Capillary Phase Transitions of Linear and Branched Alkanes in

Jul 13, 2006 - In Opticks,1 Sir Isaac Newton recorded the rise of a liquid in a capillary tube. The Kelvin equation proposed in the early 1870s,2,3 wh...
0 downloads 0 Views 238KB Size
Langmuir 2006, 22, 7391-7399

7391

Capillary Phase Transitions of Linear and Branched Alkanes in Carbon Nanotubes from Molecular Simulation Jianwen Jiang*,† and Stanley I. Sandler‡ Department of Chemical & Biomolecular Engineering, National UniVersity of Singapore, 4 Engineering DriVe 4, Singapore 117576, and Department of Chemical Engineering, UniVersity of Delaware, Newark, Delaware 19716 ReceiVed March 31, 2006. In Final Form: May 23, 2006 Capillary phase transitions of linear (from C1 to C12) and branched (C5 isomers) alkanes in single-walled carbon nanotubes have been investigated using the gauge-cell Monte Carlo simulation. The isotherm at a supercritical temperature increases monotonically with chemical potential and coincides with that from the traditional grand canonical Monte Carlo simulation, whereas the isotherm at a subcritical temperature exhibits a sigmoid van der Waals loop including stable, metastable, and unstable regions. Along this loop, the coexisting phases are determined using an Maxwell equal-area construction. A generic confinement effect is found that reduces the saturation chemical potential, lowers the critical temperature, increases the critical density, and shrinks the phase envelope. The effect is greater in a smaller diameter nanotube and is greater in a nanotube than in a nanoslit.

1. Introduction Fluids in confined spaces show phase transitions that differ significantly from bulk fluids as a consequence of the spatial confinement and the interaction with the substrate. Understanding confined phase transitions is not only of interest in the physical and surface sciences but of practical importance for a variety of applications. Historically, the study of fluids in a confined domain dates back to the 18th century. In Opticks,1 Sir Isaac Newton recorded the rise of a liquid in a capillary tube. The Kelvin equation proposed in the early 1870s,2,3 which relates the vapor pressure of a liquid to the curvature of its meniscus, was first used by Zsigmondy4 in 1911 to elucidate capillary condensation. The Kelvin equation is generally applied to infer a tube radius from low-temperature isotherms. However, the Kelvin equation is not accurate in small pores.5,6 Despite the attempts to extend the applicability of Kelvin equation,7 the macroscopic nature prevents its use in nanosized pores. In the past a few decades, remarkable advances have been achieved in elucidating capillary phase transitions from molecular aspect by statistical mechanics methods. A large number of studies have used the density functional theory (DFT). Evans, Tarazona, and Marconi8-12 investigated the capillary condensation of simple fluids in cylindrical and slit pores, and the prewetting transition * Corresponding author. E-mail: [email protected]. Tel: (65) 6516 5083. Fax: (65) 6779 1936. † National University of Singapore. ‡ University of Delaware. (1) Newton, I. Opticks, or, a Treatise of the Reflections, Refractions, IrMlections and Colours of Light; London, 1730. (2) Thomson, W. Proc. R. Soc. Edinburgh 1870, 7, 63. (3) Thomson, W. Proc. R. Soc. Edinburgh 1871, 42, 448. (4) Zsigmondy, R. Z. Anorg. Allg. Chem. 1911, 71, 356. (5) Fisher, L. R.; Israelachvili, J. N. Nature 1979, 277, 548. (6) Fisher, L. R.; Gamble, R. A.; Middlehurst, J. Nature 1981, 290, 575. (7) Gregg, S. J.; Sing, K. S. W. Adsorption, Surface Area and Porosity; Academic: London, 1982. (8) Evans, R.; Tarazona, P. Phys. ReV. Lett. 1984, 52, 557. (9) Evans, R.; Marconi, U.; Tarazona, P. J. Chem. Phys. 1985, 84, 2376. (10) Evans, R.; Marconi, U.; Tarazona, P. J. Chem. Soc., Faraday Trans. 2 1986, 82, 1763. (11) Tarazona, P.; Marconi, U.; Evans, R. Mol. Phys. 1987, 60, 573. (12) Evans, R. J. Phys. Condens. Matter 1990, 2, 8989.

on a single planar wall. Henderson and Sokolowski13 examined the phase behavior of a Lennard-Jones (LJ) fluid confined in spherical cavities and found a shifted transition toward lower density in a smaller cavity. Zgorski et al.14 calculated the phase behavior of a LJ fluid in a slitlike pore and found that the change in the interaction with surface can lead to a substantial change in the phase behavior compared to the bulk fluid. Salinger and Frink15,16 presented three algorithms for the rapid analysis of phase behavior of confined fluids, with a detailed case study in disordered porous media. Patrykiwjew et al.17 studied the confinement effect of a slit pore on the closed loop immiscibility in symmetric binary model mixtures. Pizio et al.18 considered the vapor-liquid equilibria (VLE) of the restricted primitive model for ionic fluids within charged and uncharged slit pores. Li et al.19 examined the layering, condensation, and evaporation of short LJ chains in narrow slits. Bryk et al.20 investigated the capillary condensation of LJ chains confined in slit pores and found anomalous behavior of the critical temperature for long chains. Other theoretical methods have also been used to study the phase transitions in confined media. One example is the replica Ornstein-Zernike integral equation developed by Given and Stell.21,22 With this, Pitard et al.23 predicted the VLE of a lattice gas in disordered porous media, Paschinger et al.24,25 studied the phase separation of a symmetric binary fluid in a porous matrix. Computer simulations have been widely used to investigate the behavior of confined fluids. With ever-increasing computational power, simulation can provide an improved and detailed (13) Henserson, D.; Sokolowski, S. Phys. ReV. E 1995, 52, 758. (14) Zagorski, R.; Patrykiejew, A.; Sokolowski, S. J. Phys. Condens. Matter 2002, 14, 165. (15) Salinger, A. G.; Frink, L. J. D. J. Chem. Phys. 2003, 118, 7457. (16) Frink, L. J. D.; Salinger, A. G. J. Chem. Phys. 2003, 118, 7466. (17) Patrykiejew, A.; Pizio, O.; Pusztai, L.; Sokolowski, S. Mol. Phys. 2003, 101, 2219. (18) Pizio, O.; Patrykiejew, A.; Sokolowski, S. J. Chem. Phys. 2004, 121, 11957. (19) Li, Z. D.; Cao, D. P.; Wu, J. Z. J. Chem. Phys. 2005, 122, 224701. (20) Bryk, P.; Pizio, O.; Sokolowski, S. J. Chem. Phys. 2005, 122, 194904. (21) Given, J. A.; Stell, G. Phys. ReV. A 1992, 45, 816. (22) Given, J. A.; Stell, G. J. Chem. Phys. 1992, 96, 2287. (23) Pitard, E.; Rosinberg, M. L.; Tarjus, G. Mol. Simul. 1996, 17, 399. (24) Paschinger, E.; Kahl, G. Phys. ReV. E 2000, 61, 5330. (25) Paschinger, E.; Levesque, D.; Weis, J.-J. Europhys. Lett. 2001, 55, 178.

