Publication Cover
Molecular Physics
An International Journal at the Interface Between Chemistry and Physics
Volume 121, 2023 - Issue 19-20: Thermodynamics 2022 Conference
114
Views
0
CrossRef citations to date
0
Altmetric
Thermodynamics 2022 Special Issue (by invitation only)

Quantum effects of hydrogen storage in ϵ and δ semicrystalline phases of syndiotactic polystyrene through adsorption

ORCID Icon, ORCID Icon & ORCID Icon
Article: e2244611 | Received 24 Jan 2023, Accepted 18 May 2023, Published online: 14 Aug 2023

Abstract

Syndiotactic polystyrene (sPS) is a stereoregular semicrystalline polymer which presents complex polymorphic behaviour and has been explored as an alternative for hydrogen storage using adsorption. The methodology of gas retention using a light and cheap porous material offers many advantages. To establish the adsorption capacity of a specific material, adsorption isotherms are generally measured. However, to gain a better understanding of the underlying phenomena occurring at a nanoscale level (e.g. confined interactions of H2 in sPS) and, moreover, as a predictive tool, alternative methods such as molecular simulations are required. In this work, we present Grand Canonical Monte Carlo simulation results for hydrogen (H2) adsorption isotherms of δ and ϵ phases of sPS. H2 molecule was described as a 1-site model with Lennard-Jones and Exp-6 potentials, within the Wigner–Kirkwood mean-field approximation to introduce quantum corrections. Both polymer phases were modelled using an atomistic force field and the initial configurations were relaxed using molecular dynamics for temperatures T=45K, 60K and 77K. Adsorption isotherms for each system were obtained for both classical and semiclassical potentials. Additionally, adsorption isotherms for both phases were compared with experimental adsorption data at 77K from previous work.

GRAPHICAL ABSTRACT

This article is part of the following collections:
Thermodynamics 2022 ConferenceMolecular Physics Early Career Researcher Prize

1. Introduction

Syndiotactic polystyrene (sPS) is one of the three diastereomers of polystyrene where all of the phenyl groups are positioned on alternating sides of the hydrocarbon backbone, which is not produced commercially. This order produces partially crystalline structures, giving a very rich polymorphism; until now, six crystalline structural forms have been reported. These are classified into two crystalline phase types: a trans-planar type, forming a zig-zag chain (α and β), and a helical type (γ, δ, ϵ and SI). α and β forms are obtained via thermal annealing of a melt, meanwhile (γ, δ, ϵ and SI form) are obtained from solute processing [Citation1,Citation2]. Several aspects of sPS have been investigated experimentally and using molecular simulation approach, such as structural changes when guest molecules are present, behaviour when stress is applied or a transition between phases [Citation3–5], diffusion of guest atoms or molecules in the structure [Citation6–8]. Pores (size and shape) play an important role in hydrogen storage [Citation9,Citation10] as molecule interactions are stronger in confined spaces.

Quantum effects are very noticeable under confinement since the de Broglie's thermal wavelength becomes of the order of the length size of the confinement pore and then quantum diffraction occurs, particularly at high densities, even at supercritical states not very far from the critical isotherm [Citation11,Citation12]. In a previous work [Citation13], H2 adsorption was reported for δ and ϵ sPS structures using both experimental and simulation approaches. Simulations were performed with a 3-site model for H2, with a Lennard Jones classical potential plus a Coulomb potential contribution to mimic the quadrupole moment of hydrogen. In this work we studied adsorption of H2 on δ and ϵ sPS structures, considering a 1-site H2 model. The Lennard-Jones (LJ) and Buckingham's Exponential-6 (Exp-6) pair interparticle potential models were analysed. Quantum effects were introduced at the level of the Wigner–Kirkwood (WK) mean-field approach [Citation14–17], which has been widely used to describe accurately vapour-liquid and liquid-solid phase diagrams of hydrogen and other low molecular weight substances [Citation11,Citation17,Citation18], as well as adsorption isotherms of those fluids in different substrates [Citation19,Citation20]. The details of the force field, models and equations are described in Section 2. Results obtained in this work are presented and discussed in Section 3 for T=45K, 60K and 77K, and the conclusions are given in Section 4.

