Sei sulla pagina 1di 8

Adsorption Hysteresis Dynamics

Illam Park Chemistry & Biochemistry, UCLA, CA, USA illam@chem.ucla.edu


Abstract Adsorption hysteresis and its related problems are considered with respect to adsorption dynamics, using a nonlinear, nonisothermal, dynamical adsorption system. The system successfully reproduces hysteretic isotherms of types H1 and H2. Type H1 is related to nonlinearity of the adsorption isotherm while type H2 is related to limited diffusion by pore fluid. The disappearance of hysteresis, the initial variability before the cyclic steady state, the scanning loops, and the effect of temperature on hysteresis are also demonstrated and explained.

1. Introduction
Hysteresis occurs diversely in gas adsorption and mercury porosimetry, which are used for characterization of adsorbents [1]. The occurrence of hysteresis is also common in many other important scientific areas such as magnetism (p. 529, [2]) and deformation of materials under stresses (p. 175, [3]). The term hysteresis, which is from Greek meaning shortcoming or lagging behind, refers to a dynamical system of which the state does not reversibly follow changes in an external parameter. In contrast to the original meaning, however, adsorption hysteresis has been regarded as an equilibrium characteristic rather than a shortcoming or incompleteness of equilibrium. In this paper, I argue, for the first time in the area of adsorption, that the original meaning of hysteresis should be preserved for a better understanding of hysteretic adsorption systems; hysteresis should belong to part of the nonequilibrium or dynamical adsorption. The traditional concept of adsorption hysteresis is based on capillary condensation theory. In this theory, filling of a capillary pore by adsorption is by a metastable layering process until the cylindrical core of the layer faces instability and leads to complete filling of the pore volume. In contrast, emptying the pore by desorption is a stable process associated with a hemispherical meniscus. Assuming that the metastable states in capillary tubes are unique and time-independent, hysteresis is represented by a bistable equilibrium in that the state is switched by the course of adsorption or desorption. However, this concept poorly links to the real problems such as the disappearance of hysteresis in well-defined capillary tubes (e.g., [4], [5]) and the cyclic dependence or initial variability of hysteresis (e.g., p. 233, p184, [1]; p. 1077, [6]). The purpose of this paper is to elucidate the origin of adsorption hysteresis and its related problems listed above, using a simulated dynamical system. In doing so, we attempt to characterize this well-defined system in a way that we do experimentally for real systems. We observe various hysteresis loops, and find clues to the disappearance of hysteresis, the initial variability of hysteresis, the origin of hysteresis types H1 and H2, the subtleties involved in the scanning loops, and the effect of temperature on hysteresis.