10.1021/la0608720 CCC: $33.50 © 2006 American Chemical Society Published on Web 07/13/2006

7392 Langmuir, Vol. 22, No. 17, 2006

microscopic picture of a fluid in the interaction field of a confined space and shed light on the underling physics. Grand canonical Monte Carlo (GCMC) simulation is the most commonly used method to explore adsorption and capillary phenomena in nanoscaled media.26 GCMC permits fluid molecule exchange with the reservoir at a given chemical potential, which mimics realistic processes. Using GCMC, Peterson et al.27 examined the isotherms, phase diagrams, and density profiles of a LJ fluid in cylindrical pores. Kierlik et al.28 considered the liquid-liquid equilibria (LLE) of a LJ fluid mixture in a slit pore. Monson and co-workers29-31 located the coexistence curve of a LJ fluid in a xerogel of spherical silica particles. Gelb and Gubbins32 investigated the adsorption/evaporation process of Xe and N2 in dynamically quenched porous Vycor glasses. Pellenq and coworkers33,34 studied the capillary condensation of rare gases in disordered mesoporous silica media of different morphologies and topologies generated by an off-lattice 3D reconstruction method. Gatica and Cole35 simulated the capillary condensation of Ar inside infinitely long cylindrical pores with various gassurface interaction potentials, and discussed the relation to wetting on planar surfaces. In some of the above-mentioned GCMC simulations, the coexisting phases were located using a thermodynamic integration method.36 Alternatively, a histogram-reweighting technique37 has been implemented with GCMC simulation to calculate phase transitions. Alvarez et al.38 studied the VLE of a quenchedannealed system. Potoff and Siepmann39,40 analyzed the effect of branching on the phase equilibria and critical parameters for alkane monolayers on a flat gold surface. Shi et al.41 demonstrated capillary condensation, prewetting, and layering transitions for different systems on surfaces. Grandis et al.42 presented a study of the phase diagram of a LJ fluid adsorbed in a fractal and highly porous aerogel generated from an off-lattice diffusionlimited cluster-cluster aggregation process. Hehmeyer et al.43 examined the phase transitions and structures of lattice homopolymers of up to 1024 beads in strictly 2D and quasi-2D geometries. An excellent method to simulate phase transitions in bulk fluids is Gibbs ensemble MC (GEMC) simulation developed by Panagiotopoulos,44 in which the coexisting phases are determined directly using two separated simulation boxes. GEMC has also been used for phase transitions in confined space. Panagiotopoulos45 first investigated the capillary condensation of fluids in cylindrical pores. Siepmann et al.46 studied the VLE coexistence in a Langmuir monolayer of pentadecanoic acid. Jiang and (26) Nicholson, D.; Parasonage, N. G. Computer Simulation and the Statistical Mechanics of Adsorption; Academic Press: New York, 1982. (27) Peterson, B. K.; Gubbins, K. E.; Heffelfinger, G. S.; Marconi, U.; Swol, F. v. J. Chem. Phys. 1988, 88, 6487. (28) Kierlik, E.; Fan, Y.; Monson, P. A.; Rosinberg, M. L. J. Chem. Phys. 1995, 102, 3712. (29) Page, K. S.; Monson, P. A. Phys. ReV. E 1996, 54, R29. (30) Page, K. S.; Monson, P. A. Phys. ReV. E 1996, 54, 6557. (31) Sarisov, L.; Monson, P. A. Langmuir 2000, 16, 9857. (32) Gelb, L. D.; Gubbins, K. E. Langmuir 1998, 14, 2097. (33) Pellenq, R. J.-M.; Levitz, P. E. Mol. Phys. 2002, 100, 2059. (34) Coasne, B.; Pellenq, R. J.-M. J. Chem. Phys. 2004, 121, 3767. (35) Gatica, S. M.; Cole, M. W. Phys. ReV. E 2005, 72, 041602. (36) Peterson, B. K.; Gubbins, K. E. Mol. Phys. 1987, 62, 215. (37) Ferrenberg, A. M.; Swendsen, R. W. Phys. ReV. Lett. 1988, 61, 2635. (38) Alvarez, M.; Levesque, D.; Weis, J.-J. Phys. ReV. E 1999, 60, 5495. (39) Potoff, J. J.; Siepmann, J. I. Phys. ReV. Lett. 2000, 85, 3460. (40) Potoff, J. J.; Siepmann, J. I. Langmuir 2002, 18, 6088. (41) Shi, W.; Zhao, X. C.; Johnson, J. K. Mol. Phys. 2002, 100, 2139. (42) Grandis, V. D.; Gallo, P.; Rovere, M. Phys. ReV. E 2004, 70, 061505. (43) Hehmeyer, O.; Ayra, G.; Panagiotopoulos, A. Z. J. Phys. Chem. B 2004, 108, 6809. (44) Panagiotopoulos, A. Z. Mol. Phys. 1987, 61, 813. (45) Panagiotopoulos, A. Z. Mol. Phys. 1987, 62, 701. (46) Siepmann, J. I.; Karaborni, S.; Klein, M. L. J. Phys. Chem. 1994, 98, 6675.

