates the initiation of structural changes within the carbon gel without weight loss.[37] Overall, pyrolysis produces almost the same PSD for RF carbon aerogels,[4,53] except at very low R/C ratios.[4] In the latter case, pyrolysis causes the pore size distribution to spread further towards the micropore region.[4] The temperature required for complete graphitization of RF gels may exceed 2000 C;[42] nevertheless, aerogels pyrolyzed at ~1050 C contain separated graphitic structures.[42] Partial graphitic character has also been reported by others based on XRD patterns.[25] The endothermic pyrolysis reactions are most active during the programmed temperature

RF organic aerogels and xerogels can also be activated, either subsequent to or during pyrolysis, with gases such as air, steam, or CO2. In fact, any activation method applied to activated carbons can in principal be applied to RF gels. As an example, thermal activation of RF gels is typically carried out in a flow of air, steam, or CO2 (or dilute CO2 in N2)[37] at 750 to 1000 C (or at the pyrolysis temperature) for 1 7 h.[37,60] After the activation step, pure N2 may be passed through the sample for about two hours to replace the CO2, and allow the sample time to gradually cool to room temperature.[37] Overall, increasing the activation time with CO2 increases the pore volume and the peak pore widths significantly, especially in the narrow pore size (micropore) range[37] and, to a lesser extent, in the mesopore range.[60] As an example, activated RF carbon aerogels with a surface area of 2600 m2 g1 can be obtained after activation with CO2 for seven hours.[60] However, the electrochemical double layer capacitance of RF carbon aerogels may exhibit a maximum

Adv. Mater. 2003, 15, No. 2, January 16

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

0935-9648/03/0201-0109 $ 17.50+.50/0

109

S. A. Al-Muhtaseb, J. A. Ritter/ResorcinolFormaldehyde Organic and Carbon Gels

REVIEW

after ~3 h of activation with CO2.[48] A summary of the effects of the activation time on the properties of RF carbon gels is given in Table 3. Activation with steam is also more effective than with CO2 because it causes a more significant increase in the micro- and mesoporosities.[2] A three-hour activation of an RF carbon xerogel with CO2 increases the total mass loss from 50 to 75 %. This additional weight loss increases the cumulative pore volume and surface area markedly from 1.5 cm3 g1 and 600 m2 g1 to 2.5 cm3 g1 and 1600 m2 g1, respectively.[37] Thermal activation with air also results in a comparable weight loss that depends on the activation time and temperature.[43] Overall, the thermal activation step with CO2 results in a 66 % increase in the electrochemical double-layer capacitance.[41] On the other hand, the size of the RF carbon xerogel particles is hardly affected by activation with CO2;[37] this suggests that the effect of activation and removal of surface functional groups and carbon, which creates more pores and surface area, takes place mainly within the RF carbon xerogel particles.[37] Alternatively, RF gels can be activated chemically by placing the carbon gel in an acidic solution (e.g., 60 % HNO3)[43] for two days.[43] Regardless of the R/C ratio, this step increases the mass of the carbon gel by 10 % and the oxygen content from 4 to 14 %.[43] This indicates the attachment of oxide functional groups to the surface of the carbon gel. Consequently, the chemically activated carbon gel exhibits a significant increase in both the electrochemical double-layer capacitance and the capacity for adsorbing CO2, with only a slight change of the surface area in comparison to thermal activation.[43] Electrochemical activation can also be performed by placing the RF carbon gel in an electrolyte solution (e.g., 1 M H2SO4)[50] and subjecting it to repetitive oxidationreduction cycles. Electrochemical activation results in an overall increase in the electrochemical double layer capacitance. This increase is mostly due to the effect of the reduction steps of the cycles; therefore, it may be beneficial to use reduction times that are longer than those of oxidation. On the other hand, the surface areas of the samples are relatively unaffected by the electrochemical activation procedure. This result may suggest the formation of surface functional groups during electrochemical activation that give rise to a pseudo-capacitance.[54]