2. Methodology

To determine adsorption isotherms of hydrogen on sPS in ϵ and δ phases, Grand Canonical Monte Carlo (GCMC) simulations at T=45K, 60K and 77K were performed. Bulk simulations of each value of chemical potential were carried out to obtain the relationship between pressure and chemical potential for the fluid. Each adsorption isotherm and the corresponding state point were calculated using the LJ and Exp-6 potentials, and introducing WK quantum corrections. In the following subsection, the hydrogen models are presented, explaining in Section 2.2 how the ϵ and δ structures were obtained using molecular dynamics, GCMC simulations details in Section 2.3 and adsorption isotherm calculation in Section 2.4.

2.1. Hydrogen model

Classical and semiclassical simulations were performed for a 1-site H2 model considering LJ and Exp-6 pair potentials. The corresponding semiclassical expressions were derived according to the Wigner–Kirkwood mean-field expansion [Citation14–17], (1) uWK(rij)=u(rij)+λB224π2u(rij)(1) where u(rij) is a continuous potential, uWK(rij) is the semiclassical potential energy with the Wigner–Kirkwood quantum correction and λB is the de Broglie's thermal wavelength (λB=h/2πmkBT, where m is the particle's mass, T temperature, h and kB are Planck's and Boltzmann's constants, respectively).

Explicit expressions of the WK pair potentials used are: (2) uLJ(rij)=4ϵij[(σijrij)12+11πλB22(σijrij)14(σijrij)654πλB2(σijrij)8](2) for the LJ potential, and (3) uexp6(rij)=ϵij(α6)[6eα(1rijσij)α(σijrij)6+λB224π(6α2eα(1rijσrij)σ212αeα(1rijσrij)rijσij30ασ6r8)](3) for the Exp-6 potential, where ϵij is the energy well depth between particle i and particle j, σij is the molecular diameter and α is a dimensionless parameter. The classical limit is recovered from Equations (Equation2) and (Equation3) when λB is neglected. Parameters used in this work for LJ and Exp-6 potentials are presented in Table .

Table 1. Non bonded parameters for LJ (Ref. [Citation17]) and Exp-6 (Ref. [Citation21]) potentials of hydrogen model and sPS atoms for ϵ and δ.

2.2. δ and ϵ crystalline phases

An initial box of sPS for the ϵ phase was created using the unitary cell from X-ray characterisation performed by Petraccone et al. [Citation22]. This polymorph presents an orthorhombic structure (a=1.62nm, b=2.20nm, c=0.79nm) of four chains per unitary cell, with a minimum centre-to-centre distance between channels of 1.36nm, and a density around 0.98g cm3, presenting in average two channels per unitary cell. Using the same approach reported in a previous work [Citation13], the unit cell for 6144 atoms was replicated three, two and four times in the X, Y and Z directions, respectively. The corresponding box lengths are: Lx=4.86nm, Ly=4.4nm and Lz=3.16nm. For the δ phase, the X-ray characterisation by Chatani et al. [Citation23] was used to generate the initial configuration with the fractional atomic coordinates. The structure corresponds to a monoclinic unitary cell: a=1.785nm, b=1.326nm, c=0.771nm, γ=121.2, density=1.05g cm3, two identical cavities. In this case, the unit cell was replicated three, three and four times in the X, Y and Z directions, respectively, with box lengths Lx=5.35nm, Ly=3.98nm, Lz=3.08nm and γ=121.2. Initial structures are presented in Figure  for both phases:

Figure 1. Snapshots of initial configurations taken from X-ray characterisation for (a), (c) δ and (b), (d) ϵ phase structures: (a), (b) are the projections along the a axis, (c), (d) represent the projection along the c axis.

Figure 1. Snapshots of initial configurations taken from X-ray characterisation for (a), (c) δ and (b), (d) ϵ phase structures: (a), (b) are the projections along the a axis, (c), (d) represent the projection along the c axis.