Jiang and Sandler

Gubbins47 studied the VLE of 2D LJ films, and good agreement with experimental data was obtained. Dijkstra48 simulated the VLE of normal pentane between two solid plates and found a shift of the critical temperature to lower temperatures with decreasing plate separation. Vishnyakov et al.49 considered the critical properties of LJ fluids in narrow slit-shaped pores and found a linear dependence of the critical temperature with inverse pore width. Brennan and Dong50,51 simulated the phase transitions of one-component fluid and binary fluid mixtures in random porous media and analyzed in detail the rich phase behavior. Brovchenko et al.52,53 determined the coexistence curves of H2O in cylindrical and slit nanopores and observed a significant change in the critical parameter and the shape of coexistence curve changing with H2O-substrate interactions. Puibasset54,55 examined the capillary condensation of a LJ fluid in heterogeneous pores and suggested that pore disorder has a strong influence on the properties of confined fluid. Other simulation methods have also been used. Hoffmann and Nielaba56 studied the condensation of a LJ fluid in cylindrical pores by path integral MC simulation. Dukovski et al.57 examined the coexistence curve and critical point of benzene using a tworeplica cluster MC algorithm. By means of molecular dynamics simulation, Heinz et al.58 reported the phase transitions and structures of alkyl chains on alkylammonium micas. Koga and Tanak59 presented the phase transitions among liquid, bilayer, and monolayer ice phases between hydrophobic slit surfaces. Recently, the gauge-cell MC simulation method has been developed by Neimark and Vishnyakov60,61 to investigate the phase transitions of Ar and N2 in smooth nanopores. This method can be used to simulate the complete isotherm, including stable, metastable, and unstable regions, and as a result, the coexisting phases can be determined easily. The key feature is that thermodynamically unstable states, which cannot be observed in experiment, are stabilized in simulation by suppressing density fluctuations. Gauge-cell MC simulation has been adapted in a few studies of capillary phase transitions. Seaton and coworkers62,63 investigated the phase coexistence of H2O and CH4 in slit pores. Jiang et al.64 simulated the capillary transition of short linear alkanes (from C1 to C4) in a single-walled carbon nanotube (SWNT). Kowalczyk et al.65 examined the condensation/evaporation pressures of N2 in SWNTs and sequentially evaluated the distribution of carbon nanotube sizes in a sample. In this work, our previous study64 has been extended to long linear (up to C12) and branched (C5 isomers) alkanes in three SWNTs using the gauge-cell MC. Currently it is of intense interest (47) Jiang, S.; Gubbins, K. E. Mol. Phys. 1995, 86, 599. (48) Dijkstra, M. J. Chem. Phys. 1997, 107, 3277. (49) Vishnyakov, A.; Piotrovakaya, E. M.; Brodakaya, E. N.; Votyakov, E. V.; Tovbin, Y. K. Langmuir 2001, 17, 4451. (50) Brennan, J. K.; Dong, W. J. Chem. Phys. 2002, 116, 8948. (51) Brennan, J. K.; Dong, W. Phys. ReV. E 2003, 67, 031503. (52) Brovchenko, I.; Geiger, A.; Oleinikova, A. J. Chem. Phys. 2004, 120, 1958. (53) Brovchenko, I.; Geiger, A.; Oleinikova, A. J. Chem. Phys. 2005, 123, 044515. (54) Puibasset, J. J. Phys. Chem. B 2005, 109, 4700. (55) Puibasset, J. J. Chem. Phys. 2005, 122, 134710. (56) Hoffmann, J.; Nielaba, P. Phys. ReV. E 2003, 67, 036115. (57) Dukovski, I.; Machta, J.; Saravanan, C.; Auerbach, S. M. J. Chem. Phys. 2000, 113, 3697. (58) Heinz, H.; Castelijns, H. J.; Suter, U. W. J. Am. Chem. Soc. 2003, 125, 9500. (59) Koga, K.; Tanaka, H. J. Chem. Phys. 2005, 122, 104711. (60) Neimark, A. V.; Vishnyakov, A. Phys. ReV. E 2000, 62, 4611. (61) Vishnyakov, A.; Neimark, A. V. J. Phys. Chem. B 2001, 105, 7009. (62) Jorge, M.; Seaton, N. A. Mol. Phys. 2002, 100, 3803. (63) Jorge, M.; Schumacher, C.; Seaton, N. A. Langmuir 2002, 18, 9296. (64) Jiang, J. W.; Sandler, S. I.; Smit, B. Nano Lett. 2004, 4, 241. (65) Kowalczyk, P.; Holyst, R.; Tanaka, H.; Kaneko, K. J. Phys. Chem. B 2005, 109, 14659.

Molecular Simulations of Alkanes in CNTs

Langmuir, Vol. 22, No. 17, 2006 7393

Table 1. Force Field Parameters for Linear and Branched Alkanes nonbonded LJ68,69

site

σ (Å)

/kB (K)

CH4 CH3 CH2 CH C

3.73 3.75 3.95 4.68 6.40

148.0 98.0 46.0 10.0 0.5

3

utorsion(φ) )

bending70

kθ/kB ) 62500 K/rad2

θ 0 ) 113.0o

torsion71

CHx-CH2-CH2-CHy

CHx-CH2-CH-CHyZ

V0/kB ) 1009.728 K V1/kB ) 2018.446 K V2/kB ) 136.341 K V3/kB ) -3164.520 K

V0/kB ) 373.0512 K V1/kB ) 919.0441 K V2/kB ) 268.1541 K V3/kB ) -1737.2160 K

to determine the applicability of SWNTs as adsorbents, membranes, molecular sieves, and filtering media because of their unique internal nanometric structure. Although there has been considerable literature on fluid adsorption in carbon nanotubes, far less explored are the phase transitions. The fluids under this study, alkanes, are the basic feedstock in chemical industry, and a better understanding of their phase transitions in confined environments would be useful for their storage and in elucidating their behavior in nanoporous catalysis and elsewhere. The phase transitions of three C5 isomers in a carbon nanoslit have also been simulated here in order to investigate the effect of pore geometry. For comparison with the gauge-cell MC, GCMC was also used to determine the adsorption and desorption isotherms of C1 in a SWNT. In addition, GEMC was used to calculate the coexistence curves of bulk C5 isomers. In section II, we briefly introduce the models used for alkanes, SWNTs, and nanoslit, as well as the interaction potentials used. In section III, the gaugecell MC, GCMC, and GEMC simulation methods are described. Results are presented in section IV, including isotherms, phase coexistence curves, and fluid structures within SWNT. Finally, concluding remarks are given in section V.