2.7. Additional Processing Techniques The reinforcement of RF gels with organic fibers (by adding them to the gel solution) changes the gelation rate and pore structure of the xerogel, reduces the fragility of the final product (which makes it more suitable for different applications) and minimizes the shrinkage of the gel due to pyrolysis.[7] Nevertheless, only selective fibers can be used for this purpose[33] such as Al2O3 (Saffil). Glass fibers, for example, are impractical to employ in such sheets because of their low

melting points (especially during pyrolysis) and their reactivity with carbon.[33] Reinforcement of RF gels with fibers (e.g., carbon fibers) has several advantages for use in electrochemical capacitors[30,48,58,59] or capacitive deionization.[48] The fibers provide high in-plane electrical conductivity while also improving the mechanical strength and flexibility of the carbon gel. Moreover, since the fibers prohibit linear shrinkage of the gels upon pyrolysis, cracks form in the corresponding carbon gels, which gives an enhanced accessibility for the electrolyte solution in the carbon gel. The skin formation which results when the RF solgel deposits on the narrow, hydrophilic, solid surface of the mold, also fosters the use of such materials in proton exchange membrane (PEM) fuel-cell electrodes. The coarse internal structure guarantees the fast release of water in the cathode and the skin serves as a diffusion layer.[33] Moreover, this skin allows for the finer deposition of metal particles (e.g., Pt) on the surface of RF gels.[33] RF carbon aerogel microspheres can also be produced by mixing the initial aqueous reactant solution with an organic solvent (such as cyclohexane or a mineral oil) containing a surfactant (such as SPAN80)[28] and continuously stirring at an elevated temperature (e.g., 60 C)[28] during the whole period of gelation and curing,[14,22] which lasts ~5 to 10 h.[28] Gelation can also be conducted at room temperature, but the resulting organic RF gel microspheres become very dense, almost nonporous, and exhibit very small pores on the surface (just above the molecular diameter of CO2 but less than that of N2).[28] These microspheres are made by dispersing the aqueous gel solution into the heated, surfactant-containing, organic solvent, or mineral oil just after reaching the gel point (i.e., after partial polymerization),[22] but also just before the solution loses its fluidity.[28] This approach is commonly referred to as inverse emulsion polymerization. Continuous agitation is required during the slow addition of the aqueous gel solution to form a dispersed colloidal solution of spherical droplets, where the size of the droplet depends on the rate of agitation and the amount of surfactant.[22] The surface structure of the RF gel microspheres depends strongly on the temperature of the emulsion, and potentially on other factors such as the concentration of the surfactant.[28] At the end of this process, non-sticky gel microspheres are produced with diameters ranging from micrometers to millimeters, depending on the emulsification procedure.[14,22] These gel microspheres can then be dried to form either RF organic aerogel[22] or cryogel[28] microspheres, and carbonized to form the corresponding carbon microspheres. It is noteworthy that no one has made RF carbon xerogel microspheres. Another unique feature of the RF carbon cryogel microspheres is that the largely mesoporous internal structure may be encased in an ultramicroporous layer by varying the pyrolysis temperature, as shown in Figure 7, which imparts molecular sieving properties to the RF carbon gel.[28] Another noteworthy technique for producing high pore volume RF carbon aerogels with plenty of mesopores utilizes silica-particle templates.[6163] When using 3040 wt.-% surfac-

110

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

0935-9648/03/0201-0110 $ 17.50+.50/0

Adv. Mater. 2003, 15, No. 2, January 16

S. A. Al-Muhtaseb, J. A. Ritter/ResorcinolFormaldehyde Organic and Carbon Gels

REVIEW

Fig. 7. SEM images of a) microspheres, b) surface, and c) cross-section of RF carbon cryogel microspheres pyrolyzed at 1300 C (1573 K) [28]. (Reproduced with permission from Carbon, Elsevier, 2002.)