Once the ϵ and δ initial structures were generated, NVT and NPT molecular dynamic (MD) simulations were obtained with GROMACS 2021.6 [Citation24]. Three temperatures were used: 45, 60 and 77K, and a pressure P=0.1MPa. MD simulations were performed with a force field fitted to simulate benzene-polystyrene systems [Citation25]. Non-bonded and bonded parameters for polystyrene atoms can be found in Table . These parameters were taken from [Citation25]. LJ and Exp-6 non-bonded parameters for H2 molecules can be found in Table . Cross parameters between sPS atoms and H2 molecules were calculated using the Lorentz-Berthelot (LB) combining rules. Although the molecular parameters used to obtain the corresponding values for the cross-interaction belong to different potentials, an equivalent procedure has been followed in the case of the family of Mie potentials, which are Lennard-Jones interactions with different values of the repulsive and attractive exponents n and m. By changing these exponents, the potentials are modified, and the systems are described as non-conformal since can not be mapped between them by a change of energy and spatial scales. In consequence, the thermodynamic properties do not follow corresponding states. One of the first studies about the LB rules for non-conformal potentials was reported by Calvin and Reed [Citation26] and is a standard procedure in the applications of the SAFT-VR-Mie approach (see [Citation27]). A more detailed discussion is given in the supplementary information.

Table 2. Non-bonded and bonded LJ parameters for sPS atoms (δ and ϵ phases). Cal and Hal represent aliphatic carbon and hydrogen, meanwhile Car and Har represent aromatic carbon and hydrogen respectively. The force field for sPS was derived in a previous work [Citation25].

In NVT simulations, a V-rescale thermostat [Citation28] for equilibration was employed, with a simulation length time of 5ns (timestep dt=0.002ps), and a coupling parameter of τT=0.5. For NPT runs the same parameters as in NVT simulations were applied, adding a C-rescale barostat [Citation29] with a coupling parameter of τP=2.0, as this barostat can be applied for both equilibration and sampling steps. The final configuration from NPT runs was used as the adsorbent material in the GCMC simulations. Box dimensions for each final configuration of MD runs are presented in Table .

2.3. GCMC simulations

GCMC simulations were performed for H2 in bulk and confined phases. In the first case, the total chemical potential μT given by the sum of the ideal and excess contributions, μIG and μEx respectively, was fixed at different values along the isotherm, and the pressure was calculated using the virial theorem. The goals for doing these simulations were to check the performance of the LJ and Exp-6 potentials when compared with experimental data, and also to obtain the relationship between pressure and chemical potential. In the confined systems, the final configurations of the previously equilibrated sPS structures were taken to calculate the hydrogen adsorption isotherms. These structures were kept static during the Monte Carlo simulations.

Table 3. Final size of each simulation box after molecular dynamics equilibration. Dimensions Lx,Ly,Lz, solid volume VS and solid density ρS are reported.

Each MC cycle consisted of N translations, and 100N creation/removal of a particle, where N is the number of molecules in the simulation box. After 5×104 cycles for equilibration, 5×104 cycles were required for averaging with a block size of 1000 cycles.

2.4. Adsorption isotherms

To avoid ambiguities, it is important to define the considered system clearly. Most of the following definitions were taken partially from Reference [Citation30]. The crystalline structures used for MC calculations were obtained after relaxation and equilibration performed by MD calculations at the different temperatures of interest. The adsorbate is in contact with an infinite reservoir of fluid in the bulk phase at constant temperature and constant chemical potential (or pressure). The whole volume of the adsorbate VS is fixed, and it is comprised of the solid part and the empty space of the pores. The total number of moles in the system is: (4) nT=nads+nS(4) where nads are the moles of adsorbate and nS the moles of the solid. The absolute adsorption (obtained directly from GCMC simulations) is given when the moles of the solid are removed: (5) nabs=nTnS=nads(5) To calculate net adsorption, the moles contained in a fluid at the same pressure and temperature of the system are subtracted from the absolute adsorption: (6) nnet(μ,T)=nabs(μ,T)nbulk(μ,T)(6) If we divide Equation (Equation6) by VS, the following expression is obtained (7) nnetVS=Γnet(μ,T)=Γabs(μ,T)Γbulk(μ,T)(7) where Γbulk is the molar density, obtained by running a GCMC simulation in a box of volume V at T and μ employing the following relation: (8) Γbulk(μ,T)=n(μ,T)V(8) The corresponding pressures were obtained from the same simulations using the virial equation for pressure [Citation31].