2. Models Two prototypes are commonly adapted to represent alkane molecules, the united-atom model and the all-atom model.66 Both models were found to give comparable adsorption isotherms for alkane adsorption in silicalite; however, the simpler united-atom model is computationally faster.67 Consequently, the unitedatom model, representing every CHx group as a single interaction site, was used in this work. The C-C bonds were assumed to be rigid and fixed at 1.53 Å. The nonbonded dispersive interactions between sites of different molecules and four sites apart within a molecule were modeled using the Lennard-Jones (LJ) potential68,69

uLJ(r) ) 4[(σ/r) - (σ/r) ] 12

6

(1)

For C3 and longer alkanes, the intramolecular bond bending between three successive sites was modeled using a harmonic potential70

ubending(θ) ) 0.5kθ(θ - θ0)

2

For C4 and longer alkanes, the intramolecular dihedral torsion between four successive sites was modeled using a cosine potential71

(2)

∑ Vk(cos φ)k k)0

(3)

Table 1 gives the force field parameters, as optimized by others to accurately reproduce experimental vapor-liquid coexistence curves and critical properties of pure linear and branched alkanes. The cross LJ parameters between the unlike sites were obtained using the Jorgensen combining rules σij ) (σiσj)1/2 and ij ) (ij)1/2.72 Three arm-chair type (m, m) SWNTs were considered separately with m ) 30, 25, and 20. Each SWNT was openended, 30.74 Å in length, and 1.23x3 m/π in radius. By using the periodic boundary condition in the axial dimension, the SWNT was thereby considered to be infinitely long in the simulations. Note that experimentally produced SWNTs are closed-ended and need to be uncapped to obtain open-ended SWNTs. We assume that the extent of impurities introduced during this process are rather small, and well structured open-ended nanotubes can be obtained. The carbon atoms in each SWNT were taken into account fully atomistically, different from some previous studies which used simplified smooth structureless cylindrical nanopores. During the simulation, the SWNT was assumed to be rigid with carbon atoms frozen. As our study focused on low-energy equilibrium configurations, flexibility of the SWNT allowing local small movements of the carbon atoms would not be expected to give significantly different results. The nonbonded dispersion between the carbon atom in SWNT and the alkane interaction site was also modeled using the LJ potential. The well depth and collision diameter were obtained using the Jorgensen combining rule72 with C-C/kB ) 28.0 K and σC-C ) 3.4 Å. A spherical cutoff length of 15.37 Å was used for the calculation of the LJ interaction energies without additional long-range corrections. For comparison, the capillary phase transitions of alkanes in a carbon nanoslit with a width of 40.68 Å, identical to the diameter of a (30,30) SWNT, was investigated. The carbon nanoslit was modeled as two graphitic basal surfaces separated by a width w in the z dimension. The length of the slit surface in the x or y dimension was assumed to be 40 Å. The periodic boundary conditions were employed in the transverse xy plane parallel to the slit surface. The nonbonded dispersion between a slit surface and an alkane site was modeled using the Steele 10-4-3 potential73,74

[( ) ( )

usa(z) ) 2πFssaσsa2∆

2 σsa 5 z

10

-

σsa 4 z

(

)]

σsa4

3∆(0.61∆ + z)3

, (4)

where z is the distance from the slit surface, Fs ) 0.114 Å-3 is the implicit carbon number density of the slit surface, sa and σsa are the cross interaction well depth and collision diameter

(66) Ryckaert, J. P.; Bellemans, A. Faraday Discuss. Chem. Soc. 1978, 66, 95. (67) Macedonia, M. D.; Maginn, E. J. Fluid Phase Equilib. 1999, 158-160, 19. (68) Martin, M. G.; Siepmann, J. I. J. Phys. Chem. B 1998, 102, 2569. (69) Martin, M. G.; Siepmann, J. I. J. Phys. Chem. B 1999, 103, 4508. (70) van der Ploeg, P.; Berendsen, H. J. C. J. Chem. Phys. 1982, 76, 3271.

(71) Wang, Y.; Hill, K.; Harris, J. G. J. Chem. Phys. 1993, 100, 3276. (72) Jorgensen, W. L.; Madura, J. D.; Swenson, C. J. J. Am. Chem. Soc. 1984, 106, 6638. (73) Steele, W. A. Chem. ReV. 1993, 93, 2355. (74) Steele, W. A. The Interaction of Gases with Solid Surfaces; Pergamon: Oxford, 1974.

7394 Langmuir, Vol. 22, No. 17, 2006

Figure 1. C1 isotherms in a (30, 30) SWNT at 120, 140, 160, and 180 K. The open circles are from gauge-cell MC, the solid circles connected by dashed lines are the coexisting phases. The upward and downward triangles are for adsorption and desorption, respectively, from GCMC. The dot-dashed lines indicate capillary condensation and evaporation, respectively, from GCMC.

estimated from the Jorgensen combining rule,72 and ∆ ) 3.35 Å is the intersheet distance of graphite. Equation 4 was derived by integrating LJ pair potential over graphene sheets.73,74 The overall interaction potential of an alkane site within the slit pore is usa(z) + usa(w - z). As a consequence, the atomic corrugations on graphene sheets are not taken into account; however, their effects on adsorption and phase transition within the slit pore are expected to be negligible.

3. Methods The gauge-cell MC uses two simulation boxes, each of which has a fixed volume. In this work, one box was embedded with a SWNT or a nanoslit, and the other was a gauge cell of 50 × 50 × 50 (Å)3 for bulk alkane with periodic boundary conditions in all three directions. In this respect, the gauge-cell MC is an NVT ensemble. The limited capacity of the gauge cell restrains the density fluctuations in the pore and allows the fluid in the pore in stable, metastable, or unstable state. As mentioned, GCMC was also used to determine the adsorption isotherm of C1, and GEMC was used to determine the coexistence curves of bulk C5 isomers. All three types of simulations ran separately for a total 20 000 cycles, in which the first 10 000 cycles were used for equilibration and the second 10 000 cycles to obtain ensemble averages. Each cycle consisted of a number of the following attempted trial moves: (a) Translation. A randomly selected alkane molecule was translated with a random displacement in the x, y, or z direction, and the maximum displacement was adjusted to an overall acceptance ratio of 50%. (b) Rotation. A randomly selected alkane molecule was rotated around the center-of-mass with a random angle, and the maximum angle was adjusted to an overall acceptance ratio of 50%. (c) Partial regrowth. Part of a randomly selected alkane molecule was regrown. Also, it was randomly decided which part of the molecule was to be regrown and from which bead the regrowth was started. (d) Complete regrowth. A randomly selected alkane molecule was regrown completely at a random position. (e) Swap between two boxes (only in gauge-cell MC and GEMC). A randomly selected alkane molecule was swapped into the other box at a random position. This type of trial move is in analogy to the Widom test particle method, from which the

