Вы находитесь на странице: 1из 16

Chemical Engineering Science 56 (2001) 37813796

www.elsevier.nl/locate/ces

A novel FTIR method for studying mixed gas adsorption at low concentrations: H2O and CO2 on NaX zeolite and -alumina
Salil U. Rege, Ralph T. Yang
Department of Chemical Engineering, University of Michigan, 3074F H.H. Dow Building, 2300 Hayward Street, Ann Arbor, MI 48109-2136, USA Received 11 October 2000; received in revised form 10 February 2001; accepted 22 February 2001

Abstract The modeling and measurement of mixed gas adsorption at low concentrations (ppm level) is an important part of the design of an adsorptive purication process, but has received little attention so far. This work introduces a new technique for the measurement of the equilibrium adsorbed amounts of a multicomponent mixture of gases= vapors on an adsorbent pellet based on Fourier-transform infrared spectroscopy (FTIR). The technique consists of rst calibrating the infrared absorption peak areas with known adsorbed amounts of the di erent components in the adsorbate mixture using single gases= vapors. By measuring the IR peak areas of the adsorbates on the sorbent in contact with the gas mixture, the actual amounts of the sorbate can be determined. The method is fairly simple, quick, and accurate even at low concentrations provided the adsorbates exhibit strong distinct peaks and the sorbent is at least partially infrared-transparent. The utility of this method is demonstrated by measuring the binary gas adsorption of CO2 and H2 O-vapor at low concentrations on two adsorbents: 13X (NaX) zeolite molecular sieve, and -Al2 O3 . In the low CO2 concentration range, the amount of CO2 adsorbed seemed to be signicantly enhanced in the presence of H2 O in trace amounts. The adsorption data was t to the DoongYang potential theory model, and the ideal (and real) adsorbed solution theory (IAS) of Myers and Prausnitz, with the DubininAstakhov equation as the basis. ? 2001 Elsevier Science Ltd. All rights reserved.
Keywords: FTIR studies; Mixed gas adsorption; Water-vapor; CO2 ; Zeolite; Alumina; Sorption enhancement

1. Introduction In recent years, cyclic adsorption processes have emerged to be popular choices for gas separation and purication applications. For the modeling of these adsorptive processes, multicomponent adsorption equilibrium data over a broad range of concentrations is highly desirable. For most of the bulk gas separation processes in practice, a reasonable amount of multicomponent adsorption data is available in literature (Valenzuela & Myers, 1989; Yang, 1997). However for purication processes involving the removal of one or more trace impurities (e.g., natural gas purication, air pre-purication prior to air separation, etc.) both pure component as well as mixture equilibrium data are extremely scarce. The
Corresponding author. Tel.: +1-734-936-0771; fax: +1-734-763-0459. E-mail address: yang@umich.edu (R. T. Yang).

primary reason for the lack of data for such applications is the absence of a suitable experimental method capable of measuring multicomponent adsorption data at very low adsorbate concentrations with su cient accuracy. A number of experimental methods are available for measuring mixed gas adsorption in general (Yang, 1997). These include the constant-volume method in which a sample is placed in a pre-calibrated volume containing a mixture of gases and allowed to equilibrate over a period of time. By using an appropriate P V T equation of state and by analyzing the gas composition before and after adsorption, the mixed equilibrium data can be obtained from a mass balance. A major disadvantage of this method is the slow attainment of equilibrium due to high di usional resistance in the static volumetric system. The di usional resistance can be overcome by using internal circulation (e.g., Costa, Sotelo, Calleja, & Marron, 1981; Taqvi, Appel, & LeVan, 1999). Another option is to use a dynamic ow system or a di erential adsorption bed

0009-2509/01/$ - see front matter ? 2001 Elsevier Science Ltd. All rights reserved. PII: S 0 0 0 9 - 2 5 0 9 ( 0 1 ) 0 0 0 9 5 - 1

3782

S. U. Rege, R. T. Yang / Chemical Engineering Science 56 (2001) 37813796

(DAB) (Reich, Zeigler, & Rogers, 1980; Chen, Yang, & Uawithya, 1994). In this method, a di erential adsorption bed is subject to a ow of a mixture of gases of known composition. After equilibrium is attained, signied by a breakthrough of the components, the sample bed is isolated, heated, and desorbed into a previously evacuated vessel of known volume. By measuring the pressure and temperature of the desorbed gas and by analyzing its composition, the amounts of the di erent sorbates in the adsorbed phase can be determined. A gravimetric technique involving the measurement of total weight of the adsorbate and using the Gibbs adsorption isotherm for mixtures also exists. However, it has not achieved popularity due to the various simplifying assumptions needed to be made in its analysis. Lastly, chromatographic techniques have been developed recently for measuring mixture adsorption. Despite the speed and simplicity of producing the data, these su er from the di culty of accurately analyzing the data. A review of the above techniques reveals that none of the methods would be well-suited for measuring mixture adsorption when one or more of the components in the gas mixture exists in traces. Since the amounts of the mixed adsorbates at low concentrations (in parts per million or ppm range) would be extremely small, the sensitivity of the volumetric or gravimetric technique would be very low indeed, thus leading to large errors. A high sensitivity at low concentration would be a required characteristic of an experimental method for measurements of such nature. In addition, the capability of indicating the nature of the sorbatesorbate or sorbatesorbent interactions in the mixed adsorption would be a highly desirable feature. With these features in mind, a new method based on Fourier-transform infrared (FTIR) spectroscopy is being proposed in this work. Over the last few decades, FTIR spectroscopy has emerged as a powerful tool for the study of the microscopic behavior of molecules adsorbed in a microporous adsorbent. Infrared studies have been successfully used in studying the conguration and orientation of physisorbed and chemisorbed species on a sorbent surface (e.g., Jacobs, Cauwelaert, & Vansant, 1973; Masuda, Tsutsomi, & Takahashi, 1980; Smudde, Slager, Coe, MacDougall, & Weigel, 1995; Smith et al., 1996; Mawhiney, Rossin, Gerhart, & Yates, 1999). Moreover, these have also been proven to be e ective for quantitative measurements of pure component adsorption isotherms and heats of adsorption (Delaval, Seloudoux, & de Lara, 1986; Jentys, Warecka, Derewinski, & Lercher, 1989). More recently, IR spectroscopy has been proposed as an e cient tool for studying sorption kinetics in zeolites by Karge and Nie en (1991). By measuring absorbance of the respective IR peaks, they were successful in measuring the single and binary sorption kinetics of benzene and ethylbenzene in H-ZSM-5 zeolite as well as studying the phenomenon of counter-di usion. It thus seems only

natural to extend the utility of this technique for measuring mixed gas adsorption. However no evidence of such an application of FTIR spectroscopy could be found in literature by the authors. The present study concerns the measurement of equilibrium adsorption of a binary gas mixture of CO2 and H2 O-vapor in trace levels in an inert gas on -Al2 O3 ( -alumina) and 13X (NaX) zeolite as adsorbents. This is an industrially important system frequently encountered in air purication operations such as that prior to cryogenic air separation (Sircar & Kratz, 1981; Kumar, 1987; Jain, 1993) and in space-vehicle air regeneration adsorptive modules (Dellosso & Winnick, 1969; Chang, Stonesifer, Cusick, & Hart, 1991). The spectral nature of the adsorption of these sorbates also appears to be well-documented in FTIR literature since CO2 and H2 O are considered to be useful probe molecules for studying the structural features of a microporous sorbent. The IR spectra of adsorbed CO2 on NaX (and zeolites in general) has been reported by Bertsch and Habgood (1963), Jacobs et al. (1973), Delaval et al. (1986), and Masuda et al. (1980). The spectra of adsorbed CO2 on alumina surfaces has been discussed by Parkyns (1971), Lee and Condrate (1995) and references therein. The interaction of H2 O with the zeolite framework and the charge-balancing cations present within particularly has received ample attention since it has an important e ect on the catalytic properties of a zeolite. The IR spectra of H2 O both as an adsorbed as well as a structural species has been extensively investigated for various zeolites (Bertsch & Habgood, 1963; Angell & Scha er, 1965; Ward, 1968; Jentys et al., 1989; Parker, Bibby, & Burns, 1993) as well as for alumina (Peri, 1965; Lee & Condrate, 1995). Further useful information regarding the IR spectra of these adsorbed species is also available in the reviews by Little (1966) and Hair (1967). Apart from the mixed gas adsorption experiments, this study also investigates the applicability of di erent multicomponent adsorption isotherm models available in literature to the experimental data. Previous studies on the measurement of pure component isotherms at low pressures for the systems under study (Rege, Yang, & Buzanowski, 2000) revealed that the best t to the experimental data was given by potential theory models such as the DubininAstakhov (DA) equation. It is thus obvious that the multicomponent adsorption model to be considered should incorporate the DA equation as a basis. The latter equation can be extended to gas mixtures by several methods. These include the model of Doong and Yang (1988) which is based upon maximum available pore volume, or the ideal adsorbed solution (IAS) theory of Myers and Prausnitz (1965) corrected for non-ideality by including activity coe cients. Both these models were tested for goodness of t with the experimental binary adsorption data obtained in this work.