tant-stabilized silica sols with an average particle size of 8 nm prior to the addition of the RF sol components, the resulting RF carbon gels exhibit a very narrow PSD with an average pore size of ~10 nm.[61,62] Increasing the silica-to-RF ratio increases the pore volume significantly, but it also widens the pores and broadens the corresponding PSDs.[62] Overall, maximum surface areas and pore volumes are obtained when using an initial solution pH of 8,[63] wherein the corresponding carbon matrix around the template is essentially nonporous,[63] in agreement with results reported elsewhere.[25] Stabilizing the silica-sol template with a surfactant is essential for obtaining such narrow PSDs.[61] These templated RF carbon gels can be used as adsorbents for bulky pollutants such as dyes and humic substances.[63]

3. Summary
This article presents a brief overview on the fascinating and remarkably flexible properties of RF carbon and organic gels and how these properties are related to the synthesis and processing conditions. Since the properties are uniquely related to the nanostructure, they can be easily tailored by rigidly controlling the conditions. However, slight variations in the conditions may cause drastic variations in the structural characteristics and hence properties. Therefore, the effects of the different conditions must be understood before attempting to

tailor a material to a specific application. This review article should assist in this endeavor with the understanding that the particular properties and trends reported in a particular study may be difficult to generalize because they are very much dependent on the specific conditions utilized. For this reason, a list of all the experimental conditions from each study is provided in Table 4. Unfortunately, some of the studies neglected to provide all the important details; nevertheless, generalizations are offered based on a careful analysis of the reported properties and trends. Be aware, however, that the generalization must not be taken for granted. Overall, the most significant variations of the synthesis and processing conditions can be classified into three stages. These stages are 1) the preparation of the sol mixture, and subsequent gelation and curing, 2) drying of the wet gel, and 3) carbonization or activation of the dry gel. The effects of these factors are summarized below and in Tables 1 to 3. In the first stage, the most critical factors that determine the final characteristics of the RF carbon and organic gels are the catalyst concentration (with respect to the reactants), the gel pH, and the concentration of solids in the sol. Factors that lead to a dilution effect of the sol (e.g., using excess of one reactant or reducing the concentration of solids) increase the particle size, reduce the density of the final polymer remarkably, increase the surface area of xerogels, and either increase or decrease the pore volume of xerogels, depending on the pH. Increasing the R/C ratio increases the particle size and produces a more fibrous structure; it also has a significant effect on the surface area, which may exhibit a maximum depending on the R/C ratio and other conditions. Increasing the R/C ratio also results in an equivalent effect of increasing the concentration of solids in the sol. Increasing the gel pH generally increases the surface area, pore volume, and electrochemical double layer capacitance of RF carbon aerogels, and it either increases or decreases these properties for RF carbon xerogels, depending on the concentration of solids in the solution. The proper control of the gelation and curing conditions is essential for completing the polymerization reactions and associated crosslinking of the polymerized particles. The impact of these factors is most apparent on the resulting mechanical properties of the gels. The main purpose of the second stage is to remove the solvent with minimal alteration of the polymeric structure. Depending on the solvent and the intended drying medium, the solvent may be exchanged with a more compatible solvent that is easier to evaporate, with less surface tension, and more compatible with other drying media such as supercritical CO2. Subcritical drying of the solvent produces RF organic xerogels, which experience significant shrinkage. Alternatively, the solvent can be replaced with one that can be easily transferred to its supercritical conditions (e.g., CO2). Gels dried supercritically (known as RF organic aerogels) exhibit the lowest possible shrinkage. However, supercritical drying is a high-pressure process and it is a relatively long process. Alternative methods to supercritical drying with CO2, which still minimize shrinkage, include supercritical drying with substitute solvents that require lower pressures and the freeze-drying of solvents.