Jiang and Sandler

chemical potential of the fluid in the gauge cell was estimated simultaneously. This procedure is faster than previous gaugecell simulations,60,61 in which the chemical potential in the gauge cell was determined from a series of separated GCMC simulations. (f) Exchange with the reservoir (only in GCMC). A new alkane molecule was created at a random position, or a randomly selected alkane molecule was deleted. To ensure microscopic reversibility, creation and deletion were attempted at random with equal probability. (g) Volume change (only in GEMC). While the total volume of the two boxes was kept constant, a small volume change in either box was performed, and the maximum change was adjusted to an overall acceptance ratio of 50%. In the gauge-cell MC method, each cycle consisted of 5000 trial moves (a, b, c, d, and e) with the relative probabilities of 10% translation, 10% rotation, 10% partial regrowth, 10% complete regrowth, and 60% swap. In GCMC, each cycle consisted of 2000 trial moves (a, b, c, d, and f) with the relative probabilities of 10% translation, 10% rotation, 10% partial regrowth, 10% complete regrowth, and 60% exchange with the reservoir. In GEMC, each cycle consisted of 5000 trial moves (a, b, c, d, e, and g) with the relative probabilities of 10% translation, 10% rotation, 10% partial regrowth, 10% complete regrowth, 55% exchange with the reservoir, and 5% volume change. Within the statistical uncertainty, the simulation results were found to be independent of the sequence of the trial moves. The traditional Metropolis MC algorithms are prohibitively expensive in sampling the conformations of long alkane molecules. To accelerate this process, the configurational-bias algorithm75-77 has been incorporated into the gauge-cell MC, GCMC, and GEMC simulations. The configurational-bias algorithm is used in simulations of chainlike molecules and has been widely adopted in MC simulations of phase transitions and adsorption of alkanes and polymers.78-84 It is orders of magnitude more efficient than the traditional algorithm, in which it is attempted to grow randomly the entire molecule at once. Instead of inserting a molecule at a random position, in the configurational-bias algorithm, a molecule is grown site-by-site, biasing energetically favorable configurations while avoiding overlap with other atoms, and the bias is then removed by adjusting the acceptance rules.85 In current work, eight trial positions were first generated with a probability proportional to exp(-β Uiinternal), where β ) 1/kBT and Uiinternal is the internal energy at a position i including the intramolecular bond bending and dihedral torsion interactions. One of the trial positions was then chosen for growing an atom with a probability proportional to exp(-β Uiexternal)/∑i exp(-βUiinternal), where Uiinternal is the external energy including all nonbonded intramolecular and intermolecular Lennard-Jones interactions. Also, the insertion of molecules was enhanced using the multiple first-bead scheme with fifteen trial positions.86 (75) Siepmann, J. I.; Frenkel, D. Mol. Phys. 1992, 75, 59. (76) Frenkel, D.; Mooij, G. C. A. M.; Smit, B. J. Phys.: Condensed Matter 1992, 4, 3053. (77) de Pablo, J. J.; Laso, M.; Suter, U. W. J. Chem. Phys. 1992, 96, 2395. (78) Siepmann, J. I.; Karaborni, S.; Smit, B. Nature 1993, 365, 330. (79) Smit, B.; Siepmann, J. I. Science 1994, 264, 1118. (80) Smit, B.; Maesen, T. L. M. Nature 1995, 374, 42. (81) Jiang, J. W.; Yan, Q. L.; Liu, H. L.; Hu, Y. Macromolecules 1997, 30, 8459. (82) Jiang, J. W.; Liu, H. L.; Hu, Y. Macromol. Theory Simul. 1998, 7, 105. (83) Jiang, J. W.; Sandler, S. I.; Schenk, M.; Smit, B. Phys. ReV. B 2005, 72, 045447. (84) Jiang, J. W.; Sandler, S. I. J. Chem. Phys. 2006, 124, 024717. (85) Frenkel, D.; Smit, B. Understanding Molecular Simulations: From algorithms to applications, 2nd ed.; Academic Press: San Diego, CA, 2002. (86) Esselink, K.; Loyens, L. D. J. C.; Smit, B. Phys. ReV. E 1995, 51, 1560.

Molecular Simulations of Alkanes in CNTs

Langmuir, Vol. 22, No. 17, 2006 7395

Figure 2. Saturation chemical potentials of confined and bulk C1. The solid curves are for confined C1 from gauge-cell MC in (30, 30), (25, 25), and (20, 20) SWNTs, the dashed curve is for bulk C1 predicted from the MBWR equation of state for LJ fluids.93

4. Results Figure 1 shows the isotherms of C1 in a (30, 30) SWNT at 120, 140, 160, and 180 K, respectively. The horizontal axis βµc is the dimensionless configurational chemical potential defined as βµc ) βµ - ln Λ3, and Λ is the de Broglie wavelength. The extent of adsorption is given by the overall alkane density Γ ) 2∫R0 F(r)r dr/R2 inside the SWNT of radius R, where F(r) is the local density at a radial position r. The open circles are simulation results from the gauge-cell MC. At a low temperature, say, 120 K, the isotherm exhibits a sigmoidal van der Waals loop, which is the signature of a first-order phase transition, well-known since van der Waals.87 A similar loop has been observed for the capillary condensation of LJ fluids in nanopores predicted by canonical DFT88 and for the confinement-induced phase transition of surfactant layers predicted by Scheutjens-Fleer theory.89 The coexisting vaporliquid phases are determined along the isotherm using a Maxwell equal area construction