S. U. Rege, R. T. Yang / Chemical Engineering Science 56 (2001) 37813796

3783

2. Experimental section 2.1. Materials All gases used in this study were supplied by Metro Welding Supply, Detroit, Michigan. The specications on the gases were as follows: helium (high purity grade, 99.995% purity), nitrogen (prepuried, 99.998% purity), and two CO2 = helium certied gas mixtures (316 ppm CO2 balance He, and 4.8% CO2 balance He). The two types of CO2 = He mixtures provided the exibility of studying the adsorption under two di erent CO2 concentration regimes. The supplied helium gas was further puried by passing it through a xed bed containing a molecular sieve zeolite which was periodically regenerated. A low concentration water vapor stream was made available by saturating a helium stream with water vapor at temperatures close to the freezing point of water. This was done by bubbling helium gas through two gas-wash bottles containing about 400 ml water each maintained at about 4 C by placing them in an ice-water bath. The exact temperature of the water was determined prior to each set of runs for determining the vapor pressure of water. The adsorbent materials used in this study were -Al2 O3 and 13X (NaX) zeolite. These materials were chosen by virtue of their importance as adsorbents in adsorptive air-prepurication processes. The -Al2 O3 (PSD-350, Alcoa Inc.) was supplied in a form of 1= 8 in pellets which were crushed to Mesh 80 size. The 13X zeolite (Linde, lot 945084060002) was supplied as a binderless powder. For the purpose of studying gas adsorption on these materials, the adsorbent powders were pressed into thin circular pellets using a hydraulic press under a pressure of 2000 lbs. The pellets weighed 20 mg each with a diameter of 1.3 cm. The sorbent pel lets were activated by heating to 375 C for 5 h in situ in the infrared sample cell which was simultaneously purged with a moisture-free helium stream. The pellets were heated by using a concentric glass-tube enclosed electric lament connected to a temperature controller. The temperature of the pellet was determined using an in situ thermocouple in contact with the pellet. It should be noted that the low concentration pure component CO2 and water-vapor isotherms on these sorbents were measured in a previous work (Rege et al., 2000) using a volumetric type of apparatus (ASAP 2010, Micromeritics Instrument Corporation). 2.2. FTIR measurements The FTIR measurements were carried out using an Impact 400 FTIR spectrometer (Nicolet Instrument Corporation). The infrared spectra were collected and analyzed using a data acquisition computer which used the OMNIC 4.1b software (Nicolet Corp.). The background and

spectrum measurements were averaged against 100 scans with a resolution of 4:0 cm1 . This number of scans was found to be optimal with respect to signal noise reduction and data acquisition time. The sample cell was tted with KBr windows which enabled measurements in the spectral region of interest. The background spectrum was measured after activating the adsorbent pellet as described previously, followed by cooling it down to room temperature in a puried helium stream. All subsequent sample spectra were ratioed against this background spectrum measured prior to the runs. 2.3. Calibration of absorbance peak areas Before the H2 O= CO2 mixed adsorption measurements can be carried out, it was necessary to calibrate the H2 O and CO2 peak areas with known adsorbed amounts using single gases. This was done by adsorbing pure component H2 O and CO2 on the two adsorbents under study over a range of partial pressures in separate sets of experiments. Since it was not possible to vary the total pressure inside the sample cell, the partial pressure was varied in each run by diluting the streams of the respective adsorbates of measured owrates with a puried helium stream using previously calibrated owmeters. The total ow-rates used in the FTIR measurements ranged from 300 1000 cm3 = min. The areas under the peak corresponding to the physically adsorbed adsorbate species (2357 cm1 for CO2 , and 1650 cm1 for H2 O) were then determined and correlated with the corresponding adsorbed amounts which were measured earlier (Rege et al., 2000). It must be noted that some of the observed peaks in the spectra (especially that of CO2 ) appeared at the same location for the gas phase and adsorbed (physisorbed) phase. This posed a di culty initially, but by performing some preliminary experiments it was found that including the gas-phase absorbance in the calibrations did not a ect the results to a great degree. This di culty may be overcome by using a dual pathway spectrometer with a second cell lled with gas phase only positioned in the second optical pathway for compensation, if such facility is available. The time of equilibration of the gaseous adsorbate with the sorbent was 15 min in all cases. It was earlier veried that this time was greater than the residence time of the gas in the FTIR sample cell. It must be noted however that the actual time for equilibration of sorbates such as CO2 and H2 O at very low (ppm level) concentrations is very large (Bertsch & Habgood, 1963). This is so because most of the adsorption at low partial pressures is chemical in nature and hence its kinetics is very slow. Moreover, a part of the adsorption at low partial pressures may not be entirely reversible in nature. However, the purpose of this study was to study behavior of a binary mixture of adsorbates under predominantly

3784

S. U. Rege, R. T. Yang / Chemical Engineering Science 56 (2001) 37813796

physisorption conditions. Besides the volumetric apparatus which was used to measure adsorption isotherms earlier (Rege et al., 2000) for the calibration of the spectral areas was also programmed to measure adsorption at time intervals approaching the same as the equilibration time allowed in this work. Also, preliminary experiments revealed that there was negligible change in peak areas beyond the equilibration time allowed in this work (15 min) for the systems under study. All the calibration as well as mixture adsorption experiments were performed at room temperature (295 K). 2.4. Mixture adsorption measurements Following the calibration of the adsorbents for H2 O and CO2 single-component adsorption, the mixture equilibrium measurements were carried out using FTIR spectroscopy. It must be noted that the calibration (as explained above) is specic for a particular adsorbent pellet. Any other pellet with a slightly di erent dimension is likely to have a di erent calibration curve. Hence due care was taken to perform the mixture experiments with the very same pellet used in the corresponding calibration experiments. For the H2 O= CO2 mixture experiments, two separate sorbate streams, namely a helium stream saturated with H2 O-vapor in equilibrium with ice-water (about 0.6% v= v), and a pre-diluted CO2 stream (either 300 ppm or 4% v= v), were further diluted with a third helium stream using calibrated owmeters. By manipulating the owrates of the respective streams using the owmeters, adsorption resulting from mixtures of varying concentrations of H2 O-vapor and CO2 could be studied. The range of the concentration of H2 O-vapor in the gas mixtures studied was 0 900 ppm on a volumetric basis (v= v). On the other hand, the mixed adsorption was studied for two different ranges of concentration for CO2 : 0 300 ppm v= v and 1000 40,000 ppm (0.1 4%) v= v. As will be shown later, the mixture equilibrium adsorption has a di erent behavior depending upon the relative concentration range of the adsorbates. In each progressive run of the mixture measurement, the concentration of the H2 O and CO2 was simultaneously increased. Hence it was not required to regenerate the sorbent in between the runs. The advantage of using a ow system over a static system in the IR cell for the calibration as well as mixture measurements was that the external di usional resistance was reduced and faster equilibration time was achieved. 3. Review of models for mixed gas adsorption A large number of adsorption models have been proposed in the last few decades to describe pure component and mixture adsorption (Yang, 1997). However most of the mixed isotherm models remain to be extensively

