Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
License: CC BY 4.0
arXiv:2403.03162v2 [cond-mat.soft] 21 Mar 2024

Statistical modeling of equilibrium phase transition in confined fluids

Gunjan Auti gunjanauti@thml.t.u-tokyo.ac.jp Department of Mechanical Engineering, The University of Tokyo,
7-3-1 Hongo, Bunkyo-ku, Tokyo 113-8656, Japan
   Soumyadeep Paul Department of Mechanical Engineering, The University of Tokyo,
7-3-1 Hongo, Bunkyo-ku, Tokyo 113-8656, Japan
Department of Mechanical Engineering, Stanford University, Building 530, 440 Escondido Mall, Stanford, California 94305, USA
   Wei-Lun Hsu Department of Mechanical Engineering, The University of Tokyo,
7-3-1 Hongo, Bunkyo-ku, Tokyo 113-8656, Japan
   Shohei Chiashi Department of Mechanical Engineering, The University of Tokyo,
7-3-1 Hongo, Bunkyo-ku, Tokyo 113-8656, Japan
   Shigeo Maruyama Department of Mechanical Engineering, The University of Tokyo,
7-3-1 Hongo, Bunkyo-ku, Tokyo 113-8656, Japan
   Hirofumi Daiguji daiguji@thml.t.u-tokyo.ac.jp Department of Mechanical Engineering, The University of Tokyo,
7-3-1 Hongo, Bunkyo-ku, Tokyo 113-8656, Japan
(March 21, 2024)
Abstract

The phase transition of confined fluids in mesoporous materials deviates from that of bulk fluids due to the interactions with the surrounding heterogeneous structure. For example, adsorbed fluids in metal-organic-frameworks (MOFs) have atypical phase characteristics such as capillary condensation and higher-order phase transitions due to a strong heterogeneous field. Considering a many-body problem in the presence of a nonuniform external field, we model the host-guest and guest-guest interactions in MOFs. To solve the three-dimensional Ising model, we use the mean-field theory to approximate the guest-guest interactions and Mayer’s f𝑓fitalic_f-functions to describe the host-guest interactions in a unit cell. Later, using Hill’s theory of nanothermodynamics, we define differential thermodynamic functions to understand the distribution of intensive properties and integral thermodynamic functions to explain the phase transition in confined fluids. The investigation reveals a distinct behavior where fluids confined in larger pores undergo a discontinuous (first-order) phase transition, whereas those confined in smaller pores experience a continuous (higher-order) phase transition. Furthermore, the results indicate that the free-energy barrier for phase transitions is lower in confined fluids than in bulk fluids giving rise to a lower condensation pressure relative to the bulk saturation pressure. Finally, the integral thermodynamic functions are succinctly presented in the form of a phase diagram, marking an initial step toward a more practical approach for understanding the phase behavior of confined fluids.

preprint: APS/123-QED

I Introduction

Confined fluids can be gas (vapor) trapped in nanobubbles [1], adsorbed gas in porous structures [2, 3, 4], natural gas trapped in shale and tight rock formations [5], or biomolecules trapped in cells [6]. Fluids such as these enclosed within restricted spaces have distinct physical and thermodynamic characteristics that differ from those of bulk fluids. These unique characteristics include phenomena such as an atypical phase transition and packing polymorphism [7, 8, 9, 10], a shift in the freezing and melting points [11], an anomalously low dielectric constant [12], and very high hydrodynamic slippage [13]. Although these phenomena have been widely observed and reported, how heterogeneity and their multiscale nature affect the fluid characteristics still lacks a comprehensive thermodynamic understanding.

The atypical thermodynamic properties are due to the heterogeneous interactions and steric hindrance of confined fluids. Generally, these interactions take the form of van der Waals forces and the inverse radial dependence of these cohesive interactions translates into a layered distribution of density near the surface, which creates anisotropy [14, 15, 16, 17, 18]. Gibbs laid the foundation for modeling how heterogeneous interactions affect fluid properties by formulating surface thermodynamics, where he introduced the concept of the Gibbs surface excess [19]. Later, Hill put forth a thermodynamic approach to model small systems [20]. Since then, numerous models describing how heterogeneity affects the fluid properties have been proposed. For example, various self-consistent field models have been presented based on the analogous Bogoliubov–Born–Green–Kirkwood–Yvon hierarchy describing a system containing a large number of interacting particles [21, 22, 23, 24, 25, 26, 27, 28, 29]. These models can be solved for special cases but become much more complex for a different sets of conditions.

To make the model more realizable under general conditions, Sircar and Mayers [2, 30] proposed a semi-empirical formulation for low-concentration adsorption. They coined the term “ideal adsorbed phase” to indicate a behavior that differs from that of bulk fluids. Nicholson [31, 32] later presented an elaborate molecular theory focusing on the adsorption of lattice gas. This model shows qualitatively how the thermodynamic properties vary but does not focus on the phase transition of the adsorbed fluid. Martinez et al. [33] predicted adsorption isotherms using a two-dimensional statistical associating fluid theory for a square well potential on a flat surface.

In parallel, with the advancement of molecular simulations, Evans [34, 3] undertook extensive Monte Carlo simulations to elucidate phase transitions in mesoporous slits. Schmidt and Löwen [35] introduced a hard-sphere model to investigate the freezing transition between parallel plates, presenting a phase diagram through Monte Carlo simulations. Subsequently, numerous studies used brute force molecular simulations to produce both qualitative and quantitative phase diagrams. For instance, Kimura and Maruyama [36] explored boiling phase transitions and cluster formation using molecular dynamics simulations, while Radhakrishnan et al. [37] used biased potentials and umbrella sampling in grand canonical Monte Carlo (GCMC) simulations to obtain thermodynamic properties. Takaiwa et al. [38] presented a phase diagram of water in carbon nanotubes. Zhou et al. [39] conducted ab initio simulations using density functional theory and grand canonical algorithms to illustrate a surface phase diagram.