2. Theory and Computer Simulation


We can derive a simulated dynamical system from the fundamental rules of mass and heat transfer (e.g., [7]; chap. 6, [10]). The system considered here is an open, nonisothermal system undergoing adsorption of a single species. It consists of bulk gas, pore fluid, and pore window. The pore fluid is a mixture of pore gas and liquid, which equilibrates instantly. In contrast, the pore gas and bulk gas diffuse across the pore window. The opening of pore window limits the gas diffusion and it depends on loading of the pore liquid. Any change in equilibrium of the pore fluid results in thermal interaction across the pore window by the heat of adsorption. The detailed mathematical components of the system are listed below. First, the local equilibrium of pore fluid is represented by the Dubinin-Astakhov (D-A) equation that is adequate for the nonlinear and nonisothermal nature of vapor adsorption [7], (1) xi = x 0i exp[-( Ai / E i ) ni ] where the subscript >i= denotes the properties of i-th pore while x, x0, A, E and n represent the volume occupied by pore liquid per adsorbent volume, pore volume per adsorbent volume, the adsorption potential of pore gas, the characteristic energy of the pore and the distribution factor for the potential energies over the characteristic energy, respectively. Adsorption potential Ai is a function of vapor pressure and temperature defined by (2) Ai = - RT ln( pi / p s (T )) where R is the gas constant, T temperature, pi the vapor pressure of pore-gas, and ps the saturation vapor pressure which is a function of temperature predicted by the Antoine equation . Second, a differential mass balance over the system with respect to time and using the linear driving force approximation for the rate of diffusion across the pore window may lead to dxi dt = k i g i ( p - pi ( xi, T )) (3) where ki is the linear mass transfer coefficient for xi60, p the vapor pressure of bulk gas, and gi the opening of pore window or accessibility. gi is inversely proportional to the amount of pore liquid by (4) g i = (1 - xi / x 0i ) exp[-( xi / xci ) mi ] where xci is the critical loading of the pore liquid for the threshold of accessibility and mi controls the abruptness of the threshold behavior. Third, a temporal energy balance on the system with approximation of the thermal film resistance between adsorbent material and bulk gas yields dT dt = Qi dxi dt - H (T - T 0) (5) where Qi and H are lumped parameters defined by qiCLVi/mcp and (hAh)/mcp, respectively. Here, qi denotes the isosteric heat of adsorption, CL the liquid density of the adsorptive, Vi the saturation volume of pore-liquid for pore i, m the total mass of the adsorbent and its content, cp the average heat capacity of mass m, h the linear heat transfer coefficient, and Ah the effective heat transfer area. The isosteric heat of adsorption can be related to the state of the pore fluid by qi=Ai + where denotes the heat of vaporization which can be determined by !Rd(lnps)/d(1/T) [7]. The coupled, nonlinear differential equations (3) and (5) were solved numerically for xi(t) and T(t) at constant p and T0. The numerical solver used is subroutine IVPRK of the IMSL/MATH library based on the Runge-Kutta-Verner fifth-order and sixth-order algorithm. For a linear, isothermal system xi=Kipi (Ki is the linear coefficient), the solution reduces to = exp(-t/i ) (6)

where denotes the deviation of transition state from equilibrium state defined by (xi(t)xi(4))/(xi(0)-xi(4)), and , the characteristic time of the dynamical system defined by Ki/kigi. Table I Simulation conditions and parameters. Temperature: 298-323 (K) Relative pressure: 0.1 to 1.0 D-A equation: x01=0.2756, x02=0.0731, n1=n2=2.0, E1=1.722 (KJ/mol), E2=10.930(KJ/mol) Kinetic parameters: xC1/x01=0.6-1, xC2/x02=100, mC1=4, mC2=1, k1=k2=1.0

3. Results and Discussion


The dynamical system simulates the 0.35 305 Amount adsorbed Temperature adsorption of water vapor on silica gel or porous glass, which shows a type IV 0.30 300 isotherm and a type H1 or H2 hysteresis loop ([7], [8], [3]). The parameters required 0.25 295 for the system are listed in Table 1. We will explore cyclic adsorption and desorption 0.20 isotherms of the system with various 290 equilibration times and diffusion limits. 0 500 1000 1500 2000 2500 3000 Note that the amount of adsorbed x in the Time, s isotherm plots is defined by x = xi/x0i. Figure 1 Typical equilibration processes. Figure 1 shows typical equilibration processes in terms of uptake versus time. Initially, we let the system equilibrate at p/ps=0.2 and 298 K. As we shift p/ps from 0.2 to 0.4 at t=0, the uptake increases sharply at the beginning and then asymptotically. The temperature, on the other hand, rises abruptly and then cools down to the isotherm temperature. We could observe a reversed behavior by resetting of p/ps to 0.2 at t=1500 s. Such equilibration processes are typical for many open systems (e.g., p. 198, [10]; [7]). we can collect isotherm points from the transition curves by imposing finite time of equilibration as we 500 s experiment real systems. This time is referred to here 2000 s 1.0 as equilibration time te. 5000 s g
Amount adsorbed, x
Amount adsorbed, x
true isotherm