tested due to a paucity of mixture equilibrium data in literature. In some of the simpler models such as the Langmuir or the T oth models, the extension to multicomponent mixtures is straightforward, while in others the analysis is much more complex. Most of the previous work in mixture adsorption involving water concerns the mixed adsorption of water and organics in carbon. In this respect, the association theory of Talu and Meunier (1996) and its extension to binary mixtures using the virial mixture coe cient (VMC) method of LeVan and coworkers (Appel, LeVan, & Finn, 1998) is noteworthy. Unfortunately both the association theory and the VMC method provide the partial pressure of the mixture components as an implicit function of the sorbate loadings, whereas the vice versa is preferred, especially in simulation programs. Moreover, the presence of a polynomial function in the virial equation tends to give nonphysical oscillatory solutions, which are actually mere artifacts of the method. Nevertheless, the VMC method appears to be a powerful tool for handling highly non-ideal systems (Appel et al., 1998). Before an adsorption model is chosen to t mixture data, it needs to be rst veried that the selected model ts the single component data reasonably well (Richter, Shutz, & Myers, 1989). The single-component isotherm data for the system chosen in the present work (CO2 = H2 O on a zeolite and -alumina) was studied previously by Rege et al. (2000). The data was t to the Langmuir Freundlich (or loading ratio correlation: LRC) equation, T oth equation and the DubininAstakhov (DA) potential theory model. Each of the three models is suited for describing adsorption on a sorbent with heterogeneity of surface energy. However, it was found that although the rst two models provided a good t to data in the high-pressure range, only the DA model provided a good t to the data in the entire pressure range of interest. It was thus natural to choose the DA model as the basis for extension to the binary system under study. Several approaches are available for this purpose, namely the methods proposed by Bering, Serpinsky, and Surinova, (1963), Grant and Manes (1966), Doong and Yang (1988), and the IAS theory of Myers and Prausnitz (1965). Of these, the last two methods were considered most suited for the current problem and only these will be explained in detail in the following discussion. 3.1. DoongYang (DY) potential theory model The potential theory model of Doong and Yang (1988) is based on the concept of maximum available pore volume. The advantages of this model are that it is non-iterative, simple to implement, and has been shown to favorably predict binary mixture adsorption on a wide range of microporous sorbents. Though initially proposed for an ideal system, it is possible to extend the model

S. U. Rege, R. T. Yang / Chemical Engineering Science 56 (2001) 37813796

3785

to non-ideal mixtures and systems showing hysteresis as well (Doong & Yang, 1987; Huggahalli & Fair, 1996). The basis of the DY model is the well-known Dubinin Astakhov (DA) equation which is written as follows: Vi = V0i exp Ci ln where Ci = RT : (E )i (2) Psi Pi
mi

equation is thus written as ni = n0i exp Ci ln Psi Pi


mi

(7)

(1)

The adsorbed molar amount in the mixture, is then calculated by dividing the adsorbed molar volume Vi in Eq. (6) by the molar volume of the sorbate (Vmi ): Vi : (8) ni = Vm i 3.2. Ideal adsorbed solution (IAS) theory The well-known IAS theory has been extensively reviewed in literature (Myers & Prausnitz, 1965; Richter et al., 1989; Yang, 1997) and hence only the working equations will be given here. The theory assumes that the adsorbed phase is an ideal solution of the adsorbed components and the reduced spreading pressure ( i ) of all the components in the mixture in their standard states is equal to the reduced spreading pressure of the adsorbed mixture ( ). Thus,
1

By substituting the exponential portion of Eq. (1) with a variable AI as follows: Ps Ai = exp Ci ln i Pi
ni

(3)

the pure component DA equation can be compactly rewritten as Vi = V0i Ai : (4) The DY model assumes an ideal system in which the adsorbed species do not laterally interact with each other on the sorbent surface. Hence, the parameters mi , and Ci , for any component are assumed to be unperturbed by the other components. However, the maximum available micropore volume is reduced from V0i to (V0i Vi ), where the summation is carried over all other components. The model allows for the di erence in values of the maximum micropore volume for the various components V0i . For example, the water molecules in a zeolite have access to more cavities than other molecules such as CO2 and hence have di erent available micropore volumes. However in order to obtain physically realistic results, it is suggested that the maximum micropore volume for all the components in the mixture be equal (Rege et al., 2000). Thus for a gas mixture composed of two components at equilibrium, the adsorbed amounts of each species on the sorbent can be obtained from the solution of following simultaneous linear equations: V1 = (V01 V2 )A1 ; V2 = (V02 V1 )A2 : (5)

= =

(9)

The reduced spreading pressure of each component is computed from Gibbs adsorption isotherm as follows:
i

RT

iA

Pi0 0

ni d Pi ; Pi

(10)

where Pi0 is the partial pressure of the single component i in its standard state. This parameter is ultimately xed by the spreading pressure of the resulting mixture. In Eq. (10), ni (Pi ) is the adsorption isotherm, which in the present case is the DA equation (Eq. (7)). The adsorbed-phase mole fraction of the di erent components are related to those of the gas phase by the Raoults law for ideal solutions, analogous to vaporliquid systems: Pyi = Pi0 ( )xi with the constraint that
N i=1 N i=1

(11)

The solution to the above equations is explicitly given as V1 = A1 (V01 V02 A2 ) ; 1 A1 A2 (6)

xi = 1;

yi = 1:

(12)

The total adsorbed amount of the gas nT is calculated from xi as follows:


N x 1 i = ; nT i = 1 n0 i

A2 (V02 V01 A1 ) V2 = : 1 A1 A2

(13)

It is common to express the micropore and adsorbed volumes in the DA equation in terms of molar quantities. This can be done by dividing Eq. (1) by the adsorbate molar volume which can be calculated by various methods (Reich et al., 1980). The molar form of the DA

where n0 i is the amount adsorbed in the standard state calculated from an isotherm (in this case the DA model, Eq. (7)) at the pressure Pi0 : The actual amount of each component is, therefore, n i = n T xi : (14)

3786

S. U. Rege, R. T. Yang / Chemical Engineering Science 56 (2001) 37813796

Eqs. (9) (14) provide the necessary framework for IAS mixed adsorbate predictions. The solution strategy is facilitated by rst obtaining Pi0 as an analytic function of . In case of simple isotherms such as the Langmuir model, an explicit analytic solution is possible. However in the case of the DA model, a simple analytic solution for Pi0 ( ) cannot be obtained. Hence, was rst calculated by a numerical integration of the ratio (ni =Pi0 ) as a function of Pi0 (a graphical method can also be used but is more tedious). Next, an empirical algebraic function of the following form was t to the calculated data with a; b; c; and k as tting parameters: Pi0 ( i ) = a exp b ln c
k

H2 O (190 ppm) and CO2 (40 ppm). For the -Al2 O3 sorbent, the limits were H2 O (30 ppm) and CO2 (390 ppm): These limits will be important for analyzing the performance of the IAS-DA model for mixed gas predictions for the systems under study. 3.3. Calculation of activity coe cients for real adsorbed solutions It is intuitive that a system composed of two polar components (CO2 and H2 O) would be far from ideal as the components are likely to have strong lateral interactions in the adsorbed phase. However both the Doong Yang model and the IAS theory, in their original form, assume an ideal system in which the adsorbates in the mixture do not interact with each other. Also, it has been shown that the heterogeneity of surface energy in an adsorbent can also result in an apparent deviation from Raoults law (Myers, 1983). Hence, to enable a t of the measured experimental data, it was necessary to calculate the activity coe cients of the adsorbates considering it to be a non-ideal adsorbed solution (Myers & Prausnitz, 1965; Costa et al., 1981). This was done by employing a modied form of the Raoults law given in Eq. (11): Pyi = i Pi0 ( )xi ; (17) where i is the activity coe cient for component i. In principle, the molar area of the mixed adsorbates should also be included in the analysis instead of Eq. (13) (Yang, 1997): n x RT N @ ln i 1 i = + xi : (18) 0 nT i = 1 ni A i=1 @ xi However this can greatly complicate the calculations, and in the interest of simplicity, Eq. (13) was used without the correction for the molar area for mixing following Costa et al. (1981). The activity coe cients were thus calculated from Eq. (17), using the experimental gas- and adsorbed-phase mole fractions (yi ; xi ) , and Pi0 calculated from the IAS method discussed in preceding sections. 4. Results and discussion 4.1. Pure component CO2 spectra As explained in the experimental section, the calibration for CO2 was done in two concentration ranges: 0 300 ppm and 1000 40; 000 ppm (0.1 4%). Fig. 1 shows the adsorbed CO2 spectra on 13X zeolite at various concentrations in the 0 300 ppm range. The spectra shows a peak at 2359 cm1 which corresponds to the undissociated molecular surface species. This band denoting the physisorbed CO2 species typically appears