∫µµ

Sv

V

ΓA(µ) dµ -

∫µµ

Sv

SL

ΓS(µ) dµ +

∫µµ

L

SL

ΓD(µ) dµ ) 0 (5)

where V and L are vapor and liquid equilibrated phases at a saturation chemical potential of µV ) µL, SV and SL are vapor and liquid spinodals. In this figure, the AV and DL regions are stable, the VSV and LSL regions are metastable, and the SVSL region is unstable. The coexisting phases of C1 at 120, 140, and 160 K so determined are shown in Figure 1, and the binodal curve is shown as the solid line for C1 confined in the (30, 30) SWNT. However, at 180 K, the isotherm is monotonic with increasing chemical potential, and no van der Waals loop appears. This suggests that 180 K is above the critical temperature of C1 in the (30, 30) SWNT. GCMC simulations were carried out, and the upward and downward triangles are shown in Figure 1 for adsorption and desorption, respectively. At 120, 140, and 160 K, irreversible capillary condensation and evaporation (dot-dashed lines) are clearly observed from the GCMC simulation with an IUPAC classified H1 type hysteresis.90 This type of hysteresis is the signature of phase transition in ordered cylindrical pores open (87) van der Waals, J. D. OVer de continuiteit Van den gas- en Vloeistof toestand; University of Leiden, 1873. (88) Neimark, A. V.; Ravikovitch, P. I.; Vishnyakov, A. J. Phys.-Condensed Matter 2003, 15, 347. (89) Koopal, L. K.; Leermakers, F. A. M.; Lokar, W. J.; Ducker, W. A. Langmuir 2005, 21, 10089. (90) Rouquerol, F.; Rouquerol, J.; Sing, K. Adsorption: By Powders and Porous Solids; Academic Press: London, 1999.

Figure 3. Phase diagrams of confined and bulk C1. The solid curves are the binodals of confined C1 simulated from gauge-cell MC in (30, 30), (25, 25), and (20, 20) SWNTs, the dashed and dot-dashed lines are the binodal and spinodal curves, respectively, of the bulk C1 predicted from the MBWR equation of state for LJ fluids.93 The squares and crosses are the coexisting phases of bulk C1 from GEMC68 and from experiment,94 respectively.

at both ends, as observed experimentally in mesoporous silicas MCM-41 and SBA-15 with a regular array of cylindrical pores.91 However, the experimental sample may contain pores with a wide distribution of sizes and has complex topology and connectivity. Consequently, the hysteresis behavior may be smeared out or not visible. The microscopic origin of hysteresis and its relation with pore structure is currently an issue of debate. The hysteresis associated with the mechanism of pore fillingdraining is regarded as a consequence of a metastable state, an important aspect of a first-order phase transition.92 The coexisting phases determined using the gauge-cell MC are between the capillary condensation and evaporation, i.e., inside the hysteresis loop. Though the phase transition is located within the hysteresis loop, to determine the exact position from GCMC is very difficult. With a very long GCMC run, the hysteresis loop might vanish so that the phase transition points could be determined. The salient feature of the gauge-cell MC method is that the phase transition points can be determined easily from the complete isotherm. In the stable regions AV and DL, and part of the metastable regions VSV and LSL where GCMC can be used, it gives identical results to those obtained with the gauge-cell MC. However, GCMC cannot access the unstable region SVSL. As anticipated, the hysteresis loop obtained from GCMC shrinks in size with increasing temperature from 120 to 140 K and then again at 160 K. At 180 K, the GCMC results do not exhibit hysteresis, in agreement with the gauge-cell MC results, further verifying that there is no phase transition at 180 K and that the critical temperature of C1 confined in this SWNT is below 180 K. Above the capillary critical temperature, the isotherm is reversible, indicating that the distinction between the vapor and condensed liquid phases disappears. Figure 2 shows the saturation chemical potentials of confined C1 in three SWNTs evaluated from the gauge-cell MC and of bulk C1 calculated from the MBWR equation of state for LJ fluids.93 This equation of state has been fitted to simulated data and very accurately reproduces the phase behavior and pVT data of LJ fluids. In either the SWNT or the bulk phase, the chemical potential increases (becomes less negative) as temperature rises. Compared to the bulk phase, the chemical potential in a SWNT (91) Grosman, A.; Ortega, C. Langmuir 2005, 21, 10515. (92) Gelb, L. D.; Gubbins, K. E.; Radhakrishnan, R.; Sliwinska-Bartkowiak, M. Rep. Prog. Phys. 1999, 62, 1573. (93) Johnson, J. K.; Zollweg, J. A.; Gubbins, K. E. Mol. Phys. 1993, 78, 591.

7396 Langmuir, Vol. 22, No. 17, 2006

Figure 4. Radial density profiles of C1 in a (30, 30) SWNT at 120 K and adsorbed densities of 0.048, 0.096, 0.144, 0.192, 0.240, 0.288, and 0.336 (g/mL).

Figure 5. Snapshots of C1 in a (30, 30) SWNT at 120 K and adsorbed densities of 0.048, 0.096, 0.144, 0.192, 0.240, 0.288, and 0.336 (g/mL). The arrow at the top illustrates the process from gauge-cell MC. The arrows at the bottom illustrate the processes of GCMC for adsorption and desorption, respectively, with the dot-dashed lines indicating the inaccessible regions.

is more negative, showing that the occurrence of phase separation in a confined geometry occurs at a pressure lower than the bulk saturation pressure. With decreasing SWNT diameter, the chemical potential decreases, i.e., the saturation pressure is even lower. Figure 3 shows the phase diagrams of confined and bulk C1. The gauge-cell MC was used for confined C1 in three SWNTs. For bulk C1, results from GEMC,68 the MBWR equation of state,93 and the experiment94 are plotted in the figure with good mutual agreement. Also shown is the spinodal of bulk C1 predicted from the MBWR equation of state.93 The binodal of C1 confined in the (20, 20) SWNT is inside the binodals in the (25, 25) and (30, 30) SWNTs and all are inside the binodal and spinodal of the unconfined bulk C1. Consequently, confinement shrinks the phase envelope, and the effect in a small SWNT is greater than in a large SWNT. The region within the spinodal of the unconfined C1 is unstable, and a phase transition would occur. However, if C1 is confined in a SWNT, the unstable region becomes much smaller, and part of the region unstable for the unconfined C1 is a stable state in the SWNT. This phenomenon may have relevance in the processing and storage of natural gas, in which the major component is C1, and also in interpreting data from nanoporous catalysts. With the gauge-cell MC, we were not able to simulate coexisting phases very close to the critical point due to the large density fluctuations thereby. Sophisticated algorithms have been proposed to create and destroy large clusters rather than individual (94) Smith, B. D.; Srivastava, R. Thermodynamic Data for Pure Compounds; Elsevier: Amsterdam, 1986.

Jiang and Sandler

Figure 6. Radial density profiles of C1 in a (30, 30) SWNT at 180 K and adsorbed densities of 0.048, 0.096, 0.144, 0.192, and 0.240 (g/mL).