Adsorption

Desorption

Temperature, K

f 3.1 Birth of hysteresis by insufficient equilibration c b Figure 2 shows the apparent isotherms for te=500, 0.6 2000, and 5000s, respectively. Each isotherm is of seven cycles of ascending and descending pressure 0.4 h e changes between p/ps=0.1 and 1, starting from d equilibrium at p/ps=0.1. Apparently, all the isotherms 0.2 a are hysteretic in a different way but exhibit some of the common features of hysteresis loops including (i) 0.0 0.0 0.2 0.4 0.6 0.8 1.0 the ascending path underlying the descending one, Relative pressure, p/pSAT (ii) the thermodynamically stable paths (the uptake Figure 2 Isotherm versus equilibration time. proportional to the pressure), and (iii) the cyclically reproducible loop. We will discuss each isotherm both in equilibrium (thermodynamics) and

0.8

nonequilibrium aspects. For the isotherm of te=500s (indicated by circles in figure 2) we can expect that nonequilibrium would prevail compared to other two isotherms. The ascending path (ab) of the first cycle is thermodynamically stable since the uptake increases with vapor pressure. However, it shows a significant shortcoming of equilibrium relative to the true equilibrium (dashed line). The system is subsaturated due to insufficient equilibration. Adsorption continues even in the descending path (bcd) until the system reaches equilibrium (point c, the maximum uptake for the cycle) at p/ps=0.68. Path bc is unstable because the path is a decreasing function of vapor pressure. Delayed desorption starts in the following path (cd) which is stable. The system is supersaturated and fails to return to a. So, the first cycle satisfies (i) only. The second cycle starts with the ascending path which is initially unstable (path de) due to prolonged desorption from the supersaturated state in the previous step. The uptake decreases until it reaches equilibrium at p/ps=0.31 (point e, the minimum uptake for the cycle). Then, the new ascending path (ef) lies between the first ascending path and the true isotherm and eventually merges with the path of the first cycle. The resultant loop (d-e-f-g-h) is retraced by the third and higher cycles. In other word, the cyclic path reaches a cyclic steady state. Accordingly, the second or higher cycle exhibit (i) and (iii). In contrast, the isotherms of te=2000 and 5000s (squares and triangles, respectively) exhibit all the features (i) through (iii). Unlike the case of te=500s, they show a cyclic steady state in the first cycle and show no instability. The loop area decreases with increasing te and shows the type H1 behavior. 3.2 Evolution of hysteresis with equilibration time D E Figure 3 shows how hysteresis evolves with 1.0 equilibration time. The area of cyclic steady loops versus equilibration time between te=0 to 0.8 C F 15,000 s. The loop area indicates a functional 0.6 relationship with te. The isotherms are also B G shown near the respective data points. As te 0.4 increases, the area increases sharply until it H I reaches the peak at te=500 s, and then A 0.2 decreases exponentially. We can see the two distinctive behaviors of isotherms across the 0.0 -1 0 1 2 3 4 5 peak. Those isotherms (A through D) on the Equilibration Time, hr left side of the peak indicate that they take Figure 3 Loop area versus equilibration time. more than one cycle to reach a cyclic steady loop. In addition, they are thermodynamically unstable since a part of the isotherm curve decreases as the pressure increases. As te increases, the initial variability of the loop diminishes while the area increases. In contrast, isotherms E through I are stable and show no initial variability. They strikingly resemble hysteretic isotherms from experiments. As te increases, the loop area decays exponentially but very slowly, especially near the total extinction. In other words, hysteresis near the true equilibrium may be stable enough to be reproducible by experiment. The time-dependent, nonequilibrium aspect of hysteresis, however, contradicts the thermodynamic based-theories in that hysteresis is permanent or time-invariant (Everett 1967). Everett (p. 1055, 1967) has introduced the terms permanent hysteresis and time-dependent hysteresis. In his theory, adsorption hysteresis is permanent because of its stability and
Area of Cyclic Steady Loop