Adv. Mater. 2003, 15, No. 2, January 16

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

0935-9648/03/0201-0111 $ 17.50+.50/0

111

S. A. Al-Muhtaseb, J. A. Ritter/ResorcinolFormaldehyde Organic and Carbon Gels

REVIEW

Table 4. Summary of the synthesis and processing conditions for RF organic and carbon gels [a]. Ref. [2] R/F [mol/mol] 1:1 1:2 1:3 1:2 0.25-1:1 ? 1:2 Catalyst (C) Na2CO3 Pt(NH3)4]Cl2 PdCl2 AgOOCCH3 Na2CO3 Na2CO3 ? Na2CO3 HclO4 R/C [mol/mol] 200:1 800:1 Conc. 13 % pH ? Drying [b] sup. CO2 Pyrolysis 1000 C Activation 900 C stm, 25 min CO2 2 h NA NA NA NA Remarks conc. R/W [mol-%] chosen C refers to transition metal conc. [total wt.-%] conc. [R/W] gels cast in thin sheets (0.20.5 mm) solvent = water with alkaline C, acetone with acidic C conc. [R/W] ?? conc. [units unspecified] hybrid aerogel/ xerogel conc. [w/v] thin slices of gel have reduced shrinkage used organic solvents

[3, 53, 57] [4] [7] [9]

50:1 25750:1 1500:1 50:1 200:1

5% 20% 0.050.25:1 30 wt.-% 15 or 20 wt.-% 5 or 20 wt.-%

5.5 7.0 ? ? ?

sup. CO2/ ambient sup. CO2 ambient sup. CO2

800 C 1050 C 950 C 1000 C ?

[10] [11] [14] [16]

1:2 ? 1:2 ?

Na2CO3 Na2CO3 Na2CO3 Na2CO3 Li2CO3 K2CO3 ? Na2CO3 Na2CO3

200410:1 50400:1 >50:1 ?

25% ? 3070% ?

6.57.4 ? ? ?

sup. CO2 sup. CO2/ ambient sup. CO2 sup./sub. CO2, air sup. solvent sub. acetone ambient

6001200 C 6002100 C NA 1050 C

NA NA NA NA

[17] [19] [21]

? 1:2 1:2

? 1000:1 1500:1 1500:1

? ? 2530%

? ? ?

NA 1050 C 1050 C

NA NA NA

[22]

1:2

Na2CO3

[24]

1:2

Na2CO3

50:1 200:1 300:1 25:1 200:1 50:1 ? 100:1 400:1 ? 100:1 150:1 200:1

<10%

sup. CO2

NA

NA

0.125 0.250 5% 4% 5% 0.25 0.50 ? 1:6155 0.25

5.8

[25] [27] [28]

1:2 1:2 1:2

Na2CO3 Na2CO3 Na2CO3

5.57.5 ? ?

freeze tert-butanol, sup. CO2 ambient sup. acetone/ CO2 freeze tert-butanol sub. acetone freeze water freeze tert-butanol/ water ambient freeze tert-butanol freeze tert-butanol sup. CO2 ambient sup. CO2

1000 C

NA

thin slices of gel have reduced shrinkage conc. [mass/mass] conc. [w/v] inverse emulsion polymerization conc. [w/v]

1050 C NA 7501300 C

NA NA

conc. [w/v] conc. [units unspecified] conc. R/W [g cm3] inverse emulsion polymerization

[30] [31] [32]

1:2 1:2 1:2

Na2CO3 Na2CO3 Na2CO3

? ? ?

800 C NA 1000 C

CO2, 950 C NA NA

conc. [mol/mol] W: (R+F+C) conc. R/W [g cm3]

[33] [34] [35] [36] [37] [38]

1 2:3 1:2 1:2 1:2 1:2 ?

Na2CO3 Na2CO3 Na2CO3 Na2CO3 Na2CO3 Alkaline (Na2CO3?)

1500 50200:1 200:1 200:1 50:1 50:1 300:1 1500:1 50:1 200:1 200:1

2530% 0.1250.500 0.25 ? 5% 15% 5% 30% ?