The adsorption isotherms presented in the Results section were represented in mass/mass percentage. These are the units generally used by many experimentalists in hydrogen storage. To calculate it, the next relations were employed: (9) Total amount adsorbed (wt \%)=ΓabsMH2ρS100(9) (10) Net amount adsorbed (wt\%)=MH2(ΓabsΓbulk)ρS100=MH2ΓnetρS100(10) where MH2 is the molecular weight of hydrogen in gmol1 and ρS=mSVS is the solid density in gm3. Solid density values for each system can be found in Table 3.

3. Results and discussion

Performance of the potential models was studied first in bulk phase, i.e. no sPS adsorbent was present. Low temperatures for H2 were considered, with pressures ranging from P=0.05MPa to 10 MPa, modifying the chemical potential accordingly in each simulation point. As observed in Figure , the classical and semiclassical potentials agree very well with experimental data for temperature T=45K and low pressures below P=1.5MPa. This data was obtained using the database REFPROP [Citation32]. However, at higher pressures, both classical potentials tend to overestimate the experimental density, whereas the semiclassical WK models give very accurate predictions, even for very high pressures, P=10MPa. This behaviour is expected since quantum effects should be evident at low temperatures and high pressures, whereas classical potentials do not include quantum effects.

Figure 2. Density (ρ) vs. pressure (P) plot for H2 in bulk phase at T=45K in linear (left) and semi-log (right) scale. Symbols represent data obtained by GCMC simulations for each intermolecular potential (circles for LJ, triangles down for LJ-3 sites model, squares for LJ-WK, triangles for Exp-6 and diamonds for Exp-6-WK). Dashed lines are only a visual guide. Solid line represents experimental data retrieved from REFPROP [Citation32].

Figure 2. Density (ρ) vs. pressure (P) plot for H2 in bulk phase at T=45K in linear (left) and semi-log (right) scale. Symbols represent data obtained by GCMC simulations for each intermolecular potential (circles for LJ, triangles down for LJ-3 sites model, squares for LJ-WK, triangles for Exp-6 and diamonds for Exp-6-WK). Dashed lines are only a visual guide. Solid line represents experimental data retrieved from REFPROP [Citation32].

For T=60K (Figure ), the calculated densities using GCMC simulations (classical and semiclassical potentials) agree with the experimental data reported up to a pressure P=2.0MPa. At higher pressures, the densities calculated by adding quantum corrections agree with the experimental densities; meanwhile, the curve obtained from classic potentials deviates significantly as pressure increases. However, this deviation from experimental densities is smaller than the one observed in the previous temperature (T=45K), which agrees with the fact that quantum effects become more relevant at lower temperatures. As the temperature increases, the de Broglie thermal wavelength λB decreases, implying a lower contribution of quantum effects.

Figure 3. ρ vs. P plot for H2 in bulk phase at T=60K in linear (left) and semi-log (right) scale. Symbols represent data obtained by GCMC simulations for each intermolecular potential (circles for LJ, triangles down for LJ-3 sites model, squares for LJ-WK, triangles for Exp-6 and diamonds for Exp-6-WK). Dashed lines are only a visual guide. Solid line represents experimental data retrieved from REFPROP [Citation32].

Figure 3. ρ vs. P plot for H2 in bulk phase at T=60K in linear (left) and semi-log (right) scale. Symbols represent data obtained by GCMC simulations for each intermolecular potential (circles for LJ, triangles down for LJ-3 sites model, squares for LJ-WK, triangles for Exp-6 and diamonds for Exp-6-WK). Dashed lines are only a visual guide. Solid line represents experimental data retrieved from REFPROP [Citation32].