More recently, these molecular simulation techniques have been improved by incorporating various machine-learning algorithms [40, 41, 42, 43]. Such methodologies aim to extract specific properties of adsorbed fluids, reducing computational demands and facilitating the prediction of certain thermodynamic properties. However, even the most current deep-learning models are black boxes [44, 45], meaning that they perform the majority of their calculations internally, so that significant thermodynamic information is overlooked. This hinders a comprehensive understanding of the physics underlying adsorption and confinement. Moreover, deep-learning algorithms are essentially the methods of statistical mechanics, as explained by Lin et al. [46]. Consequently, understanding the thermodynamics of confined fluids requires statistical methods and analytical models based on appropriate approximations.

Statistical models have been used to understand various multiscale processes. For example, Košmrlj and Nelson [47] gave a model for the thin shells and argued that large spherical shells are unstable due to thermally generated pressure using statistical mechanics. Goodrich et al. [48] formulated a statistical model for nanocluster formation in the crystallization process, and Molina et al. [49] experimentally described the many-body interactions that occur in confined space for self-organizing of droplets. In a similar way, to understand the multiscale process of phase transition in confined fluids, we have developed a semi-analytical statistical model.

Refer to caption
Figure 1: Statistical model for adsorbed fluid. (a) Schematic representation of the system under consideration. The argon molecules (red) in bulk with chemical potential μbulksuperscript𝜇bulk\mu^{\text{bulk}}italic_μ start_POSTSUPERSCRIPT bulk end_POSTSUPERSCRIPT are in equilibrium with the system of adsorbed molecules with chemical potential μadssuperscript𝜇ads\mu^{\text{ads}}italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT in a metal-organic framework (MOF) with chemical potential μframesuperscript𝜇frame\mu^{\text{frame}}italic_μ start_POSTSUPERSCRIPT frame end_POSTSUPERSCRIPT. The relative density distribution obtained from the proposed model for an MOF with a 12 Å unit cell is shown at (b) low relative pressure (p/p0=0.04𝑝subscript𝑝00.04p/p_{0}=0.04italic_p / italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.04) and (c) after saturation (p/p0=0.20𝑝subscript𝑝00.20p/p_{0}=0.20italic_p / italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.20). Similar distribution functions are plotted for a 24 Å unit cell MOF at (d) low relative pressure (p/p0=0.10𝑝subscript𝑝00.10p/p_{0}=0.10italic_p / italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.10) and (e) after saturation (p/p0=0.70𝑝subscript𝑝00.70p/p_{0}=0.70italic_p / italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.70). The deep blue shade represents the heterogeneity (metal or ligand) and all distributions are plotted for a cross section of the unit cell at z=a/2𝑧𝑎2z=a/2italic_z = italic_a / 2.

The present paper introduces a three-dimensional (3D) Ising model for argon confined in a cubic metal-organic framework [MOF, see Fig. 1(a)]. This approach considers the nonuniformity of the external field by decoupling homogeneous and heterogeneous interactions. The homogeneous interactions are considered through mean-field theory and the heterogeneous interactions are approximated using Mayer’s f𝑓fitalic_f-functions. The nonuniform interactions lead to a nonuniform density distribution in the pores, as depicted in Figs. 1(b)–1(e). To account for the thermodynamic properties of such a distribution, differential (local) and integral (global) intensive thermodynamic functions can be defined. For example, Figs. 1(b) and 1(c) show the expected relative density distribution (a differential property) of argon at relative pressures, p/p0=0.04𝑝subscript𝑝00.04p/p_{0}=0.04italic_p / italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.04 and p/p0=0.20𝑝subscript𝑝00.20p/p_{0}=0.20italic_p / italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.20, respectively, for a small pore (a=12𝑎12a=12italic_a = 12 Å). Even for extremely low relative pressure, the pores are close to saturation. Conversely, Figs. 1(d) and 1(e) show the expected relative density distribution of argon at relative pressures p/p0=0.10𝑝subscript𝑝00.10p/p_{0}=0.10italic_p / italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.10 and p/p0=0.70𝑝subscript𝑝00.70p/p_{0}=0.70italic_p / italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.70, respectively, for a large pore (a=24𝑎24a=24italic_a = 24 Å). Here, the layered adsorption occurs near the heterogeneity at lower relative pressure and a uniform density distribution develops at higher relative pressure. Later in this paper, we derive the integral thermodynamic functions using Hill’s thermodynamics for small systems. Based on these integral properties, we discuss the phase transition of confined fluids and compare it with that of bulk fluids.

This paper discusses the statistical modeling of the confined fluid while highlighting the phase transition during capillary condensation. We also showcase the phase diagram of the adsorbed fluid and discuss the similarities and key differences vis-á-vis the bulk fluid. The remainder of the paper is organized as follows: Section II.2 details a mathematical derivation of the Ising model in a grand canonical ensemble for a confined fluid. Section II.3 derives the relevant thermodynamic properties of the adsorbed fluids. Section III.1 explains the benchmarking of the proposed model with GCMC simulations. Section III.2 discusses the phase transition as a function of pore size based on a double-well potential. Sec. III.3 introduces the phase diagram for the adsorbed fluid. Section IV.1 then discusses how this model may be used in different contexts (e.g., where the confinement effect is significant). Finally, Sec. IV.2 presents the conclusions of this paper.

II Model

II.1 General assumptions

We focus on the classical regime, where quantum effects may be disregarded. Consequently, the van der Waals potentials generated by different sources are treated as additive, following the established principles of intermolecular forces [50]. Furthermore, the analysis assumes equilibrium conditions throughout. This assumption is based on the premise that external conditions, such as temperature and pressure, are time-independent. Given such conditions, macroscopic quantities can be expressed in terms of microscopic average values, distribution functions, or probabilities.