reproducibility for a significantly long time. The time-dependent cases not considered as adsorption hysteresis were (i) a steady drift of the dependent variable upon oscillation of the independent variable between two boundaries; (ii) a hysteresis loop degenerating (or shrinking) into a line as the frequency of oscillation decreases. (The frequency of oscillation is inversely proportional to the equilibration time). Case (i) describes the behavior of isotherm A in figure 3. On the other hand, case (ii) describes the decay of hysteresis shown by isotherms E-I. Accordingly, all the loops from the present dynamical system should not be considered as adsorption hysteresis. However, a slowly evolving hysteresis may be assumed permanent. In fact, equilibration can take place very slowly due to two major factors, nonlinear equilibrium and limited diffusion by pore blocking. 3.3 Nonlinear equilibrium: Hysteresis loop of type H1 Nonlinear equilibrium can cause multiplicity of equilibration times. Equilibration in steep regions of the isotherm may take substantially long. For instance, in [4]), nitrogen adsorption of MCM-41 equilibrated within 20 min, apart from the steep middle region of the isotherm where 2 hr was required. Also, [11] reported that in the nitrogen adsorption of aggregated nonporous silica spheres, equilibration lasted several hours in the steep region of their hysteretic isotherm while it was achieved quickly within 1 hr in the other region. Physically, the slope or first derivative of the isotherm curve represents the adsorption capacity or the pore volume for its respective vapor pressure. It will take longer to fill or empty a pore with a larger volume. We can explain the relationship and equilibration time, quantitatively. Consider a nonlinear isotherm curve consisting of two distinctive slopes K1 and K2, and let K2/K1=4. Here, the subscripts 1 and 2 denote two different regions. From Eq. (6), we know Ki scales time, t, assuming kigi is constant for two regions. For instance, let equilibration time te = 41. Then, Eq. (6) will give 0.18 for region 1. This is, the isotherm measured for that region will represent the true equilibrium within 1.8%. The isotherm for region 2, however, will depart from the true one by 36%. Accordingly, a cyclic isotherm for the entire region will have a hysteresis loop in region 2. In cases E -I in figure 3, notice that the hysteresis loop is distributed in the steep region of the 1.0 true isotherm. In conclusion, hysteresis is probable in 0.8 the systems with nonlinearity in equilibrium. Such systems have two or more equilibration time scales depending on the 0.6 vapor pressure. A hysteresis loop may appear around the steep region of the true 0.4 Equilibrium equilibrium, and its ascending and Xc = 1.0 descending branches depart evenly from the Xc = 0.8 0.2 true isotherm (see figure 3). This behavior Xc = 0.7 resembles type H1 hysteresis. Xc = 0.6

Amount adsorbed, x

3.4 Limited diffusion by pore liquid: 0.0 0.2 0.4 0.6 0.8 1.0 Hysteresis loop of type H2 Relative Pressure, p/pSAT Equilibration takes longer if diffusion is Figure 4 Transition of hysteresis from H1 to H2 limited due to blocking of the pore window by pore liquid. This effect is represented by the critical loading xci in Eq. (4). Figure 5 shows a

0.0

series of hysteresis loops for xc2=1, 0.8, 0.7, and 0.6, respectively. The shape of the loop changes from type H1 to H2 as xc2 decreases. The loop of xc2=0.6 shows the feature of type H2: the descending branch of the loop shows a pronounced plateau followed by a steep fall. This behavior agrees with experimental isotherms of powder compacts compressed at gradually elevated pressures (p. 102, Fig. 2.30; p. 155, Fig. 3.22; p. 205, Fig. 4.5; in [1]). The uncompressed powder shows an isotherm of type II with no hysteresis. As compaction increases, a hysteresis loop of type H1 develops, and then evolves into type H2. Meanwhile, the type II shape of the isotherm evolves into a type IV. The extent of uptake at the saturation vapor pressure becomes lower. In sum, the type H2 behavior may be described as follows. The ascending path deviates further from the true isotherm at the onset of the critical uptake where diffusion becomes seriously limited. Since adsorption takes place minimally, the uptake never reaches the possible maximum at the saturation vapor pressure. Desorption delayed until the path crosses the true equilibrium, starts very slowly at the beginning because of the limited diffusion. As the uptake reaches the critical loading, the trapped (supersaturated) pore liquid diffuses away quickly to equilibrate with the bulk gas. This behavior agrees with the pore-blocking theory by Everett (p. 150, [1]) in that evaporation from the pore is assumed to delay until the pore neck is cleared. However, the pore-blocking theory assumes complete filling of pores across the pore neck for the course of adsorption. Lack of the transport barrier in the pore-filling process is the major difference from the present theory.
Scanning Curves of Type H1
1.0 0.8 0.6 0.4 0.2 0.0