The density obtained for temperature T=77K for the classical and WK models is presented in Figure , In both cases there is agreement with experimental data but deviations are noticeable at high pressures, P=5.0MPa, where the classical potentials also overestimate density with respect to semiclassical potentials. It is important to remark that a 1-site hydrogen model (no charges present) agrees with experimental data and deviations at low temperatures are greatly reduced when adding a quantum correction term to the classical potentials.

Figure 4. ρ vs. P plot for H2 in bulk phase at T=77K in linear (left) and semi-log (right) scale. Symbols represent data obtained by GCMC simulations for each intermolecular potential (circles for LJ, triangles down for LJ-3 sites model, squares for LJ-WK, triangles for Exp-6 and diamonds for Exp-6-WK). Dashed lines are only a visual guide. Solid line represents experimental data retrieved from REFPROP [Citation32].

Figure 4. ρ vs. P plot for H2 in bulk phase at T=77K in linear (left) and semi-log (right) scale. Symbols represent data obtained by GCMC simulations for each intermolecular potential (circles for LJ, triangles down for LJ-3 sites model, squares for LJ-WK, triangles for Exp-6 and diamonds for Exp-6-WK). Dashed lines are only a visual guide. Solid line represents experimental data retrieved from REFPROP [Citation32].

As mentioned in Section 2, bulk simulations allow us to relate density, chemical potential and pressure for every temperature that is required to calculate adsorption isotherms for H2 confined in δ and ϵ phases. Moreover, quantum effects in H2 bulk system can be analysed, which seems to be relevant even at T=77K.

For sPS systems, both structures at different temperatures were analysed, at the same chemical potentials and corresponding pressures used in bulk simulations. Figure  compares adsorption isotherms for each potential at T=45K for δ and ϵ structures. Figure a and c present the total amount adsorbed in sPS structures, calculated by Equation (Equation9). Figure b and d give the net amount adsorbed, calculated by Equation (Equation10). As it is observed in Figure a, the adsorption calculated using the Exp-6 potential is noticeably higher than the one calculated using the other potentials. It is expected that classic potentials tend to have higher adsorption than semiclassical ones, as it is known that the inclusion of quantum corrections introduces a higher value of the effective diameter, that at low densities goes as σeffσ+λB/22 [Citation33]. This effective diameter reduces the available adsorption volume since the excluded volume increases with respect to the classical model, an effect that we have reported previously for a semiclassical thermodynamic perturbation theory model for square-well particles [Citation18]. In Figure a it is possible to note that, for δ phase, the sPS saturates approximately at P=0.2MPa for LJ and Exp-6-WK potentials. For the Exp-6 and LJ-WK models, adsorption isotherms saturate approximately at P=1.5MPa. In Figure b, we observe a maximum of net adsorption at very low pressure, P<0.1MPa for all potentials.

Figure 5. Adsorption isotherms of H2 in sPS at T=45K for LJ, LJ-WK, Exp-6 and Exp-6-WK potentials calculated by GCMC simulations. Figure (a) and (b) present total adsorption and net adsorption isotherms of H2 (in percentage) respectively, for δ phase. Figure (c) and (d) present total adsorption and net adsorption isotherms of H2 (in percentage) respectively, for the ϵ phase.

Figure 5. Adsorption isotherms of H2 in sPS at T=45K for LJ, LJ-WK, Exp-6 and Exp-6-WK potentials calculated by GCMC simulations. Figure (a) and (b) present total adsorption and net adsorption isotherms of H2 (in percentage) respectively, for δ phase. Figure (c) and (d) present total adsorption and net adsorption isotherms of H2 (in percentage) respectively, for the ϵ phase.

For the ϵ phase, values for adsorption have a significant variation for each potential, as observed in Figure c. LJ and Exp-6 systems predict higher adsorption with respect to the semiclassical fluids. The adsorption isotherms saturate at lower pressures when the quantum correction is applied (0.5MPa). As in the δ phase, net adsorption is at a maximum in pressures not greater than 0.1MPa for all potentials (see Figure d).