Furthermore, mean-field theory is used to solve the many-body problem of adsorbents in a potential well of nonuniform depth. This assumption is valid under conditions of low adsorbent concentration because interactions between adsorbed molecules are negligible at such concentrations. Moreover, the mean-field theory remains valid at and above the saturation point for capillary condensation because the distribution of fluid molecules within pores becomes uniform, leading to a mean-field effect. However, the mean-field theory may not be accurate in a high-density gas-like regime. In such cases where the adsorbents are relatively concentrated, a density distribution around the heterogeneity creates an anisotropy. However, this effect is not considered in the current model.

Given these assumptions, the current work provides a framework for analyzing and understanding the behavior of fluids in confined spaces, particularly in the context of adsorption in MOFs. These assumptions allow for simplified models and calculations, enabling insights into the thermodynamic properties and phase transitions of the adsorbed fluids.

To clarify the formulation of the model, we briefly revisit the fundamental concepts of Hill’s nanothermodynamics [51] in the context of the current problem (a detailed derivation and discussion are available in Ref. [52]). In our case, fluid confined in the pore of a modeled MOF is in equilibrium with the surrounding bulk fluid. The argon molecules confined within the MOF are the “system” in this investigation. To understand the thermodynamic characteristics of this system, we subdivide it into an ensemble of η𝜂\etaitalic_η small, equivalent, distinguishable, independent systems, as shown in Fig. 1(a). Therefore, assuming the total volume is constant, the system described here at equilibrium gives

dEt=TdSt+μdNt+ξdη,𝑑subscript𝐸𝑡𝑇𝑑subscript𝑆𝑡𝜇𝑑subscript𝑁𝑡𝜉𝑑𝜂\displaystyle dE_{t}=TdS_{t}+\mu dN_{t}+\xi d\eta,italic_d italic_E start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = italic_T italic_d italic_S start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_μ italic_d italic_N start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_ξ italic_d italic_η , (1)

where Etsubscript𝐸𝑡E_{t}italic_E start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT is the total energy of the system, T𝑇Titalic_T is the temperature, Stsubscript𝑆𝑡S_{t}italic_S start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT is the total entropy of the system, μ𝜇\muitalic_μ is the chemical potential, Ntsubscript𝑁𝑡N_{t}italic_N start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT are the total number of molecules in the system, ξ𝜉\xiitalic_ξ is the subdivision potential, and η𝜂\etaitalic_η is the number of subdivisions.

Equation (1) resembles Gibb’s equilibrium equation for a two-component system with Ntsubscript𝑁𝑡N_{t}italic_N start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT being the number of molecules. Since η𝜂\etaitalic_η is the number of subdivisions then the total volume Vt=ηVsubscript𝑉𝑡𝜂𝑉V_{t}=\eta Vitalic_V start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = italic_η italic_V, where V𝑉Vitalic_V is the volume of each subdivision. Therefore, we consider the work associated with varying η𝜂\etaitalic_η at pressure p𝑝pitalic_p by adding the work of expansion, pηdV𝑝𝜂𝑑𝑉-p\eta dV- italic_p italic_η italic_d italic_V, in Eq.  (1) to obtain

dEt=TdStpηdV+μdNt+ξdη,𝑑subscript𝐸𝑡𝑇𝑑subscript𝑆𝑡𝑝𝜂𝑑𝑉𝜇𝑑subscript𝑁𝑡𝜉𝑑𝜂\displaystyle dE_{t}=TdS_{t}-p\eta dV+\mu dN_{t}+\xi d\eta,italic_d italic_E start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = italic_T italic_d italic_S start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT - italic_p italic_η italic_d italic_V + italic_μ italic_d italic_N start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_ξ italic_d italic_η , (2)

where pηEt/V𝑝𝜂subscript𝐸𝑡𝑉-p\eta\equiv\partial E_{t}/\partial V- italic_p italic_η ≡ ∂ italic_E start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT / ∂ italic_V. We now use Hill’s definition of subdivision potential, ξp^V𝜉^𝑝𝑉\xi\equiv-\hat{p}Vitalic_ξ ≡ - over^ start_ARG italic_p end_ARG italic_V, where p^^𝑝\hat{p}over^ start_ARG italic_p end_ARG is the integral pressure [52]. Integrating Eq. (2) gives the equilibrium equation for the total system:

Et=TSt+μNtp^Vη.subscript𝐸𝑡𝑇subscript𝑆𝑡𝜇subscript𝑁𝑡^𝑝𝑉𝜂\displaystyle E_{t}=TS_{t}+\mu N_{t}-\hat{p}V\eta.italic_E start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = italic_T italic_S start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT + italic_μ italic_N start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT - over^ start_ARG italic_p end_ARG italic_V italic_η . (3)

From here, it is straightforward to show for a grand canonical ensemble that

p^V=kBTlnΞ,^𝑝𝑉subscript𝑘B𝑇Ξ\displaystyle\hat{p}V=k_{\text{B}}T\ln\Xi,over^ start_ARG italic_p end_ARG italic_V = italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T roman_ln roman_Ξ , (4)

where kBsubscript𝑘Bk_{\text{B}}italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT is the Boltzmann constant and ΞΞ\Xiroman_Ξ is the grand partition function for the small chosen system of volume V𝑉Vitalic_V.