(15)

This function actually resembles the inverse form of the DA equation and was found to t the reduced pressure data reasonably well. The rest of the solution strategy is fairly straightforward (Yang, 1997). The IAS theory in conjunction with the DA equation (henceforth referred to as the IAS-DA model) was recently shown to be more accurate for mixed gas adsorption prediction of the CH4 = CO2 on coal system (Clarkson & Bustin, 2000) than conventional models such as the extended Langmuir equation. A similar conclusion was reached by Richter et al. (1989), who concluded that the DubininRadushkevich equation (a special case of the DA equation) was the best choice for a single-component isotherm as a basis for mixed gas predictions by IAS theory. This was attributed to the e ectiveness of the Dubinin-type models in tting isotherm data over a wide range of concentrations. However Richter et al. also caution about an important pitfall of using the DA-type of models in the IAS theory. This relates to the asymptotic behavior of the isotherm at low coverage. In principle, any adsorption equation should reduce to the Henrys law at low coverage in order to be thermodynamically consistent. In other words, lim ni Pi = Hi : (16)

Pi 0

However, the ratio ni =Pi in case of the DA model rst goes through a maximum and thereafter reduces to zero as Pi tends to zero. Thus, the calculated spreading pressure ( ) (and hence the mixture loading prediction by IAS theory) would be inaccurate at pressures lower than that corresponding to the maximum in the ni =Pi , since the limits of the DA equation are in error. For the systems under study, the ratio of ni =Pi was plotted as a function of Pi and the limits of partial pressure of the sorbates below which the IAS predictions would be inaccurate were therefore determined from the position of the respective maxima. For the 13X zeolite sorbent, the limits were as follows:

S. U. Rege, R. T. Yang / Chemical Engineering Science 56 (2001) 37813796

3787

Fig. 1. FTIR spectra of pure component CO2 at di erent concentrations (0 300 ppm v= v) on 13X (NaX) zeolite at 295 K. The inset gures denote the corresponding concentration in ppm by volume.

Fig. 2. FTIR spectra of pure component CO2 at di erent concentrations (0.1 4% v= v) on 13X (NaX) zeolite at 295 K. The inset gures denote the corresponding concentration in percentage (%) by volume.

in the region of 2350 2370 cm1 and is known as the 3 band. In addition, several peaks corresponding to the chemisorbed species appear in the region of 1200 1700 cm1 . These typically correspond to CO2 molecules adsorbed either in a linear conguration on the cations ( 2 band), or in a non-linear (or bent) manner in form of carbonate= bicarbonate ions. The observed peaks are in accordance with previous observations in literature (Bertsch & Habgood, 1963; Jacobs et al., 1973). In the low concentration range, it appears that the intensity of the chemisorbed peaks is much higher than that of the physisorbed species. This is intuitive since most of the adsorption at very low partial pressures is known to be chemical in nature characterized by long equilibration times. The single component CO2 adsorption on 13X zeolite in the higher concentration of 0.1 4% is shown in Fig. 2. In this case, the 3 peak at 2359 cm1 corresponding to the physisorbed CO2 molecules, is much more dominant than those corresponding to the chemisorbed species in the 1200 1700 cm1 range. It was also observed that the 3 peak broadens as the concentration of CO2 in the gas phase is increased and gets resolved into several peaks. This ne structure of the CO2 band probably re ects adsorption on the di erent sites available for CO2 adsorption. The FTIR spectra of adsorbed CO2 in -Al2 O3 yielded is expected to be similar in some respects to that observed for 13X zeolite. Similar to the zeolite surface, the alumina surface also presents oxide ion sites for CO2 adsorption, especially at locations of surface defects. Moreover, when the temperature of activation of the alumina sample is below 500 C, a relatively large hydroxyl concentration is believed to exist on the surface (Parkyns, 1971). This results in the surface reaction of CO2 with the hydroxyl groups forming bicarbonates on the surface, characterized

Fig. 3. FTIR spectra of pure component CO2 at di erent concentrations (0 300 ppm v= v) on -alumina at 295 K. The inset gures denote the corresponding concentration in ppm by volume.

by adsorption bands at 3605, 1640, 1480 and 1235 cm1 . Also uncoordinated, unidentate, and bidentate carbonate species may result at isolated oxide ions on the surface. At room temperature, the formation of these carbonate compounds is expected to be rather slow, but may account for a signicant amount of adsorbed CO2 at low pressures. The low-pressure (0 300 ppm) CO2 adsorption spectra on -Al2 O3 are shown in Fig. 3. As discussed before, a peak at 2361 cm1 was observed which refers to the undissociated CO2 molecules physisorbed on the surface. Also, a few absorption bands were observed in the 1200 1700 cm1 range which correspond to the chemisorbed species. The IR spectra of the adsorption at a higher CO2 concentration range (0.1 4%) is shown in Fig. 4. The physisorption peaks are seen to be of greater intensity than

3788

S. U. Rege, R. T. Yang / Chemical Engineering Science 56 (2001) 37813796

Fig. 4. FTIR spectra of pure component CO2 at di erent concentrations (0.1 4% v= v) on -alumina at 295 K. The inset gures denote the corresponding concentration in percentage (%) by volume.

Fig. 5. FTIR spectra of pure component H2 O at di erent concentrations (0 800 ppm v= v) on 13X (NaX) zeolite at 295 K. The inset gures denote the corresponding concentration in ppm by volume.

the chemisorption peaks. The chemisorption peaks were observed at 1644, 1432, and 1226 cm1 which are in accordance with previously reported data (Lee & Condrate, 1995). 4.2. Pure component H2 O spectra The zeolite surface is composed of microporous cavities which are lined with oxide ions attached to the SiAl framework. In addition to oxide ions, a number of charge balancing metal cations (such as Na+ -ions in case of 13X zeolite) exist at di erent site locations within the crystal lattice. Thus several congurations are available for an adsorbed molecule depending on the location and nature of the interaction of the sorbate with either the oxide ion or the cation. In case of H2 O molecules, three basic absorption frequencies have been observed in zeolites in general (Bertsch & Habgood, 1963). These include a single sharp band of isolated OH stretching vibration in the 3690 3700 cm1 , a broad band characteristic of hydrogen-bonded OH stretching vibration in the 3250 3400 cm1 , and the usual HOH bending vibration band at 1645 1660 cm1 . The pure component H2 O spectra on 13X zeolite are shown in Fig. 5. As discussed above, a prominent sharp band of HOH bending was seen at 1644 cm1 . In addition, there was a broad band related to the H-bonded OH stretching between 2900 3500 cm1 . It was observed that there was a shift of the isolated OH stretching peaks above 3500 cm1 towards the lower frequencies as the H2 O concentration in the gas phase was increased from 80 to 900 ppm. This was due to the increase in amount of H2 O molecules in the zeolite cavity which resulted in an increase in the hydrogen bonded H2 O molecules in comparison to isolated H2 O molecules.

Fig. 6. FTIR spectra of pure component H2 O at di erent concentrations (0 800 ppm v= v) on -alumina at 295 K. The inset gures denote the corresponding concentration in ppm by volume.