Figure 7. Snapshots of C1 in a (30, 30) SWNT at 180 K and adsorbed densities of 0.048, 0.096, 0.144, 0.192, and 0.240 (g/mL). The arrow at the top illustrates the process from gauge-cell MC. The arrows at the bottom illustrate the processes of adsorption and desorption, respectively, from GCMC.

particles in order to sample large fluctuations.95 In this work, we simply used

Γ( ) Γc + A(Tc - T) ( 0.5B(Tc - T)β

(6)

based on the renormalization-group theory96 to estimate the critical temperature Tc and critical density Γc by fitting the coexisting phases, where the + sign is for the liquid phase and the - sign is for the vapor phase. The amplitudes A and B and the critical exponent β are the adjustable parameters. The uppermost point of each curve in Figure 3 indicates the critical point. Compared to the bulk phase, the capillary Tc in the SWNT is consistently lower, whereas Γc is higher. The shift of the critical point from that of the bulk phase becomes greater with decreasing SWNT radius. The observed shift in the critical point upon confinement is generic. When a fluid is confined by a surface, the overall fluid-fluid intermolecular attraction becomes weaker since the number of neighbors is reduced; this leads to a lower Tc. Moreover, from the figure, the shift in Tc with SWNT radius was found to obey a scaling law

∆Tc ∝ R-y

(7)

with the scaling exponent y ) 1.9, in close agreement with the experimentally determined y ) 2.1 for SF6 in interconnected (95) Auerbach, S. M. Int. ReV. Phys. Chem. 2000, 19, 155. (96) Sengers, J. V.; Sengers, J. M. H. L. In Progress in Liquid Physics; Croxton, C. A., Ed.; Wiley: Cichester, U.K., 1978; p 103.

Molecular Simulations of Alkanes in CNTs

Langmuir, Vol. 22, No. 17, 2006 7397

Figure 8. Isotherms of n-alkanes C2, C3, C4, C5, C8, and C12 in a (30, 30) SWNT. Legend as in Figure 1.

Figure 9. Binodal curves of confined and bulk linear alkanes C1, C2, C3, C4, C5, C8, and C12. The solid curves with circles are for confined nCn in a (30, 30) SWNT from gauge-cell MC, and the dotted curves with squares are for bulk nCn from GEMC.68

cylindrical controlled-pore-glass materials.97 In slitlike pores, a similar scaling relation was theoretically predicted but with y ) 1.59 for large slits and y ) 1 for small slits.12 It is generally accepted that a true phase transition does not occur at any nonzero temperatures in an infinite one-dimensional (1D) system with finite-ranged interactions.98 This scenario seems applicable to fluid confined in a SWNT as the latter is a 1D nanostructure in the axial direction. However, the SWNT being considered here is sufficiently large, and several alkane molecules can fit in the cross section of the SWNT as will be shown in Figures 5 and 7. Therefore, the confined alkane here is rigorously not an 1D system. In addition, the SWNT in our simulation is finite in length without the periodic images, and the finite-size effect can lead to sharp phase transition between two states of (97) Findenegg, G. H.; Thommes, M. High-Pressure Physisorption of Gases on Planar Surfaces and in Porous Materials. In Physical Adsorption: Experiment, Theory and Applications; Fraissard, J., Ed.; Kluwer Academic Publishers: The Netherlands, 1997; p 151. (98) Landau, L. D.; Lifshitz, E. M. Statistical Physics I; Pergamon: New York, 1980.

Figure 10. Ratios of the critical density and critical temperature of linear alkanes in a (30, 30) SWNT to those in bulk.

distinctly different densities and structures.99 However, we expect that phase transition will not occur in a rather small SWNT. Realistically, fluid molecules cannot enter into an extremely small SWNT.84 Figure 4 shows the radial density profiles of C1 in a (30, 30) SWNT at 120 K and adsorbed densities of 0.048, 0.096, 0.144, 0.192, 0.240, 0.288, and 0.336 (g/mL). At the lowest density, the preferred location of C1 molecules is in a single annular layer near the SWNT wall. With increasing density, the number of layers increases with each subsequent layer being closer to the SWNT center, and the peak in density profile in each layer also increases. At the highest density, there are up to five layers. Figure 5 shows the snapshots of C1 molecules in these systems obtained by accumulating 50 widely spaced equilibrium configurations (the SWNT is not shown). The solid line and arrow above the snapshots are to indicate that gauge-cell MC can simulate the complete isotherm including the stable, metastable, and unstable regions. The arrows and lines below the snapshots represent the adsorption and desorption processes, respectively, that can be accessed by GCMC. The dot-dashed portions of the (99) Privman, V.; Fisher, M. E. J. Stat. Phys. 1983, 33, 385.

7398 Langmuir, Vol. 22, No. 17, 2006

Jiang and Sandler

Figure 11. Binodal curves of confined and bulk nC5, iC5, and neoC5. The circles and diamonds are for confined C5 in a (30, 30) SWNT and a nanoslit from gauge-cell MC, the squares are for bulk C5 from GEMC, and the dashed curves are for bulk C5 from experiments.102-104 Table 2. Critical Temperatures and Densities of nCn in a (30, 30) Nanotube and Bulk Phase

Table 3. Critical Temperatures and Densities of C5 Isomers in a (30, 30) Nanotube, a Nanoslit, and Bulk Phase

bulk68

(30,30) SWNT

(30,30) SWNT

nCn

Tc (K)

Γc (g/mL)

Tc (K)

Γc (g/mL)