Hereinafter, we use Hill’s notation [52] where the hat (^^absent\hat{~{}}over^ start_ARG end_ARG) denotes the integral intensive thermodynamic function and for any extensive thermodynamic function α𝛼\alphaitalic_α, and

α¯1VVα𝑑V¯𝛼1𝑉subscript𝑉𝛼differential-d𝑉\displaystyle\bar{\alpha}\equiv\frac{1}{V}\int_{V}\alpha dVover¯ start_ARG italic_α end_ARG ≡ divide start_ARG 1 end_ARG start_ARG italic_V end_ARG ∫ start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT italic_α italic_d italic_V (5)

denotes the integral extensive thermodynamic function. In contrast, symbols without hat or bar represent differential thermodynamic functions.

II.2 Ising framework

To understand the phase transition, the Ising model plays a crucial role. The inherent complexity in the 3D Ising model coupled with the external non-uniform field presents significant challenges. However, we have made certain assumptions, mentioned in the prior section to obtain an approximate solution. The formulation for the confined fluids is as follows:

Let 𝐩=(p1,p2,,pN)𝐩subscript𝑝1subscript𝑝2subscript𝑝𝑁\mathbf{p}=(p_{1},p_{2},\dots,p_{N})bold_p = ( italic_p start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_p start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_p start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ), 𝐪=(q1,q2,,qN)𝐪subscript𝑞1subscript𝑞2subscript𝑞𝑁\mathbf{q}=(q_{1},q_{2},\dots,q_{N})bold_q = ( italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_q start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_q start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ) be the momentum and position coordinates, respectively, in phase space for a system of N𝑁Nitalic_N molecules confined in a framework creating a potential Uma(𝐪)subscript𝑈𝑚𝑎𝐪U_{ma}(\mathbf{q})italic_U start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT ( bold_q ). The Hamiltonian \mathcal{H}caligraphic_H is then

(𝐩,𝐪)𝐩𝐪\displaystyle\mathcal{H}(\mathbf{p},\mathbf{q})caligraphic_H ( bold_p , bold_q ) =𝒦(𝐩,𝐪)+𝒰(𝐩,𝐪)absent𝒦𝐩𝐪𝒰𝐩𝐪\displaystyle=\mathcal{K}(\mathbf{p},\mathbf{q})+\mathcal{U}(\mathbf{p},% \mathbf{q})= caligraphic_K ( bold_p , bold_q ) + caligraphic_U ( bold_p , bold_q ) (6a)
=iNpi22m+iN1j>iNΦ(𝐫i𝐫j)+iNUma(qi),absentsuperscriptsubscript𝑖𝑁subscriptsuperscript𝑝2𝑖2𝑚superscriptsubscript𝑖𝑁1superscriptsubscript𝑗𝑖𝑁Φsubscript𝐫𝑖subscript𝐫𝑗superscriptsubscript𝑖𝑁subscript𝑈𝑚𝑎subscript𝑞𝑖\displaystyle=\sum_{i}^{N}\frac{p^{2}_{i}}{2m}+\sum_{i}^{N-1}\sum_{j>i}^{N}% \Phi(\mathbf{r}_{i}-\mathbf{r}_{j})+\sum_{i}^{N}U_{ma}(q_{i}),= ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT divide start_ARG italic_p start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG 2 italic_m end_ARG + ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_j > italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT roman_Φ ( bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - bold_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) + ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_U start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT ( italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) , (6b)

where 𝒦𝒦\mathcal{K}caligraphic_K and 𝒰𝒰\mathcal{U}caligraphic_U are kinetic and potential energy contributions to the Hamiltonian, respectively. m𝑚mitalic_m is mass of the particle, Φ(𝐫i𝐫j)Φsubscript𝐫𝑖subscript𝐫𝑗\Phi(\mathbf{r}_{i}-\mathbf{r}_{j})roman_Φ ( bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - bold_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) is the potential between molecules i𝑖iitalic_i and j𝑗jitalic_j as a function of the spatial coordinate 𝐫𝐫\mathbf{r}bold_r. This intermolecular potential can be approximated as a field defined in terms of the phase-space coordinate Uaa(𝐪)subscript𝑈𝑎𝑎𝐪U_{aa}(\mathbf{q})italic_U start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT ( bold_q ):

iNUaa(qi)iN1j>iNΦ(𝐫i𝐫j).superscriptsubscript𝑖𝑁subscript𝑈𝑎𝑎subscript𝑞𝑖superscriptsubscript𝑖𝑁1superscriptsubscript𝑗𝑖𝑁Φsubscript𝐫𝑖subscript𝐫𝑗\displaystyle\sum_{i}^{N}U_{aa}(q_{i})\equiv\sum_{i}^{N-1}\sum_{j>i}^{N}\Phi(% \mathbf{r}_{i}-\mathbf{r}_{j}).∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_U start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT ( italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) ≡ ∑ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_j > italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT roman_Φ ( bold_r start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT - bold_r start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ) . (7)

II.2.1 Canonical ensemble

Equation (6b) shows that the kinetic-energy term is independent of the position coordinate 𝐪𝐪\mathbf{q}bold_q and the potential-energy term is independent of the momentum coordinate 𝐩𝐩\mathbf{p}bold_p. Assuming that the confining framework is stationary and only fluid particles contribute to the kinetic energy, we separate the variables and define the canonical partition function \mathbb{Z}blackboard_Z as the product of the kinetic contribution Zkaasubscript𝑍𝑘𝑎𝑎Z_{k-aa}italic_Z start_POSTSUBSCRIPT italic_k - italic_a italic_a end_POSTSUBSCRIPT and the configurational contribution Zqsubscript𝑍𝑞Z_{q}italic_Z start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT [refer Eq. 8]. Additionally, the kinetic energy in turn depends only on the temperature of the reservoir. Therefore, we focus on the solution of the configurational partition.