The adsorbed H2 O spectra on -Al2 O3 is shown in Fig. 6. As was observed for the zeolite sample, there is a peak at 1643 cm1 and a broad band of hydrogen-bonded OH stretching vibrations in the region of 2800 3500 cm1 . Unlike the HOH bending band at 1650 cm1 , the stretch bands did not appear to be well-dened. This is probably due to the abundance of H2 O molecules present on the surface which obscure the isolated OH stretch band. It was di cult to verify the observed IR spectra with those published in literature since most of the previous studies considered a nearly dehydrated surface which was only sparsely inhabited by hydroxyl groups. Also most of the studies were carried out at high temperatures (100 C) since the emphasis was on the catalytic properties of alumina rather than on the adsorptive properties. In contrast, the

S. U. Rege, R. T. Yang / Chemical Engineering Science 56 (2001) 37813796

3789

Fig. 7. Calibration of the amount of CO2 adsorbed on 13X (NaX) zeolite vs. the corresponding area of the FTIR spectrum peak at 2358 cm1 (2230 2390 cm1 region).

Fig. 8. Calibration of the amount of H2 O adsorbed on -alumina vs. the corresponding area of the FTIR spectrum peak at 1650 cm1 (1550 1760 cm1 region).

present study involves adsorption of H2 O molecules as much as 3:0 mmol= g on -Al2 O3 , at room temperature, and hence a slightly di erent spectral picture was observed. 4.3. Calibration of single-component peak areas In principle, it is possible to determine true isotherms from infrared measurements by a mathematical analysis of the absorption peak areas using the BeerLamberts law. However, it involves making certain assumptions which may not be valid under all experimental conditions, as well as it requires accurate measurement of pellet dimensions (Delaval et al., 1986). Instead, it was decided to calibrate the spectra of the adsorbed single-component CO2 and H2 O molecules on the NaX zeolite and the -Al2 O3 with the known adsorbed amount at the corresponding partial pressure. As mentioned earlier, the isotherms for these sorbatesorbent systems were measured in a previous work using an independent experimental technique (Rege et al., 2000). Since the areas of the IR absorption peaks are proportional to molar adsorbed amounts, and the loading on the sorbents were known at di erent partial pressures from the earlier isotherm measurements, a calibration could be made relating the absorbance area to amount adsorbed. For the sake of brevity, only a few of the calibration curves will be shown here. Fig. 7 shows the adsorbed amount of CO2 on 13X zeolite versus the area of the absorbance peak at 2358 cm1 (2350 2370 cm1 region), while Fig. 8 shows a similar calibration for H2 O on -Al2 O3 for the peak at 1650 cm1 (1550 1760 cm1 region). It can be seen that the calibration curves were of fairly linear nature. It must be noted however that

they may not necessarily be linear over the entire range as was observed by Bertsch and Habgood (1963) for absorbance of adsorbed H2 O molecules on NaX. This might be an explanation as to why the calibration line in Fig. 8 does not pass through the origin. Also, the measured peak areas correspond to sorbate molecules in the adsorbed phase as well as the gas phase contained within the IR cell. This fact is specically relevant for CO2 since gas-phase CO2 molecules have a considerable IR absorbance as well. The gas-phase absorbance of water-vapor, on the other hand, was found to be relatively negligible. Deducting the gas-phase CO2 absorbance did pose a di culty, but it was found that reasonably good results were obtained without correcting for the same. Another assumption made in the calibrations was that the physisorbed CO2 ( 3 ) peak is representative of the total adsorbed amount including the chemisorbed species. This assumption, though questionable, is necessary since the isotherm measured by conventional techniques cannot di erentiate between physisorption and chemisorption and the measured adsorption is a sum of the two types of adsorption. A complication arises specically for the CO2 H2 O system since one of the CO2 chemisorption peaks appear in the 1650 cm1 region which is actually supposed to indicate the H2 O adsorbed amount in the mixture experiments. Since the area of the surface-coordinated CO2 species will add to the area of the adsorbed H2 O molecules, an overestimated value for the H2 O adsorbed amount may be obtained during the mixture measurements. This can be overcome by obtaining a calibration of the CO2 chemisorption peak with the CO2 adsorbed amount as predicted by the physisorption peak at 2350 cm1 . This calibration is shown in Figs. 9 and 10 for 13X zeolite and -Al2 O3 , respectively, which

3790

S. U. Rege, R. T. Yang / Chemical Engineering Science 56 (2001) 37813796

Fig. 9. Calibration of the peak area in the 2230 2390 cm1 region for the CO2 species chemisorbed on 13X (NaX) zeolite vs. the corresponding CO2 adsorbed amount as determined by the peak at 2358 cm1 .

Fig. 11. FTIR spectra of various compositions of a H2 O-vapor= CO2 gas mixture adsorbed on 13X (NaX) zeolite at 295 K. The inset gures denote the corresponding run number described in Table 1.

in conjunction with the previous pure component H2 O calibration. 4.4. Measurement of CO2 =H2 O mixed adsorption The FTIR spectra of H2 O-vapor = CO2 mixed gas of varying concentration in an inert carrier (helium) on 13X zeolite is shown in Fig. 11. The concentration of H2 O vapor in the mixture was gradually increased from 0 1000 ppm v= v in each run. The concentration of CO2 was simultaneously increased from 0.1 to 4% in each subsequent run. The actual mixture composition in each run is given in Table 1. As expected, the gure appears like a superposition of the pure component spectra of CO2 and H2 O. Again, the CO2 physisorbed amount appeared in the IR peak at 2360 cm1 while the H2 O adsorbed amount in the mixture was visible from the 1640 cm1 peak and the OH stretching region around 3000 3500 cm1 . There were also some CO2 chemisorption peaks in the 1200 1800 cm1 region, some of which also appeared in the HOH bending region (1650 cm1 ). The FTIR spectra for the same composition on -Al2 O3 shown in Fig. 12 also show similar features. The spectra for the mixed H2 O= CO2 adsorption at a lower CO2 concentration range (0 300 ppm v= v) was also measured for both the sorbents, but are not shown here since they are of a similar nature. The results for these runs are instead summarized in Tables 1 and 2. The measured CO2 and H2 O absorbance peak areas were next converted to adsorbed molar amounts (mmol= g) using the calibration carried out for pure component adsorbed amounts. The CO2 adsorbed amount was calculated from the 2350 cm1 peak, while that of H2 O was calculated from the 1650 cm1 peak area after correcting for the CO2 chemisorption peak area in the same wavenumber region. The results of the

Fig. 10. Calibration of the peak area in the 2230 2390 cm1 region for the CO2 species chemisorbed on -alumina vs. the corresponding CO2 adsorbed amount as determined by the peak at 2358 cm1 .

have a nature inverse of that of Figs. 7 and 8. For the 13X zeolite it appears that the CO2 chemisorption peak grows in direct proportion to the physisorption peak. In case of -Al2 O3 though, the growth of the chemisorbed CO2 peak shows a weaker dependence and it levels o at higher CO2 partial pressures. The actual amount of H2 O adsorbed was then estimated as follows: (1) the CO2 adsorbed amount was estimated from the peak at 2350 cm1 , (2) the absorbance area of the corresponding chemisorption peak at 1650 cm1 was estimated assuming the calibration shown in Figs. 9 and 10 is valid, (3) the absorbance peak area of the chemisorbed CO2 peak was deducted from the measured peak area at 1650 cm1 to give the absorbance purely due to the H2 O molecules. The amount of adsorbed H2 O in the mixture was thus calculated using the corrected H2 O absorbance at 1650 cm1

S. U. Rege, R. T. Yang / Chemical Engineering Science 56 (2001) 37813796

3791

Table 1 Adsorption equilibrium results for H2 O-vapor= CO2 mixtures on 13X (NaX) zeolite at 295 K using the FTIR technique. The corresponding single-component loadings predicted by the DubininAstakhov equation are also shown for comparison Run No. Adsorbate concentration H2 O (ppm) 1 2 3 4 5 6 7 8 9 10 11 219 336 560 685 814 1070 133 206 379 551 760 CO2 18 ppm 38 ppm 102 ppm 149 ppm 198 ppm 260 ppm 0.16% 0.34% 1.22% 2.39% 3.59% Expt. adsorbed amounts H2 O (mmol= g) 2.89 6.13 8.84 10.08 10.07 10.55 4.51 6.52 8.09 7.39 4.96 CO2 (mmol= g) 0.091 0.166 0.249 0.293 0.299 0.328 0.33 0.37 0.92 1.68 2.94 Single-component predicted loading H2 O (mmol= g) 4.08 5.89 8.17 9.04 9.67 10.83 2.33 3.85 6.43 8.10 9.48 CO2 (mmol= g) 0.021 0.049 0.127 0.178 0.226 0.281 0.96 1.43 2.48 3.12 3.52