C1 C2 C3 C4 C5 C8 C12

169.8 270.0 322.7 365.5 402.7 485.9 553.2

0.200 0.264 0.289 0.311 0.322 0.340 0.356

191.4 304 368 423 470 570 667

0.160 0.206 0.221 0.231 0.238 0.239 0.235

lines indicate those regions inaccessible by GCMC. Consequently, from GCMC, one is unable to access states with three and four annular layers along either the adsorption or desorption path. Capillary condensation or evaporation is commonly regarded as a secondary process, which is always preceded by adsorption or desorption, and a hysteresis loop usually appears in the multilayer range of a physisorption isotherm. All of these feature are observed here in the sequence of simulation snapshots. Figure 6 shows the radial density profiles of C1 in a (30, 30) SWNT at 180 K and adsorbed densities of 0.048, 0.096, 0.144, 0.192, and 0.240 (g/mL). Analogous to the results at 120 K, the most preferred location of C1 molecules is in a single annular layer near the SWNT wall, and the number of layers increases upon increased density with each subsequent layer being closer to the SWNT center. However, here the layers are less distinct and the peaks are smaller because of the enhanced thermal motion. Figure 7 shows the snapshots of these same systems at 180 K showing that the distribution of C1 molecules is more dispersed than at 120 K. Because 180 K is above the critical temperature of the capillary phase transition, the gauge-cell MC and GCMC give the same results, and the complete isotherm and a series of snapshots can be obtained by both simulation methods. Figure 8 shows the isotherms of linear alkanes (C2, C3, C4, C5, C8, and C12) at various temperatures in a (30, 30) SWNT from the gauge-cell MC. For each alkane, at a low temperature, the complete isotherm shows a van der Waals loop, from which the vapor-liquid coexisting phases were estimated; above the critical temperature, the isotherm does not have a van der Waals loop. Figure 9 shows the binodal curves of confined and bulk linear alkanes (C1, C2, C3, C4, C5, C8, and C12). The circles are the confined alkanes obtained from gauge-cell MC, and the squares are the bulk alkanes from GEMC.68 As we could not simulate the coexisting phases close to the critical point, the critical parameters were estimated by fitting to the coexisting phases away from the critical point. Table 2 lists the estimated critical temperatures and densities of C1-C12 in a (30, 30) SWNT and in the bulk. Compared to the bulk, the confined fluid has a critical

nC5 iC5 neoC5

nanoslit

bulk102-104

Tc (K)

Γc (g/mL)

Tc (K)

Γc (g/mL)

Tc (K)

Γc (g/mL)

402.7 391.6 376.7

0.322 0.329 0.334

424.0 422.1 398.5

0.354 0.352 0.358

469.7 460.4 433.8

0.237 0.235 0.232

temperature that is consistently lower, a critical density that is higher, and binodal curve that is narrower. Figure 10 shows the ratios of the critical parameters (density and temperature) of linear alkanes (C1, C2, C3, C4, C5, C8, and C12) in a (30, 30) SWNT relative to those in bulk. The ratio of nanotube-to-bulk critical density increases and the ratio of critical temperature decreases as a function of carbon number. Two linear relations were used to correlate these trends

Γnanotube /Γbulk ) 1.2385 + 0.0236Nc c c

(8)

/Tbulk ) 0.8916 - 0.0053Nc Tnanotube c c

(9)

where Nc is the carbon number of the alkane molecule. From these, the critical parameters of longer linear alkanes in the (30, 30) SWNT can be predicted approximately. We expect that similar trends with different values may be found for linear alkanes in other SWNTs. The phase transitions of three C5 isomers in a carbon nanoslit of w ) 40.68 Å were simulated to investigate the difference from those in a (30, 30) SWNT which has a diameter of 40.68 Å. Figure 11 shows the binodals of nC5, iC5, and neoC5 in the SWNT, nanoslit, and bulk, with the corresponding critical temperatures and densities listed in Table 3. Whether in the SWNT, the nanoslit, or the bulk, the critical densities of the three isomers are close in value, but nC5 has the highest critical temperature and neoC5 has the lowest. Due to configurational variations, the linear nC5 is the most densely packed, whereas the branched neoC5 is the least densely packed among the three isomers. Consequently, nC5 has the strongest self-interactions and hence the highest critical temperature. This also suggests that the minimal surface area of a spherical structure results in weaker van der Waals attractive forces between fluid molecules compared to an elongated isomer structure. Such a branching effect is well demonstrated in the normal boiling temperatures of bulk C5 isomers (309.2, 301.1, and 282.6 K, respectively, for nC5, iC5, and neoC5).94 Monte Carlo simulation of lattice homopolymers100 and experimental study of star polystyrenes101 reveal a similar effect. For each isomer, the critical temperature (100) Arya, G.; Panagiotopoulos, A. Z. Macromolecules 2005, 38, 10596. (101) Alessi, M. L.; Bittner, K. C.; Greer, S. C. J. Polym. Sci. B 2004, 42, 129.

Molecular Simulations of Alkanes in CNTs

in the SWNT is lower than in the nanoslit, and both are lower than in the bulk. For a cylindrical SWNT with a diameter of 2R, the ratio of surface area to volume is 2/R; for a nanoslit with a width of 2R, the ratio is 1/R; for bulk, there is no surface, and therefore, the ratio is 0. The smaller this ratio, the weaker the pore effect, and the less the shift of the phase boundary from the bulk. As already shown in Figure 3, with an increased SWNT radius and consequently a smaller ratio of surface area to volume, the confined phase behavior gradually approaches that of the bulk fluid.

5. Conlusions We have investigated the capillary phase transitions of linear and branched alkanes in nanopores. Using the gauge-cell MC simulation, not only are the complete isotherms over stable, metastable, and unstable regions obtained, but also the coexisting phases are determined easily. In the regions where GCMC is applicable, consistent results with the gauge-cell MC are obtained. (102) Das, T. R.; Reed, C. O.; Eubank, P. T. J. Chem. Eng. Data 1977, 22, 3. (103) Das, T. R.; Reed, C. O.; Eubank, P. T. J. Chem. Eng. Data 1977, 22, 9.

Langmuir, Vol. 22, No. 17, 2006 7399

The capillary phase transition can be regarded as a surfacedriven phase transition at a lower saturation pressure than the bulk fluid. Due to geometry constraints and surface interactions, the critical point in a confined space is shifted from the bulk phase with a lowered critical temperature, an increased critical density, and a shrunken binodal curve. The effect of confinement can be qualitatively characterized by the ratio of the surface area to volume of the nanopores. The larger this ratio, the stronger the effect. The gauge-cell MC used in this work can be extended further to study phase transitions in 3D porous media, such as nanotube bundles, porous glasses, aerogels, zeolites, etc. Note that for 3D periodic structures, the application of GEMC is formidable, in which the volume change of the simulation boxes leads to the destruction of periodicity. However, GCMC can still be used in conjugation with the thermodynamic integration or histogramreweighting technique. Acknowledgment. Support from the National University of Singapore to J.J. and from the U.S. Department of Energy and U.S. National Science Foundation to S.I.S. is graciously acknowledged.

(104) Das, T. R.; Reed, C. O.; Eubank, P. T. J. Chem. Eng. Data 1977, 22, 16.

LA0608720