=ZkaaZq,subscript𝑍𝑘𝑎𝑎subscript𝑍𝑞\displaystyle\mathbb{Z}=Z_{k-aa}Z_{q},blackboard_Z = italic_Z start_POSTSUBSCRIPT italic_k - italic_a italic_a end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT , (8)

where the configurational contribution is

Zq=1VNVe𝒰(q1,q2,,qN)/τ𝑑q1𝑑qN,subscript𝑍𝑞1superscript𝑉𝑁subscript𝑉superscript𝑒𝒰subscript𝑞1subscript𝑞2subscript𝑞𝑁𝜏differential-dsubscript𝑞1differential-dsubscript𝑞𝑁\displaystyle Z_{q}=\frac{1}{V^{N}}\int_{V}e^{-\mathcal{U}(q_{1},q_{2},\dots,q% _{N})/\tau}dq_{1}\dots dq_{N},italic_Z start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_V start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT end_ARG ∫ start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - caligraphic_U ( italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_q start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , … , italic_q start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ) / italic_τ end_POSTSUPERSCRIPT italic_d italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT … italic_d italic_q start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT , (9)

with τ=kBT𝜏subscript𝑘B𝑇\tau=k_{\text{B}}Titalic_τ = italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T. The potential energy is a combination of potentials created by adsorbate-adsorbate interaction and the MOF-adsorbate interaction as described in Eq. 10

𝒰(𝐪)=Uaa(𝐪)+Uma(𝐪).𝒰𝐪subscript𝑈𝑎𝑎𝐪subscript𝑈𝑚𝑎𝐪\displaystyle\mathcal{U}(\mathbf{q})=U_{aa}(\mathbf{q})+U_{ma}(\mathbf{q}).caligraphic_U ( bold_q ) = italic_U start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT ( bold_q ) + italic_U start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT ( bold_q ) . (10)

Umasubscript𝑈𝑚𝑎U_{ma}italic_U start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT is the potential energy due to the framework at any given position, which implies

Uma(𝐪)subscript𝑈𝑚𝑎𝐪\displaystyle U_{ma}(\mathbf{q})italic_U start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT ( bold_q ) =Uma(q1)+Uma(q2)++Uma(qN).absentsubscript𝑈𝑚𝑎subscript𝑞1subscript𝑈𝑚𝑎subscript𝑞2subscript𝑈𝑚𝑎subscript𝑞𝑁\displaystyle=U_{ma}(q_{1})+U_{ma}(q_{2})+\dots+U_{ma}(q_{N}).= italic_U start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT ( italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) + italic_U start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT ( italic_q start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) + ⋯ + italic_U start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT ( italic_q start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ) . (11)

The configurational partition function [Eq. (9)] is

Zq=1VNVsubscript𝑍𝑞1superscript𝑉𝑁subscript𝑉\displaystyle Z_{q}=\frac{1}{V^{N}}\int_{V}italic_Z start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_V start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT end_ARG ∫ start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT exp{[Uma(q1)++Uma(qN)\displaystyle\exp\{-[U_{ma}(q_{1})+\cdots+U_{ma}(q_{N})roman_exp { - [ italic_U start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT ( italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) + ⋯ + italic_U start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT ( italic_q start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT )
+\displaystyle++ Uaa(q1)++Uaa(qN)]/τ}dq1dqN.\displaystyle U_{aa}(q_{1})+\cdots+U_{aa}(q_{N})]/\tau\}dq_{1}\cdots dq_{N}.italic_U start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT ( italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) + ⋯ + italic_U start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT ( italic_q start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ) ] / italic_τ } italic_d italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⋯ italic_d italic_q start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT . (12)

For the extreme case in the absence of any external field, where Uma(q1,,qN)=0subscript𝑈𝑚𝑎subscript𝑞1subscript𝑞𝑁0U_{ma}(q_{1},\dots,q_{N})=0italic_U start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT ( italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , … , italic_q start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ) = 0, Eq. (12) resembles the configurational partition function for a bulk fluid.

The complexity due to the inclusion of an external field may be treated in several ways. For example, Travalloni et al. [53] used the extension of the generalized van der Waals theory to model the confined fluids. Simon et al. [54, 55] took as a lattice model the adsorbed gas and the different orientations of the flexible framework and defined a transfer matrix for a rather complex problem. Poluektov [28] analyzed the self-consistent field model for classical systems using a one-dimensional perturbation theory. Singh et al. [56] proposed decoupling the two types of interactions and approximating the solution. Recently, Dong et al. [57] used Gibbs-surface thermodynamics to define the problem in appropriate independent variables and obtained an analytical solution for the special case of confinement between parallel sheets.

We follow the approach of Singh et al. [56] of decoupling the interactions and approximating the effect of the external potential using Mayer’s f𝑓fitalic_f-functions [58, 59]. Let fisubscript𝑓𝑖f_{i}italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT be defined as follows:

fieUma(qi)/τ1.subscript𝑓𝑖superscript𝑒subscript𝑈𝑚𝑎subscript𝑞𝑖𝜏1\displaystyle f_{i}\equiv e^{-U_{ma}(q_{i})/\tau}-1.italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≡ italic_e start_POSTSUPERSCRIPT - italic_U start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT ( italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) / italic_τ end_POSTSUPERSCRIPT - 1 . (13)