? 5.8 ? ? 6 ?

1050 C 1000 C NA

NA NA NA

conc. [wt.(RF/total)-%] conc. R/W [g cm3] conc. in R/W [g cm3]

4001050 C NA 6001200 C 1050 C, CO2, 0.53 h 1050 C NA

conc. [wt.-%] conc. [wt.-%]

[39]

Na2CO3 HClO4 Na2CO3

sup. CO2

1050 C

NA

solvent: water (with Na2CO3) or acetone (with HClO4)

[40]

1:2

NA

NA

112

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

0935-9648/03/0201-0112 $ 17.50+.50/0

Adv. Mater. 2003, 15, No. 2, January 16

S. A. Al-Muhtaseb, J. A. Ritter/ResorcinolFormaldehyde Organic and Carbon Gels

REVIEW

Table 4. continued Ref. [41] R/F [mol/mol] 1:2 Catalyst (C) Na2CO3 R/C [mol/mol] 50:1 200:1 2001500:1 50400:1 100200:1 50300 200:1 100210:1 ? Conc. 25% 40% 30, 50% >10% ? 7% 40, 70% 25% ? ? ? 6.57.4 ? ? pH Drying [b] sup. CO2 Pyrolysis 10502500 C Activation Chem., HNO3 Chem., HNO3 NA NA NA CO2, 1050 C NA CO2, 950 C, 2 h/ electrochem. Remarks inverse emulsion polymerization conc. [w/v] conc. in mass ratio conc. [units unspecified] conc. [units unspecified] conc. [w/v] conc. [units unspecified]

[43] [45] [46] [47] [48] [49] [50]

? 1:2 1:2-3 1:2 1:2 1:2 1:2

? Na2CO3 ? Na2CO3 Na2CO3 Na2CO3 Na2CO3

sub. acetone sup. CO2 NA sup. CO2 sup. CO2 sup. CO2 sub. acetone

1050 C 1050 C NA 1050 C 6001100 C NA 800, 1050 C

[51] [52]

? ?

Na2CO3 NaOH

? ?

? 10%

? ?

sup. acetone sup. CO2

? ? conc. [w/v] solvent: ethanol, n-propanol, n-butanol, or n-pentanol conc. [w/v]

[54] [60] [61] [62] [63]

1:2 ? 1:2 1:2 1:2

Na2CO3 Na2CO3 Na2CO3 ? ?

50:1 200:1 500:7 ? ?

5% ? 35% ? ?

? ? 7.3 8 4.39.6

ambient sup. CO2 ? ? ?

6001200 C 1050 C 850 C 850 C 850 C

1050 C, CO2, 3 h 900 C, CO2, 17 h

conc. [mol-% (silica-free)]

[a] ?: Conditions are not given explicitly. [b] sup: supercritical; sub: subcritical.

The pyrolysis (or carbonization) step, i.e., the third stage, transforms the dried gels into relatively pure carbon structures through the thermal decomposition and removal of oxide and hydrogen groups, which depends on the pyrolysis temperature. The transformed carbon structures are referred to as RF carbon gels. Overall, increasing the pyrolysis temperature decreases the surface area of carbon gels (or increases it in the case of low concentrations of solids); but they are always higher than those of RF organic gels. Similarly, the pore volume of RF carbon gels generally decreases when increasing the pyrolysis temperature, possibly due to decreases in the macropore region. Activation of the carbon gels can be performed in conjunction with or after pyrolysis of the RF gel with gases such as air, steam, or CO2 at elevated temperatures (in the same range as those of pyrolysis). Overall, the activation of carbon gels enhances the electrochemical properties significantly and is very effective at increasing the pore volume, pore width (especially in the narrow pore range) and surface area. In conclusion, RF carbon gels are receiving considerable attention towards their use in many commercial applications such as adsorbent materials,[28] electrodes for capacitive deionization of aqueous solutions,[56] ion exchange resins,[55] electrochemical double layer capacitors and supercapacitors,[56] gas diffusion electrodes in PEM fuel cells,[7,33] and anodes in rechargeable lithium ion batteries.[59,72,73] This large variety of potential and existing commercial applications is