Scanning Curves of Type H2


Amount adsorbed, x
1.0 0.8 0.6 0.4 0.2 0.0

Desorption scanning 4 3 2 1

Desorption scanning 1 3 2

Amount adsorbed, x

5 6 8 7

5 6 7 9 8

Amount adsorbed, x

Amount adsorbed, x

1.0 0.8 0.6 0.4 0.2 0.0 0.0

X Data Adsorption scanning

7 6 5 4 1 2 3

1.0 0.8 0.6 0.4 0.2 0.0 0.0

Adsorption scanning 4 3 5 6 7 8 9

1 2
0.2 0.4 0.6 0.8 1.0

0.2

0.4

0.6

0.8

1.0

Figure 5 Scanning curves of type H1 hysteresis.

Relative pressure, p/pSAT

Relative pressure, p/pSAT

Figure 6 Scanning curves of type H2 hysteresis.

3.5 Scanning curves Scanning curves are cyclic paths in a narrow range of vapor pressure, in contrast to boundary (B)

curves in the full range of vapor pressure. Figures 5 and 6 show scanning curves for type H1 and H2 hysteresis loops, respectively. Nine consecutive cycles with various lower or upper pressure limits were examined. The upper or lower pressure limits are labeled with the number of the cycle. The first cyclic path will be considered as a boundary curve. Descending Scanning (DS) and ascending scanning (AS) curves for type H1 hysteresis (from figure 4a) are shown in figures 5a and 5b, respectively. All the DS curves behave similarly in the following way. They follow a horizontal path toward the equilibrium isotherm as the vapor pressure decreases. After crossing the equilibrium curve, they fall and merge with the upper B-curve. The AS-curves in Fig. 6b behave similarly but in the reverse direction. As the vapor pressure increases at a point on the upper B-curve, the path switches its direction horizontally toward the equilibrium curve, and then rises to merge with the lower B-curve. The scanning curves of type H2 in figure 6 show a more complex behavior. Of the DS-curves in figure 6a, curves 2 and 3 resemble the upper B-curve 1. They fall directly onto the lower B-curve (actually onto the true isotherm) rather than merging with the upper B-curve. In contrast, curves 6-9 behave similarly to those of type H1 in figure 5. The rest of the curves (4, 6) indicate a transitional behavior between types H1 and H2. These contrasting behaviors indicate the significance of either nonlinearity or limited diffusion for the particular region of the isotherm. On the other hand, the AScurves in figure 6b show the prominent instability, especially near the sharp transition of the upper B-curve. This instability is caused by the substantial departure of the upper boundary curves from the true equilibrium. The behavior of scanning curves generally agrees with the experimental result presented in Fig. 36-12 of Everett (1967). 3.6 Effect of temperature on hysteresis The effect of temperature on adsorption 1.0 hysteresis has been one of the issues discussed recently. It has been reported that 0.8 the size of hysteresis loop gradually 0.6 decreases as temperature increases [7,12]. This experimental behavior agrees with the 0.4 298 K result shown in figure 7. The true isotherms 310.5 K 0.2 at 298, 310.5 indicate that there is no 323 K noticeable difference although the amount 0.0 0.0 0.2 0.4 0.6 0.8 1.0 of uptake slightly increases as temperature Relative pressure, p/pSAT increases. However, the hysteresis loops of type H2 apparently show that the area of Figure 7 Effect of temperature on hysteresis (H2) hysteresis loop decreases with increasing temperature. The major change resides in the steep region of the desorption curve near the point of critical loading. Interestingly, a close look at the region near the saturation vapor pressure will reveal that the uptake increases with increasing vapor pressure. This behavior is against the thermodynamic nature of adsorption. However, since the system is subsaturated and the diffusion is limited in that region, an increased temperature will increase the diffusivity of molecules and therefore will help the system moving toward the true equilibrium. In other words, kinetic effects may dominate in the diffusion-limited region.