Equation (12) can then be written as

Zq=1VNVeUaa(𝐪)/τi=1N(1+fi)dq1dqN,subscript𝑍𝑞1superscript𝑉𝑁subscript𝑉superscript𝑒subscript𝑈𝑎𝑎𝐪𝜏superscriptsubscriptproduct𝑖1𝑁1subscript𝑓𝑖𝑑subscript𝑞1𝑑subscript𝑞𝑁\displaystyle Z_{q}=\frac{1}{V^{N}}\int_{V}e^{-U_{aa}(\mathbf{q})/\tau}\prod_{% i=1}^{N}(1+f_{i})dq_{1}\cdots dq_{N},italic_Z start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_V start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT end_ARG ∫ start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_U start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT ( bold_q ) / italic_τ end_POSTSUPERSCRIPT ∏ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( 1 + italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) italic_d italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⋯ italic_d italic_q start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT , (14)

where

i=1N(1+fi)=1+i=1Nfi+i=1N1j>iNfifj+O(fi3).superscriptsubscriptproduct𝑖1𝑁1subscript𝑓𝑖1superscriptsubscript𝑖1𝑁subscript𝑓𝑖superscriptsubscript𝑖1𝑁1superscriptsubscript𝑗𝑖𝑁subscript𝑓𝑖subscript𝑓𝑗𝑂superscriptsubscript𝑓𝑖3\displaystyle\prod_{i=1}^{N}(1+f_{i})=1+\sum_{i=1}^{N}f_{i}+\sum_{i=1}^{N-1}% \sum_{j>i}^{N}f_{i}f_{j}+{O}(f_{i}^{3}).∏ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT ( 1 + italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) = 1 + ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N - 1 end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_j > italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT + italic_O ( italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) . (15)

Including the individual molecular terms and ignoring the higher-order terms in Eq. (15) yields,

Zq=subscript𝑍𝑞absent\displaystyle Z_{q}=italic_Z start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = 1VNVeUaa(𝐪)/τ(1+i=1Nfi)𝑑q1𝑑qN1superscript𝑉𝑁subscript𝑉superscript𝑒subscript𝑈𝑎𝑎𝐪𝜏1superscriptsubscript𝑖1𝑁subscript𝑓𝑖differential-dsubscript𝑞1differential-dsubscript𝑞𝑁\displaystyle\,\frac{1}{V^{N}}\int_{V}e^{-U_{aa}(\mathbf{q})/\tau}\left(1+\sum% _{i=1}^{N}f_{i}\right)dq_{1}\cdots dq_{N}divide start_ARG 1 end_ARG start_ARG italic_V start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT end_ARG ∫ start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_U start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT ( bold_q ) / italic_τ end_POSTSUPERSCRIPT ( 1 + ∑ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) italic_d italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⋯ italic_d italic_q start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT (16)

This expansion is analogous to the first-order fluid-fluid interactions described by Mayer [59]. In this formulation, Mayer’s f𝑓fitalic_f-functions are applied to simplify the heterogeneous interactions and not the fluid-fluid interactions.

Assuming that, despite the density distribution resulting from the external potential, fluid particles collectively establish a mean field to render the analytical solution tractable, we can write

Uaa(𝐪)=Nuaa,subscript𝑈𝑎𝑎𝐪𝑁subscript𝑢𝑎𝑎\displaystyle U_{aa}(\mathbf{q})=Nu_{aa},italic_U start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT ( bold_q ) = italic_N italic_u start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT , (17)

where uaasubscript𝑢𝑎𝑎u_{aa}italic_u start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT is the mean field independent of the positional coordinate in phase space. Thus, uaasubscript𝑢𝑎𝑎u_{aa}italic_u start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT can be moved outside the integral and Eq. (16) takes the form

Zq=subscript𝑍𝑞absent\displaystyle Z_{q}=italic_Z start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT = 1VNVeNuaa(𝐪)/τ𝑑q1𝑑qN1superscript𝑉𝑁subscript𝑉superscript𝑒𝑁subscript𝑢𝑎𝑎𝐪𝜏differential-dsubscript𝑞1differential-dsubscript𝑞𝑁\displaystyle\,\frac{1}{V^{N}}\int_{V}e^{-Nu_{aa}(\mathbf{q})/\tau}dq_{1}% \cdots dq_{N}divide start_ARG 1 end_ARG start_ARG italic_V start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT end_ARG ∫ start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - italic_N italic_u start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT ( bold_q ) / italic_τ end_POSTSUPERSCRIPT italic_d italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⋯ italic_d italic_q start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT
+eNuaa(𝐪)/τV(Vf1𝑑q1++VfN𝑑qN).superscript𝑒𝑁subscript𝑢𝑎𝑎𝐪𝜏𝑉subscript𝑉subscript𝑓1differential-dsubscript𝑞1subscript𝑉subscript𝑓𝑁differential-dsubscript𝑞𝑁\displaystyle+\frac{e^{-Nu_{aa}(\mathbf{q})/\tau}}{V}\left(\int_{V}f_{1}dq_{1}% +\cdots+\int_{V}f_{N}dq_{N}\right).+ divide start_ARG italic_e start_POSTSUPERSCRIPT - italic_N italic_u start_POSTSUBSCRIPT italic_a italic_a end_POSTSUBSCRIPT ( bold_q ) / italic_τ end_POSTSUPERSCRIPT end_ARG start_ARG italic_V end_ARG ( ∫ start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_d italic_q start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + ⋯ + ∫ start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT italic_d italic_q start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT ) . (18)

The first term in Eq. (18) describes the configurational partition function of the bulk phase in the absence of any external field; let it be Zqaasubscript𝑍𝑞𝑎𝑎Z_{q-aa}italic_Z start_POSTSUBSCRIPT italic_q - italic_a italic_a end_POSTSUBSCRIPT. In addition, we can define