largely due to the tunable properties of RF carbon gels, especially the surface area, pore volume, and pore size distribution. The tunable properties of RF carbon gels are easily related to the synthesis and processing conditions, which produce a wide spectrum of nanostructured materials with unique properties. It is anticipated that many new and exciting applications will arise for RF organic and carbon gels as more fundamental research is done to expose their unique and tunable characteristics.
Received: June 15, 2002 Final version: November 15, 2002

[1] [2] [3] [4] [5] [6] [7] [8] [9]

A. Garziella, L. A. Pilato, A. Knop, Phenolic Resins, 2nd ed., Springer, New York 2000. F. J. Maldonado-Hdar, M. A. Ferro-Garca, J. Rivera-Utrilla, C. Moreno-Castilla, Carbon 1999, 37, 1199. E. J. Zanto, S. A. Al-Muhtaseb, J. A. Ritter, Ind. Eng. Chem. Res. 2002, 41, 3151. H. Tamon, H. Ishizaka, M. Mikami, M. Okazaki, Carbon 1997, 35, 791. S. Y. Kim, D. H. Yeo, J. W. Lim, K. Yoo, K. Lee, H. Kim, J. Chem. Eng. Jpn. 2001, 34, 216. University of Wisconsin Space Sciences Engineering Center, Reduced Gravity Aerogel Formation. http://www.cae.wisc.edu/ ~ aerogel/, last accessed in June 2002. M. Glora, M. Wiener, R. Petricevic, H. Prbstle, J. Fricke, J. Non-Cryst. Solids 2001, 285, 283. A. C. Pierre, Introduction to Sol-Gel Processing, Kluwer Academic Publishers, Norwell, MA 1998. S. Berthon, O. Barbieri, F. Ehrbuger-Dolle, E. Geissler, P. Achard, F. Bley, A. Hecht, F. Livet, G. Pajonk, N. Pinto, A. Rigacci, C. Rochas, J. NonCryst. Solids 2001, 285, 154.