Table 2 Adsorption equilibrium results for H2 O-vapor= CO2 mixtures on -Al2 O3 at 295 K using the FTIR technique. The corresponding single-component loadings predicted by the DubininAstakhov equation are also shown for comparison Run No. Adsorbate concentration H2 O (ppm) 12 13 14 15 16 17 18 19 20 21 219 336 560 685 814 133 206 379 551 760 CO2 18 ppm 38 ppm 102 ppm 149 ppm 198 ppm 0.16% 0.34% 1.22% 2.39% 3.59% Expt. adsorbed amounts H2 O (mmol= g) 1.16 1.19 1.35 1.57 2.71 1.29 1.63 1.78 1.81 1.96 CO2 (mmol= g) 3:24103 4:27103 5:64103 5:84103 6:84103 0.06 0.09 0.13 0.19 0.31 Single-component predicted loading H2 O (mmol= g) 1.66 1.95 2.27 2.39 2.48 1.31 1.61 2.03 2.26 2.45 CO2 (mmol= g) 2:14104 6:20104 2:23103 3:52103 4:88103 0.04 0.07 0.18 0.27 0.33

mixture adsorption experiments for 13X zeolite are given in Table 1, while those for -Al2 O3 are given in Table 2. 4.5. Sorption enhancement The single-component DA isotherm prediction for the respective adsorbates at the corresponding partial pressure in the mixture are given alongside the actual mixture measurement results in Tables 1 and 2. This would give the respective amounts of CO2 and H2 O that would be adsorbed for the various runs assuming that the adsorption was exclusive of the other species. In other words, these predictions would hold true for a hypothetical situation in which there is an absence of competition between the adsorbates for the sorbent volume or surface area. A comparison of the experimental and single-component

isotherm predicted data sets in Tables 1 and 2 reveals an interesting phenomenon. For the data sets in which the CO2 is of a relatively lower concentration (300 ppm), the experimentally measured adsorbed amount of CO2 was found to be greater than the pure component predicted amount at the same partial pressure. This observation is highly counter-intuitive since typically there is a reduction in the amount of adsorption of both sorbates in a mixture due to increased competition for adsorption sites on the sorbent surface. The enhancement factor, dened here as the ratio of experimental adsorbed loading to single-component predicted loading, ranged from 1.2 to 4.3 for the 13X zeolite and from 1.4 to 15 for -Al2 O3 . This factor was the highest for the lowest CO2 partial pressure in the data sets and decreased steadily as the partial pressure of H2 O-vapor and CO2 in the mixture increased. It needs to be mentioned that the CO2 = H2 O experiments of both the zeolite and alumina samples were

3792

S. U. Rege, R. T. Yang / Chemical Engineering Science 56 (2001) 37813796

Fig. 12. FTIR spectra of various compositions of a H2 O -vapor= CO2 gas mixture adsorbed on -alumina at 295 K. The inset gures denote the corresponding run number described in Table 2.

repeated several times using di erent sorbent samples, but the strong enhancement was reproducibly observed each time. A study of the data sets for CO2 in the higher concentration range (0.1 4% v= v) reveals that there is little or no enhancement in the CO2 sorption at the same amount of H2 O partial pressure in the mixture. In fact, the adsorbed CO2 amount was found to be suppressed in most of the runs, as is usually the case. However for the higher CO2 concentration range there was instead a mild enhancement in the amount of H2 O adsorbed, especially for the zeolite sorbent. The enhancement factor for H2 O ranged from 1.4 to 2.3 for 13X zeolite. The enhancement in sorption of one component in a mixture by virtue of a presence of a second component (usually water) is not an entirely new phenomenon. Similar enhancement in the adsorption of organic vapors on activated carbons with coadsorption of H2 O has been previously reported. For example, Matsumara, Yamabe, and Takahashi (1985) observed a cooperative enhancement in methanol adsorption on carbon in a humid atmosphere. Water-vapor promoted enhancements on carbon-based sorbents have also been reported for trichloroethylene based on xed-bed breakthrough experiments (Kane, Bushong, Foley, & Brendley, 1998), and for lower alcohols by LeVan and co-workers (Taqvi & LeVan, 1996; Taqvi et al., 1999). In the latter case, there was also an enhanced water adsorption coupled with the increased alcohol sorption. In the inorganic domain as well, there is evidence of increased NH3 sorption on carbon in presence of H2 O vapor for heat-pump cycles (Meunier, 2000). While most of the cooperative enhancement in sorption of polar organic compounds on carbon can be

attributed to factors such as hydrogen bonding (Taqvi et al., 1999), the explanation for increased loading of an inorganic compound such as CO2 is less straightforward. Interestingly, Bertsch and Habgood (1965) also observed a dramatic acceleration in the adsorption of CO2 on NaX zeolite in presence of small amounts of preadsorbed water. While dehydrated NaX zeolite took more than 60 h to achieve equilibrium for CO2 at low partial pressure, true equilibrium was found to be attained within a few minutes when the zeolite was preadsorbed with small amounts of H2 O. It is quite possible that the isotherms measured using the static volumetric technique earlier (Rege et al., 2000) were actually pseudo-equilibrium results since the time for equilibration was impractically long. Thus, the enhancement in the adsorbed equilibrium amount may simply be attributed to increase in the uptake rate of the sorbent in presence of moisture. Bertsch and Habgood (1965) suggest that the increase in uptake rate of CO2 may possibly be due to the decrease in energy of activation for di usion. Molecules need to pass through channels with high-energy sites in order to di use through the lattice. By preferentially adsorbing onto the high-energy sites, the CO2 can bypass these and hence the energy of activation for di usion may be reduced. However, since the NaX zeolite has large pores compared to the size of a CO2 molecule, di usion is probably not a limiting factor. The slow rate of uptake at low pressures (in absence of moisture) is likely to be due to the chemical nature of the adsorption under these conditions. The enhanced adsorption may also be due to the faster formation of bicarbonate compounds on the sorbent surface. The presence of water molecules on the surface probably catalyzes the formation of carbonate ion by stabilizing the charged complex. Further experiments will be necessary to verify this fact. However since the H2 O spectra remain unchanged, it probably does not directly interact with the surface coordinated CO2 species. 4.6. Mixture adsorption model predictions As a natural follow-up to the measurement of the mixture equilibria of CO2 and H2 O on NaX zeolite and -alumina, the experimental data was tested against two model predictions, namely the DoongYang (DY) and the IAS-DA models reviewed earlier. The values of the pure component DA model (Eq. (7)) as well as some physical properties of the sorbates at room temperature required by the models are given in Table 3. A comparison of the model predictions with the experimental data is shown in Table 4. It can be seen from the table that the results from the DY model are reasonably close to the IAS-DA models, especially at higher sorbate concentrations. A graphical

S. U. Rege, R. T. Yang / Chemical Engineering Science 56 (2001) 37813796 Table 3 Parameters of the single-component DubininAstakhov isotherm (Eq. (7)) at 295 K used in predicting mixture adsorption Sorbent 13X (NaX) Zeolite -Al2 O3 Sorbate H2 O CO2 H2 O CO2 n0 (mmol= g) 16.00 7.056 3.364 1.483 C 0.125 0.120 0.111 0.160 m 6.0 3.0 5.0 2.5 Ps (atm) 1.00 57.53 1.00 57.53