ϕitalic-ϕ\displaystyle\phiitalic_ϕ 1VVfi𝑑qiabsent1𝑉subscript𝑉subscript𝑓𝑖differential-dsubscript𝑞𝑖\displaystyle\equiv\frac{1}{V}\int_{V}f_{i}dq_{i}≡ divide start_ARG 1 end_ARG start_ARG italic_V end_ARG ∫ start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT italic_d italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT
=1VV(eUma(qi)/τ1)𝑑qi.absent1𝑉subscript𝑉superscript𝑒subscript𝑈𝑚𝑎subscript𝑞𝑖𝜏1differential-dsubscript𝑞𝑖\displaystyle=\frac{1}{V}\int_{V}\left(e^{-U_{ma}(q_{i})/\tau}-1\right)dq_{i}.= divide start_ARG 1 end_ARG start_ARG italic_V end_ARG ∫ start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT ( italic_e start_POSTSUPERSCRIPT - italic_U start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT ( italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ) / italic_τ end_POSTSUPERSCRIPT - 1 ) italic_d italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT . (19)

Using this definition,

ZqZqaa(1+Nϕ),subscript𝑍𝑞subscript𝑍𝑞𝑎𝑎1𝑁italic-ϕ\displaystyle Z_{q}\approx Z_{q-aa}\left(1+N\phi\right),italic_Z start_POSTSUBSCRIPT italic_q end_POSTSUBSCRIPT ≈ italic_Z start_POSTSUBSCRIPT italic_q - italic_a italic_a end_POSTSUBSCRIPT ( 1 + italic_N italic_ϕ ) , (20)

where Zqaasubscript𝑍𝑞𝑎𝑎Z_{q-aa}italic_Z start_POSTSUBSCRIPT italic_q - italic_a italic_a end_POSTSUBSCRIPT is the configurational partition function of the mean-field bulk fluid. Putting this back into Eq. (8) gives

\displaystyle\mathbb{Z}blackboard_Z =ZkaaZqaa(1+Nϕ)absentsubscript𝑍𝑘𝑎𝑎subscript𝑍𝑞𝑎𝑎1𝑁italic-ϕ\displaystyle=Z_{k-aa}Z_{q-aa}\left(1+N\phi\right)= italic_Z start_POSTSUBSCRIPT italic_k - italic_a italic_a end_POSTSUBSCRIPT italic_Z start_POSTSUBSCRIPT italic_q - italic_a italic_a end_POSTSUBSCRIPT ( 1 + italic_N italic_ϕ )
=bulk(1+Nϕ).absentsubscriptbulk1𝑁italic-ϕ\displaystyle=\mathbb{Z}_{\text{bulk}}\left(1+N\phi\right).= blackboard_Z start_POSTSUBSCRIPT bulk end_POSTSUBSCRIPT ( 1 + italic_N italic_ϕ ) . (21)

Given that Umasubscript𝑈𝑚𝑎U_{ma}italic_U start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT is a spatially varying potential inside the mesopores, we consider an infinitesimal volume dV=dxdydz𝑑𝑉𝑑𝑥𝑑𝑦𝑑𝑧dV=dxdydzitalic_d italic_V = italic_d italic_x italic_d italic_y italic_d italic_z at coordinate qi=(x,y,z)subscript𝑞𝑖𝑥𝑦𝑧q_{i}=(x,y,z)italic_q start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT = ( italic_x , italic_y , italic_z ). We assume a uniform potential Uma(x,y,z)subscript𝑈𝑚𝑎𝑥𝑦𝑧U_{ma}(x,y,z)italic_U start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT ( italic_x , italic_y , italic_z ) in the infinitesimal volume, which implies that the canonical partition function for an infinitesimal volume is

=bulk[1+Nϕ(x,y,z)].subscriptbulkdelimited-[]1𝑁italic-ϕ𝑥𝑦𝑧\displaystyle\mathbb{Z}=\mathbb{Z}_{\text{bulk}}[1+N\phi(x,y,z)].blackboard_Z = blackboard_Z start_POSTSUBSCRIPT bulk end_POSTSUBSCRIPT [ 1 + italic_N italic_ϕ ( italic_x , italic_y , italic_z ) ] . (22)

II.2.2 Grand canonical ensemble

The equilibrium assumption implies that the chemical potential μbulksuperscript𝜇bulk\mu^{\text{bulk}}italic_μ start_POSTSUPERSCRIPT bulk end_POSTSUPERSCRIPT of the bulk phase equals the chemical potential μtotalsuperscript𝜇total\mu^{\text{total}}italic_μ start_POSTSUPERSCRIPT total end_POSTSUPERSCRIPT of the argon inside the nanospace. The chemical potential μtotalsuperscript𝜇total\mu^{\text{total}}italic_μ start_POSTSUPERSCRIPT total end_POSTSUPERSCRIPT of argon inside the nanospace consists of contributions from other argon inside nanospace (μadssuperscript𝜇ads\mu^{\text{ads}}italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT) and from the framework atoms (μframesuperscript𝜇frame\mu^{\text{frame}}italic_μ start_POSTSUPERSCRIPT frame end_POSTSUPERSCRIPT) that form the heterogeneity. Both contributions are made through intermolecular forces, not intramolecular forces:

μbulk(pext,T)superscript𝜇bulksubscript𝑝ext𝑇\displaystyle\mu^{\text{bulk}}(p_{\text{ext}},T)italic_μ start_POSTSUPERSCRIPT bulk end_POSTSUPERSCRIPT ( italic_p start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT , italic_T ) =μtotalabsentsuperscript𝜇total\displaystyle=\mu^{\text{total}}= italic_μ start_POSTSUPERSCRIPT total end_POSTSUPERSCRIPT (23a)
=μads+μframe.absentsuperscript𝜇adssuperscript𝜇frame\displaystyle=\mu^{\text{ads}}+\mu^{\text{frame}}.= italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT + italic_μ start_POSTSUPERSCRIPT frame end_POSTSUPERSCRIPT . (23b)