Adv. Mater. 2003, 15, No. 2, January 16

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

0935-9648/03/0201-0113 $ 17.50+.50/0

113

S. A. Al-Muhtaseb, J. A. Ritter/ResorcinolFormaldehyde Organic and Carbon Gels

REVIEW

[10] R. W. Pekala, US Patent 4 873 218, 1989. [11] J. L. Kaschmitter, S.T. Mayer, R.W. Pekala, US Patent 5 789 338, 1998. [12] R. Jansen, B. Kessler, J. Wonner, A, Zimmermann, US Patent 5 811 031, 1998. [13] R. S. Mendenhall, G. R. Andrews, J. W. Bruno, D. F. Albert, US Patent 6 077 876, 2000. [14] S. T. Mayer, F. Kong, R. W. Pekala, J. L. Kaschmitter, US Patent 5 508 341, 1996. [15] P. R. Coronado, J. F. Poco, US Patent 6 087 407, 2000. [16] S. T. Mayer, J. L. Kaschmitter, R. W. Pekala, US Patent 5 420 168, 1995. [17] R. S. Mendenhall, G. R. Andrews, J. W. Bruno, D. F. Albert, US Patent 6 090 861, 2000. [18] R. W. Pekala, S. T. Mayer, J. F. Poco, J. L. Kaschmitter, Mater. Res. Soc. Symp. Proc. 1994, 349, 79. [19] R. Saliger, V. Bock, R. Petricevic, T. Tilloston, S. Geis, J. Fricke, J. NonCryst. Solids 1997, 221, 144. [20] V. Bock, U. Fischer, U. Klett, J. Fricke, GDCh Monograph 1995, 3, 471. [21] R. Petricevic, G. Reichenauer, V. Bock, A. Emmerling, J. Fricke, J. NonCryst. Solids 1998, 225, 41. [22] X. Lu, R. Caps, J. Fricke, C. T. Alviso, R. W. Pekala, J. Non-Cryst. Solids 1995, 188, 226. [23] X. Lu, M. C. Arduini-Schuster, J. Kuhn, O. Nilsson, J. Fricke, R. W. Pekala, Science 1992, 225, 971. [24] H. Tamon, H. Ishizaka, T. Yamamoto, T. Suzuki, Carbon 1999, 37, 2049. [25] C. Lin, J. A. Ritter, Carbon 1997, 35, 1271. [26] R. W. Pekala, C. T. Alviso, J. D. Lemay, in Chemical Processing of Advanced Materials (Eds: L. L. Hench, J. K. West), John Wiley and Sons, New York 1992, p. 671. [27] C. Liang, G. Sha, S. Guo, J. Non-Cryst. Solids 2000, 271, 167. [28] Y. Yamamoto, T. Sugimoto, T. Suzuki, S.R. Mukai, H. Tamon, Carbon 2002, 40, 1345. [29] G. C. Ruben, Micron 1998, 29, 359. [30] H. Prbstle, C. Schmitt, J. Frick, J. Power Sources 2002, 105, 87. [31] R. Kocklenberg, B. Mathieu, S. Blacher, R. Pirard, J. P. Pirard, R. Sobry, G. Van den Bossche, J. Non-Cryst. Solids 1998, 225, 8. [32] H. Tamon, H. Ishizaka, T. Yamamoto, T. Suzuki, Carbon 2000, 38, 1099. [33] R. Petricevic, M. Glora, J. Fricke, Carbon 2001, 39, 857. [34] Y. Yamamoto, T. Nishimura, T. Suzuki, H. Tamon, J. Non-Cryst. Solids 2001, 288, 46. [35] Y. Yamamoto, T. Nishimura, T. Suzuki, H. Tamon, Carbon 2001, 39, 2369. [36] S. Q. Zhang, J. Wang, J. Shen, Z. S. Deng, Z. Q. Lai, B. Zhou, S. M. Attia, L. Y. Chen, Nanostruct. Mater. 1999, 11, 375. [37] C. Lin, J. A. Ritter, Carbon 2000, 38, 849. [38] C. I. Merzbacher, S. R. Meier, J. R. Pierce, M. L. Korwin, J. Non-Cryst. Solids 2001, 285, 210. [39] O. Barbieri, F. Ehrburger-Dolle, T. P. Rieker, G. M. Pajonk, N. Pinto, A. V. Rao, J. Non-Cryst. Solids 2001, 285, 109. [40] F. Despetis, K. Barral, K. Kocon, J. Phalippou, J. Sol-Gel Sci. Tech. 2000, 19, 829. [41] R. W. Pekala, S. T. Mayer, J. L. Kaschmitter, F. M. Kong, in Sol-Gel Processing and Applications (Ed: Y. A. Attia), New York 1994, p. 369. [42] J. Kuhn, R. Brandt, H. Mehling, R. Petricevic, J. Fricke, J. Non-Cryst. Solids 1998, 225, 58.