3793

Vm (cm3 = mol) 18.78 42.61 18.78 42.61

Table 4 Comparison of experimental adsorption equilibrium results for H2 O-vapor= CO2 mixtures for the runs described in Tables 1 and 2, with Doong Yang (DY) model and ideal adsorbed solution theory using the DA equation as basis (IAS-DA) Run No. Expt. adsorbed amounts H2 O (mmol= g) 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 2.89 6.13 8.84 10.08 10.07 10.55 4.51 6.52 8.09 7.39 4.96 1.16 1.19 1.35 1.57 2.71 1.29 1.63 1.78 1.81 1.96 CO2 (mmol= g) 0.091 0.166 0.249 0.293 0.299 0.328 0.33 0.37 0.92 1.68 2.94 3:24103 4:27103 5:6103 5:84103 6:84103 0.06 0.09 0.13 0.19 0.31 DY Model predicted values H2 O (mmol= g) 4.08 5.87 8.10 8.94 9.64 10.68 2.05 3.22 4.86 5.82 6.74 1.66 1.95 2.27 2.39 2.48 1.29 1.57 1.92 2.11 2.27 CO2 (mmol= g) 0.016 0.031 0.063 0.079 0.090 0.094 0.84 1.14 1.73 1.99 2.04 1:09104 2:61104 7:26104 1:02103 1:28103 0.02 0.04 0.08 0.10 0.11 IAS-DA predicted values H2 O (mmol= g) 4.04 5.84 8.10 8.97 9.69 10.78 1.37 2.10 3.26 4.22 5.40 1.66 1.95 2.27 2.39 2.48 1.30 1.61 2.02 2.26 2.40 CO2 (mmol= g) 0.021 0.027 0.033 0.034 0.033 0.025 1.02 1.40 2.10 2.38 2.36 3:36105 2:88105 1:67105 8:67106 7:56105 0.007 0.007 0.007 0.004 0.002

comparison of the parity of the two models with the measured experimental data is shown in Figs. 13 and 14. Both the models were reasonably successful in predicting the H2 O loading for the mixture on both NaX zeolite and -Al2 O3 , although the IAS-DA model tended to underpredict a part of the data to a greater degree than the DY model. The CO2 predictions were more instrumental in discriminating between the models. As such both the models were fairly o as far as predicting the CO2 adsorption at very low partial pressures was concerned. This was expected since both the mixed-adsorbate theories assume ideality and the lateral interactions among the components is neglected. Among the two models, the DY predicted data appears to be closer to the experimental data as can be adjudged from the proximity to the parity diagonal in Fig. 13.

The IAS-DA results were grossly in error as far as the very low partial pressure predictions were concerned. This is primarily due to the erroneous limits of the DA model which does not tend to the Henrys law as pressure tends to zero. As explained in an earlier section which reviewed the IAS theory, there is an error in calculating the spreading pressure when the DA equation is used as a basis below a certain partial pressure which corresponds to a maximum in the ni =Pi ratio. Below this partial pressure, the ni =Pi ratio tends to deviate from real behavior and goes towards zero. Hence the IAS-DA theory cannot be used for predicting the mixed gas adsorption behavior below a certain concentration. In case of -Al2 O3 , this limiting partial pressure for CO2 for deviation from Henrys law was as high as 390 ppm. Hence most of the predictions for CO2 on -Al2 O3 with a CO2 concentration below this limit had a large error.

3794

S. U. Rege, R. T. Yang / Chemical Engineering Science 56 (2001) 37813796

Fig. 13. Parity plot of the predictions of the H2 O-vapor= CO2 mixture equilibrium results on 13X (NaX) zeolite and -alumina sorbents using the DoongYang model vs. the experimentally measured values.

Fig. 15. Activity coe cients ( i ) calculated using the non-ideal adsorbed solution theory vs. the adsorbed phase mole fractions of the two components (1: CO2 and 2: H2 O) for (A) 13X (NaX) zeolite, and (B) -alumina. The trend lines shown in the gure show the Wilson equation ts to the calculated i values.

Fig. 14. Parity plot of the predictions of the H2 O-vapor= CO2 mixture equilibrium results on 13X (NaX) zeolite and -alumina sorbents using the IAS-DA model vs. the experimentally measured values.

4.7. Calculation of activity coe cients As was discussed previously, the CO2 = H2 O systems displayed a strong deviation from ideality as predicted by the DoongYang and the IAS models. In an attempt to t the experimental data to a predictive model, it was decided to utilize the non-ideal adsorbed solution theory following Costa et al. (1981). The calculated activity coe cients ( i ) vs. the adsorbed-phase mole fractions of the two components (CO2 and H2 O) for both the sorbents are shown in Fig. 15. The Wilson equation (Smith & Van Ness, 1987) was used to t the calculated activity coe cients and is depicted by the trend lines in the gure. As can be seen, the i values could be t with only moderate success. It must be noted how-

ever that these activity coe cients in fact correct for the errors associated with using the DA model as the basis equation in the IAS theory as well as account for the apparent non-ideality due to sorbent heterogeneity, etc. Hence, the calculated i values should be considered as merely tting parameters and their physical or thermodynamic interpretation should be treated with caution. 5. Conclusions This work introduced FTIR spectroscopy has an efcient tool for the qualitative as well as quantitative analysis of mixed gas adsorption systems particularly at very low adsorbate concentrations. The utility of the FTIR technique was demonstrated by studying the equilibria of CO2 and H2 O-vapor mixtures on NaX (13X) zeolite and -Al2 O3 sorbents at low partial pressures. Apart from measuring the adsorption equilibrium data,

S. U. Rege, R. T. Yang / Chemical Engineering Science 56 (2001) 37813796

3795

this method was capable of distinguishing between the physisorbed and chemisorbed species on the surface and can also indicate any sorbatesorbate interactions if present. The results of the CO2 = H2 O equilibria on the zeolite and alumina samples studied in this work revealed that there was a distinct enhancement in the amount of CO2 and H2 O adsorbed at very low partial pressures. This enhancement gradually decreased and nally disappeared as the sorbate partial pressure is increased. It is hypothesized that the enhancement in the CO2 adsorbed amount in presence of traces of moisture was due to the catalysis of the formation of bicarbonate species on the sorbent surface by the adjacent H2 O molecules. Enhancement of the sorbate di usion due to the occupation of high energy sites by water molecules is also a possibility. Further work is necessary to completely understand and model the enhancement in this adsorption system as well as in other systems displaying a similar behavior. The experimental mixture adsorption data was t to the DoongYang model and the ideal adsorbed solution (IAS) theory which employed the DubininAstakhov equation as a basis. It was found that the DoongYang model gave a marginally better t to the mixed adsorption data than the IAS model. Although the H2 O loading was reasonably well-predicted by both the models, they failed to predict the enhancement in CO2 adsorption at very low partial pressures in presence of traces of moisture. This was due to the assumption of ideality which fails to account for lateral sorbatesorbate interactions. The experimental data was hence t by calculating activity coe cients using the real adsorbed solution theory. Notation A C E H m n n0 nT N P Ps R T V Vm V0 x y specic surface area of sorbent, cm2 = g DubininAstakhov equation parameter characteristic energy of the sorbent, J= mol Henrys constant, mmol= g= atm DubininAstakhov equation parameter adsorbed molar amount, mmol= g limiting molar amount, mmol= g total adsorbed amount, mmol= g total number of components partial pressure of adsorbate, atm saturation pressure of adsorbate, atm gas constant, J= mol= K absolute temperature, K adsorbed volume in micropore, cm3 = g molar volume of adsorbate, cm3 = mol limiting micropore volume, cm3 = g adsorbed-phase mole-fraction gas-phase mole-fraction

Greek letters

a nity coe cient of the sorbent activity coe cient spreading pressure of adsorbate, N= m reduced spreading pressure, mmol= g

Acknowledgements This work was supported by NSF under Grant CTS-9819008.