4. Conclusions

Amount adsorbed, x

Adsorption hysteresis on the phase plane (isotherm) of uptake versus vapor pressure has been investigated using a simulated dynamical adsorption system, which is nonlinear and nonisothermal in nature. The simulation has successfully reproduced the hysteresis and its related phenomena often found in experiments, and elucidated the underlying mechanisms straightforwardly in the aspect of adsorption dynamics. The major findings of the present study are listed below. First, hysteresis may be attributed to insufficient equilibration in the measurement of isotherm points. The hysteresis loop, which is often claimed experimentally reproducible, may be due to a cyclic steady state of premature equilibrium points. Second, the hysteresis loop is not necessarily a unique, equilibrium property. It can be diverse in shape and extent depending on equilibration time. A relatively shorter equilibration may cause the initial variability of the hysteresis loop [1] that are observed commonly in both adsorption and mercury porosimetry. In contrast, a longer equilibration gives rise to a direct formation of the stable loop, but the loop gradually shrinks and disappears as the equilibration time increases further. Finally, the present model suggested two major factors for the multiplicity of time scales. The one is nonlinearity of the true isotherm and the other is limited diffusion by the adsorbed phase or pore-liquid. References [1] S.J. Gregg, K.S.W. Sing, Adsorption, surface area and porosity, Academic Press, New York 1982 [2] R. G. Lerner, G. L. Trigg, Encyclopedia of Physics, VCH Publishers, New York, 1991. [3] M. Brokate, J. Sprekels, Hysteresis and Phase Transitions, Springer-Verlag, New York, 1996. [4] P.J. Branton, P.G. Hall, K.S.W. Sing, H. Reichert, F. Schth, and K.K. Unger, Physisorption of Argon, Nitrogen, Oxygen by Mcm-41, a Model Mesoporous Adsorbent, J. Chem. Soc. Faraday Trans., 90 (19), (1994) 2965-2967. [5] P. J. Branton, K. Kaneko, N. Setoyama, K. W. Sing, S. Inagaki, Y. Fukusima, Physisorption of Nitrogen by Mesoporous Modified Kanemite, Langmuir, 12, (1996) 599-600. [6] D.H.Everett, Adsorption Hysteresis, The Solid-Gas Interface, Vol. 2, E.A. Flood (Ed.), pp. 10551113, Marcel Dekker, Inc., New York, 1967. [7] I. Park, K.S. Knaebel, Adsorption Breakthrough Behavior: Unusual Effects and Possible Causes, AIChE J., 38 (5), (1992) 660-670. [8] P. Rajniak, R. T. Yang, A Simple Model and Experiments for Adsorption-desorption Hysteresis: Water Vapor on Silica Gel, AIChE J., 39 (5), (1993) 774-786. [9] P. Porion, A.M. Faugre, P. Levitz, H. van Damme, A. Raoof, J.P. Guilbaud, F. Chevoir, A NMR Investigation of Adsorption/desorption Hysteresis in Porous Silica Gels, Magnetic Resonance Imaging, 16 (5/6), (1998) 679-682. [10] D. M. Ruthven, Principles of Adsorption and Adsorption Processes, John Wiley & Son, New York, 1984 [11] H. Giesche, K.K. Unger, U. Mller, U. Esser, Hysteresis in Nitrogen Sorption and Mercury Porosimetry on Mesoporous Model Adsorbents Made of Aggregated Monodisperse Silica Spheres, Colloids and Surfaces, 37, (1989) 93-113. [12] W. D. Machin, Temperature Dependence of Hysteresis and the Pore Size Distributions of Two Mesoporous Adsorbents, Langmuir, 10, (1994) 1235-1240.

Potrebbero piacerti anche