[43] R. Saliger, U. Fischer, C. Herta, J. Fricke, J. Non-Cryst. Solids 1998, 225, 81. [44] F. M. Kong, S. R. Buckley, C. L. Giles Jr., B. L. Haendler, L. M. Hair, S. A. Letts, G. E. Overturf III, C. W. Price, R. C. Cook, Final Report UCRL-LR-106946-DE92-004848, Lawrence Livermore National Laboratory, Livermore, CA 1991. [45] S. T. Mayer, R. W. Pekala, J. L. Kaschmitter, J. Electrochem. Soc. 1993, 140, 446. [46] R. C. Cook, S. A. Letts, G. E. Overturf, III, S. M. Lambert, G. Wilemski, D. Schroen-Carey, Final Report UCRL-LR-105821-97-1, Lawrence Livermore National Laboratory, Livermore, CA 1997. [47] R. W. Pekala, C. T. Alviso, F. M. Kong, S. S. Hulsey, J. Non-Cryst. Solids 1992, 145, 90. [48] R. W. Pekala, J. C. Farmer, C. T. Alviso, T. D. Tran, S. T. Mayer, J. M. Miller, B. Dunn, J. Non-Cryst. Solids 1998, 225, 74. [49] R. W. Pekala, J. Mater. Sci. 1989, 24, 3221. [50] H. Prbstle, R. Saliger, J. Fricke, in Studies in Surface Science and Catalysis (Eds: K. K. Unger, G. Kreysa, J. P. Baselt), Elsevier Science, Amsterdam 2000, p. 371. [51] Qin, S. Guo, Carbon 1999, 37, 1168. [52] Qin, S. Guo, Carbon 2001, 39, 1935. [53] S. A. Al-Muhtaseb, C. E. Holland, J. A. Ritter, unpublished. [54] C. Lin, J. A. Ritter, B. N. Popov, J. Electrochem. Soc. 1999, 146, 3639. [55] M. V. Ernest, Jr., J. P. Bibler, R. D. Whitley, N.-H. L. Wang, Ind. Eng. Chem. Res. 1997, 36, 2775. [56] J. C. Farmer, D. V. Fix, G. V. Mack, R. W. Pekala, J. F. Poco, J. Appl. Electrochem. 1996, 26, 1007. [57] S. A. Al-Muhtaseb, E. J. Zanto, B. N. Popov, J. A. Ritter, unpublished. [58] C. Schmitt, H. Prbstle, J. Fricke, J. of Non-Cryst. Solids 2001, 285, 277. [59] E. Frackowiak, F. Bguin, Carbon 2001, 39, 937. [60] Y. Hanzawa, K. Kaneko, R. W. Pekala, M. S. Dresselhaus, Langmuir 1996, 12, 6167. [61] S. Han, T. Hyeon, Chem. Commun. 1999, 19, 1955. [62] S. Han, T. Hyeon, Carbon 1999, 37, 1645. [63] S. Han, K. Sohn, T. Hyeon, Chem. Mater. 2000, 12, 3337. [64] J. Brinker, G. Scherrer, in Sol-Gel Science: The Physics and Chemistry of Sol-Gel Processing (Eds: C. J. Brinker, G. W. Scherer), Academic Press, New York 1990. [65] Y. A. Attia, Sol-Gel Processing and Applications, Plenum Press, New York 1994. [66] A. C. Pierre, Introduction to Sol-Gel Processing, Kluwer Academic Publishers, Norwell, MA 1998. [67] C. Lin, J. A. Ritter, B. N. Popov, J. Electrochem. Soc. 1999, 146, 3155. [68] J. M. Miller, B. Dunn, Langmuir 1999, 15, 799. [69] E. Bekyarova, K. Kaneko, Langmuir 1999, 15, 7119. [70] C. Moreno-Castilla, F. J. Maldonado-Hdar, F. Carrasco-Marn, E. Rodrguez-Castelln, Langmuir 2002, 18, 2295. [71] E. Bekyarova, K. Kaneko, Adv. Mater. 2000, 12, 1625. [72] S. Escribano, S. Berthon, J. L. Ginoux, P. Achard, Extended abstracts, Eurocarbon '98, Strasbourg 1998, p. 841. [73] B. Huang, Y. Huang, Z. Wang, L. Chen, R. Xue, F. Wang, J. Power Sources 1996, 58, 231.

______________________

114

2003 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

0935-9648/03/0201-0114 $ 17.50+.50/0

Adv. Mater. 2003, 15, No. 2, January 16

Potrebbero piacerti anche