References
Angell, C. L., & Scha er, P. C. (1965). Infrared spectroscopic investigations of zeolites and adsorbed molecules, I. Structural OH groups. Journal of Physical Chemistry, 69, 3463. Appel, W. S., LeVan, M. D., & Finn, J. E. (1998). Nonideal adsorption equilibria described by pure component isotherms and virial mixture coe cients. Industrial and Engineering Chemistry Research, 37, 4774. Bering, B. P., Serpinsky, V. V., & Surinova, S. I. (1963). Doklady Akademii Nauk SSSR, 153, 129. Bertsch, L., & Habgood, H. W. (1963). An infrared spectroscopic study of the adsorption of water and carbon dioxide by Linde molecular sieve X. Journal of Physical Chemistry, 67, 1621. Chang, C. H., Stonesifer, G. T., Cusick, R. J., & Hart, J. M. (1991). Comparison of metal oxide adsorbents for regenerative carbon dioxide and water vapor removal for advanced portable life support systems. Space Station and Advanced EVA: SAE Special Publications, 872, 1. Chen, Y. D., Yang, R. T., & Uawithya, P. (1994). Di usion of oxygen, nitrogen and their mixtures in carbon molecular sieve. A.I.Ch.E. Journal, 40, 577. Clarkson, C. R., & Bustin, R. M. (2000). Binary gas adsorption= desorption isotherms: e ect of moisture and coal composition upon carbon dioxide selectivity over methane. International Journal of Coal Geology, 42, 241. Costa, E., Sotelo, J. L., Calleja, G., & Marron, C. (1981). Adsorption of binary and ternary hydrocarbon gas mixtures on activated carbon: experimental determination and theoretical prediction of the ternary equilibrium data. A.I.Ch.E. Journal, 27, 5. Delaval, Y., Seloudoux, R., & de Lara, E. C. (1986). Determination of isotherms and initial heat of adsorption of CO2 and N2 O in four a zeolites from infrared measurements. Journal of Chemical Society of Faraday Transactions I, 82, 365. Dellosso Jr., L., & Winnick, J. (1969). Mixed-gas adsorption and vacuum desorption of carbon dioxide on molecular sieve: bed design for use in a humid atmosphere. Industrial and Engineering Chemistry Process Design and Development, 8, 469. Doong, S. J., & Yang, R. T. (1988). A simple potential-theory model for predicting mixed-gas adsorption. Industrial and Engineering Chemistry Research, 27, 630. Doong, S. J., & Yang, R. T. (1987). Adsorption of mixtures of water vapor and hydrocarbons by activated carbon beds: thermodynamic model for adsorption equilibrium and adsorber dynamics. A.I.Ch.E. Symposium Series No. 259, 83, 87. Grant, R. J., & Manes, M. (1966). Adsorption of binary hydrocarbon gas mixtures on activated carbon. Industrial and Engineering Chemistry Research, 5, 490. Hair, M. L. (1967). Infrared spectroscopy in surface chemistry. New York: Dekker.

3796

S. U. Rege, R. T. Yang / Chemical Engineering Science 56 (2001) 37813796 Peri, J. B. (1965). Infrared and gravimetric study of the surface hydration of -alumina. Journal of Physical Chemistry, 69, 211. Rege, S. U., Yang, R. T., & Buzanowski, M. A. (2000). Sorbents for air prepurication in air separation. Chemical Engineering Science, 55(21), 4789. Reich, R., Zeigler, W. T., & Rogers, K. A. (1980). Adsorption of methane, ethane, and ethylene gases and their binary and ternary mixtures and carbon dioxide on activated carbon at 212301 K and pressures to 35 atmospheres. Industrial and Engineering Chemistry Process Design and Development, 19, 336. Richter, E., Shutz, W., & Myers, A. L. (1989). E ect of adsorption equation on prediction of multicomponent adsorption equilibria by the ideal adsorbed solution theory. Chemical Engineering Science, 44, 1609. Sircar, S., & Kratz, W. C. (1981). Removal of water and carbon dioxide from air. US Patent 4,249,915. Smith, J. M., & Van Ness, H. C. (1987). Introduction to Chemical Engineering Thermodynamics (p. 379). New York: McGraw-Hill. Smith, L., Cheetham, A. K., Morris, R. E., Marchese, L., Thomas, J. M., Wright, P. A., & Chen, J. (1996). On the nature of water bound to a solid acid catalyst. Science, 271, 799. Smudde Jr., G. H., Slager, T. L., Coe, C. G., MacDougall, J. E., & Weigel, S. J. (1995). DRIFTS and Raman investigation of N2 and O2 adsorption on zeolites at ambient temperature. Applied Spectroscopy, 49, 1747. Talu, O., & Meunier, F. (1996). Adsorption of associating molecules in micropores and application to water on carbon. A.I.Ch.E. Journal, 42, 809. Taqvi, S. M., Appel, W. S., & LeVan, M. D. (1999). Coadsorption of organic compounds and water vapor on BPL activated carbon. 4. methanol, ethanol, propanol, butanol, and modeling. Industrial and Engineering Chemistry Research, 38, 240. Taqvi, S. M., & LeVan, M. D. (1996). Nonidealities in vapor-phase coadsorption of organic compounds and water on activated carbon. In: LeVan, M. D. (Ed.), Fundamentals of adsorption (p. 969). Boston: Kluwer. Valenzuela, D. P., & Myers, A. L. (1989). Adsorption equilibrium data handbook.. Englewood Cli s, NJ: Prentice-Hall. Ward, J. W. (1968). A spectroscopic study of the surface of zeolite Y. II. infrared spectra of structural hydroxyl groups and adsorbed water on alkali, alkaline earth, and rare earth ion-exchanged zeolites. Journal of Physical Chemistry, 72, 4211. Yang, R. T. (1997). Gas separation by adsorption processes (Chapter 3). London: Imperial College Press.

Huggahalli, M., & Fair, J. R. (1996). Prediction of equilibrium adsorption of water onto activated carbon. Industrial and Engineering Chemistry Research, 35, 2071. Jacobs, P. A., Cauwelaert, F. H. V., & Vansant, E. F. (1973). Surface probing of synthetic faujasites by adsorption of carbon dioxide. Journal of Chemical Society of Faraday Transactions I, 69, 2130. Jain, R. (1993). Pre-purication of air for separation. US Patent 5,232,474. Jentys, A., Warecka, G., Derewinski, M., & Lercher, J. A. (1989). Adsorption of water on ZSM-5 zeolites. Journal of Physical Chemistry, 93, 4837. Kane, M. S., Bushong, J. H., Foley, H. C., & Brendley Jr., W. H. (1998). E ect of nanopore size distributions on trichloroethylene adsorption and desorption on carbogenic adsorbents. Industrial and Engineering Chemistry Research, 37, 2416. Karge, H. G., & Nie en, W. (1991). A new method for the study of di usion and counter-di usion in zeolite. Catalogue Today, 8, 451. Kumar, R. (1987). Removal of water and carbon dioxide from atmospheric air. US Patent 4,711,645. Lee, D. H., & Condrate Sr., R. A. (1995). An FTIR spectral investigation of the structural species found on alumina surfaces. Materials Letters, 23, 241. Little, L. H. (1966). Infrared spectra of adsorbed species.. London: Academic Press. Matsumara, Y., Yamabe, K., & Takahashi, H. (1985). The e ects of hydrophilic structures of active carbon on the adsorption of benzene and methanol vapors. Carbon, 23, 263. Masuda, T., Tsutsomi, K., & Takahashi, H. (1980). Infrared and calorimetric studies of adsorbed carbon dioxide on NaA and CaNaA zeolites. Journal of Colloid and Interface Sciences, 77, 232. Mawhiney, D. B., Rossin, J. A., Gerhart, K., & Yates Jr., J. T. (1999). Adsorption studies by transmission IR spectroscopy: a new method for opaque materials. Langmuir, 15, 4617. Meunier, F. (2000). Private communication. Myers, A. L. (1983). Activity coe cients of mixtures adsorbed on heterogeneous surfaces. A.I.Ch.E. Journal, 29, 691. Myers, A. L., & Prausnitz, J. M. (1965). Thermodynamics of mixed-gas adsorption. A.I.Ch.E. Journal, 11, 121. Parker, L. M., Bibby, D. M., & Burns, G. R. (1993). An infrared study of H2 O and D2 O on HZSM-5 and DZSM-5. Zeolites, 13, 107. Parkyns, N. D. (1971). The in uence of thermal pretreatment on the infrared spectrum of carbon dioxide adsorbed on alumina. Journal of Physical Chemistry, 75, 526.

Вам также может понравиться