Adsorption at temperature T=60K continues to be higher when the classical models are used. The LJ-WK potential predicts the lowest adsorption in the sPS structure. This is also an expected behaviour, thus the prediction for H2 packing in δ cavities is lower than for the Exp6-WK model.

Adsorption of H2 in these conditions saturates at pressure P=2.0MPa for both WK potential models, P<0.1MPa and P=1MPa for the LJ and Exp-6 models, respectively, see Figure a. In the case of the δ phase, the maximum net adsorption is presented at very low pressures, basically P<0.1MPa for all the potential models, similarly to what occurs at temperature T=45K, as observed in Figure b. For the ϵ structure a lower adsorption for all the potential models is obtained when compared with the δ phase, see Figure c. In this case, the saturation range in pressure is 1.0MPa<P<1.5MPa, with the maximum net adsorption values at pressures below 0.1MPa for the LJ model and both WK systems, whereas P=0.3MPa for the classical Exp-6 (see Figure d).

Figure 6. Adsorption isotherms of H2 on sPS at T=60K for LJ, LJ-WK, Exp-6 and Exp-6-WK potentials calculated by GCMC simulations. Figure (a) and (b) present total adsorption and net adsorption isotherms of H2 (in percentage) respectively, for δ phase. Figure (c) and (d) present total adsorption and net adsorption isotherms of H2 (in percentage) respectively, for the ϵ phase.

Figure 6. Adsorption isotherms of H2 on sPS at T=60K for LJ, LJ-WK, Exp-6 and Exp-6-WK potentials calculated by GCMC simulations. Figure (a) and (b) present total adsorption and net adsorption isotherms of H2 (in percentage) respectively, for δ phase. Figure (c) and (d) present total adsorption and net adsorption isotherms of H2 (in percentage) respectively, for the ϵ phase.

A relevant feature comparing the adsorption isotherms for the sPS structures are very similar for both WK potential models, whereas for the δ structure they are very well differentiated. A plausible physical explanation for this is that the cavities and channels present in the δ and ϵ phases, respectively, give different response for different potential models. H2 molecules are closer each other in cavities than in channels, as can be seen in Figure  for the LJ and LJ-WK models. The effective higher molecule predicted by the quantum correction modifies the H2 adsorption, as also predicted in a previous theoretical work for adsorption of H2 onto carbon-based structures [Citation18]. Comparing Figure a and b we observe that LJ-WK molecules are more packed in the cavities of the δ phase when compared to the corresponding classical model. The same occurs for the channels in the ϵ phase, see Figure c and d.

Figure 7. Snapshots of final GCMC configurations at T=60K for (a) δ phase LJ potential, (b), δ phase LJ-WK potential, (c) ϵ phase LJ potential and (d) ϵ phase LJ-WK potential, using projections along the a axis in each snapshot.

Figure 7. Snapshots of final GCMC configurations at T=60K for (a) δ phase LJ potential, (b), δ phase LJ-WK potential, (c) ϵ phase LJ potential and (d) ϵ phase LJ-WK potential, using projections along the a axis in each snapshot.

Adsorption isotherms for temperature T=77K are presented in Figure . A rough comparison in both crystalline phases with experimental adsorption isotherms [Citation13] is given. Due to the nature of the material, with amorphous and crystalline conformations, it is not possible to do a direct comparison with experimental data, since the experimental adsorption isotherms were obtained by subtracting from the total amount adsorbed in both phases the adsorption on the dense β sPS phase at temperature T=77K. All the sPS semi-crystalline materials were prepared as aerogels of the same nano fibrillar morphology, with a 35-40 % degree of crystallinity and with the same range of porosity, 88-95 %.

Figure 8. Adsorption isotherms of H2 in sPS at T=77K for LJ, LJ-WK, Exp-6 and Exp-6-WK potentials calculated by GCMC simulations. Figure (a) and (b) present total adsorption and net adsorption isotherms of H2 (in percentage) respectively, for δ phase. Figure (c) and (d) are total adsorption and net adsorption isotherms of H2 (in percentage) respectively, for the ϵ phase. In Figure (a) and (c), also experimental (solid line) and simulation (cross symbols) values from a previous work [Citation13] are plotted.

Figure 8. Adsorption isotherms of H2 in sPS at T=77K for LJ, LJ-WK, Exp-6 and Exp-6-WK potentials calculated by GCMC simulations. Figure (a) and (b) present total adsorption and net adsorption isotherms of H2 (in percentage) respectively, for δ phase. Figure (c) and (d) are total adsorption and net adsorption isotherms of H2 (in percentage) respectively, for the ϵ phase. In Figure (a) and (c), also experimental (solid line) and simulation (cross symbols) values from a previous work [Citation13] are plotted.

The adsorption on the β phase is due to the amorphous part because the crystalline part has a very high packing and is without space to adsorb hydrogen. A more detailed explanation of these experiments can be found elsewhere in Ref. [Citation13], Figures 2 and 9. In this previous work, simulations were also developed but a different model was used; a 3-site hydrogen model with explicit Coulomb's charges. Experimental data was measured for both structures up to a pressure P=1.75MPa. In Figure a, the potentials used in this work present higher adsorption than the 3-site hydrogen model. The LJ-WK potential curve seems to follow better the experimental data for δ phase at low pressures (P<2.0MPa). Classic potentials have a very similar adsorption behaviour. Figure b indicates maximum net adsorption at P=0.3MPa for Exp-6-WK potential and P<0.1MPa for other potentials.

In Figure c we can observe that the WK models give a better prediction of the low pressures adsorption isotherms, up to P=1.0MPa and particularly for the Exp-6 potential. Using semiclassical potentials for the ϵ phase does predict the saturation at P=0.5MPa and classical potentials do allow more adsorption, saturating approximately at P=1.5MPa. In Figure d, the maximum net adsorption is observed at P=0.5MPa for Exp-6 potential and P<0.1MPa for the other potentials. It is relevant to note that, even at T=77K where bulk simulations do not present a significant difference between potentials (up to P=5.0MPa), adsorption isotherms in sPS structures do change for all potentials. Also, in the δ phase, differences between LJ-WK and Exp-6-WK are noticeable at each temperature; however, for the ϵ phase, adsorption isotherms tend to have similar values as temperature increases.

4. Conclusions

In this work we have presented the relevance of introducing quantum effects in order to describe adsorption isotherms of H2 in δ and ϵ phases of sPS. Using 1-site model classical simulations for H2 in bulk phases can reproduce experimental data only at low pressures, P<2.0MPa, whereas the Wigner–Kirkwood mean-field semiclassical potentials can be used even for pressures as high as 10MPa. For H2 in syndiotactic polystyrene (sPS) structure, adsorption is higher on δ phase than ϵ phase for temperatures T=45, 60 and 77K. For these isotherms, the difference between LJ and Exp-6 potential are more notable in δ phase than ϵ phase. As already mentioned before, these differences could be consequence of the different pore morphology in the δ and ϵ phases, since stronger confinement is present in the ϵ structure, and then the differences between potentials must be more noticeable, particularly when quantum corrections are introduced. More experimental data of H2 adsorption isotherms at lower temperatures (T=45K and 60K) in a wider range of pressures for these sPS structures are required to determine which potential has a better prediction. Experimental data at T=77K suggests that, for both δ and ϵ phases, a 1-site H2 model using a LJ-WK potential could perform a reasonable prediction of H2 adsorption at low pressures (P<1MPa), followed by Exp-6-WK potential. Finally, for each temperature studied in this work, adsorption isotherms for classical potentials tend to be higher than their semiclassical counterparts.

Supplemental material

Supplemental Material

Download PDF (404.7 KB)

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

We acknowledge support from the University of Guanajuato, Dirección de Investigación y Posgrado (grant CIIC 151/2022 and 250/2023). Also, we gratefully acknowledge the computing time granted by LANCAD/CONACyT on the supercomputer Miztli at DGTIC UNAM. We thank CONACyT for Kevin R. Arriola Gonzalez's PhD scholarship, number 819666.

References

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.