Similar to the excess chemical potential defined by Widom [60], the chemical potential of the adsorbed phase combines the intramolecular chemical potential μadssuperscript𝜇ads\mu^{\text{ads}}italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT from the adsorbed phase and the excess chemical potential μframesuperscript𝜇frame\mu^{\text{frame}}italic_μ start_POSTSUPERSCRIPT frame end_POSTSUPERSCRIPT due to the framework. We use the mean value of the interaction energies in the unit cell to calculate the excess chemical potential μframesuperscript𝜇frame\mu^{\text{frame}}italic_μ start_POSTSUPERSCRIPT frame end_POSTSUPERSCRIPT:

μframe=τlnexp(Uma(𝐪)τ).superscript𝜇frame𝜏subscript𝑈𝑚𝑎𝐪𝜏\mu^{\text{frame}}=-\tau\ln\left\langle\exp\left(\frac{-U_{ma}(\mathbf{q})}{% \tau}\right)\right\rangle.italic_μ start_POSTSUPERSCRIPT frame end_POSTSUPERSCRIPT = - italic_τ roman_ln ⟨ roman_exp ( divide start_ARG - italic_U start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT ( bold_q ) end_ARG start_ARG italic_τ end_ARG ) ⟩ . (24)

In addition, the unit-cell volume V𝑉Vitalic_V is fixed and the temperature T𝑇Titalic_T is controlled externally. Therefore, the modeling is done in a grand canonical ensemble (μads,V,Tsuperscript𝜇ads𝑉𝑇\mu^{\text{ads}},V,Titalic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT , italic_V , italic_T). Moreover, if a partition function of a system of particles can be obtained, the relevant quantities of interest, such as density, pressure, entropy, and free energy can be derived. To this end, we start the model by defining the grand partition function ΞadssubscriptΞads\Xi_{\text{ads}}roman_Ξ start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT for the adsorbed fluid [52]:

ΞadssubscriptΞads\displaystyle\Xi_{\text{ads}}roman_Ξ start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT =N=0dN𝐪dN𝐩h3NN!e[(𝐩,𝐪)μadsN]/τabsentsuperscriptsubscript𝑁0superscript𝑑𝑁𝐪superscript𝑑𝑁𝐩superscript3𝑁𝑁superscript𝑒delimited-[]𝐩𝐪superscript𝜇ads𝑁𝜏\displaystyle=\sum_{N=0}^{\infty}\int\frac{d^{N}\mathbf{q}d^{N}\mathbf{p}}{h^{% 3N}N!}e^{-[\mathcal{H}(\mathbf{p},\mathbf{q})-\mu^{\text{ads}}N]/\tau}= ∑ start_POSTSUBSCRIPT italic_N = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT ∫ divide start_ARG italic_d start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT bold_q italic_d start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT bold_p end_ARG start_ARG italic_h start_POSTSUPERSCRIPT 3 italic_N end_POSTSUPERSCRIPT italic_N ! end_ARG italic_e start_POSTSUPERSCRIPT - [ caligraphic_H ( bold_p , bold_q ) - italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT italic_N ] / italic_τ end_POSTSUPERSCRIPT (25a)
=N=0eNμads/τ,absentsuperscriptsubscript𝑁0superscript𝑒𝑁superscript𝜇ads𝜏\displaystyle=\sum_{N=0}^{\infty}\mathbb{Z}e^{N\mu^{\text{ads}}/\tau},= ∑ start_POSTSUBSCRIPT italic_N = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT blackboard_Z italic_e start_POSTSUPERSCRIPT italic_N italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT / italic_τ end_POSTSUPERSCRIPT , (25b)

where hhitalic_h is Planck’s constant and \mathbb{Z}blackboard_Z is the canonical partition function for an ensemble of N𝑁Nitalic_N molecules. Combining Eqs. (22) and (25b) gives

Ξads(μads,V,T)=Ξbulk(μads,V,T)(1+ϕNbulkμads).subscriptΞadssuperscript𝜇ads𝑉𝑇subscriptΞbulksuperscript𝜇ads𝑉𝑇1italic-ϕsubscriptsuperscriptdelimited-⟨⟩𝑁superscript𝜇adsbulk\displaystyle\Xi_{\text{ads}}(\mu^{\text{ads}},V,T)=\Xi_{\text{bulk}}(\mu^{% \text{ads}},V,T)\left(1+\phi\langle N\rangle^{\mu^{\text{ads}}}_{\text{bulk}}% \right).roman_Ξ start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ( italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT , italic_V , italic_T ) = roman_Ξ start_POSTSUBSCRIPT bulk end_POSTSUBSCRIPT ( italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT , italic_V , italic_T ) ( 1 + italic_ϕ ⟨ italic_N ⟩ start_POSTSUPERSCRIPT italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bulk end_POSTSUBSCRIPT ) . (26)

Under extreme conditions where there is no external potential (i.e., no confinement), the partition function ΞadssubscriptΞads\Xi_{\text{ads}}roman_Ξ start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT simplifies to the partition function ΞbulksubscriptΞbulk\Xi_{\text{bulk}}roman_Ξ start_POSTSUBSCRIPT bulk end_POSTSUBSCRIPT of the bulk fluid. This ensures the consistency and coherence of the equation, particularly in scenarios where confinement effects are negligible.

II.3 Thermodynamic properties

When examining intensive thermodynamic functions for a bulk fluid, such as pressure or chemical potential, it is common to assume that the fluid is homogeneous, meaning that the properties of the fluid are uniform throughout the entire volume under consideration. However, in the context of confined fluids, the presence of heterogeneous interactions introduces nonuniformities in the intensive thermodynamic functions. To address this distribution of properties, Hill introduced both differential and integral thermodynamic functions [61, 52]. Differential thermodynamic functions are defined at a specific point in space, indicating their local nature. Conversely, integral thermodynamic functions extend their definition across the entire volume of the system, providing a global characterization.

Given that the phase of any substance is defined for a group of molecules [62, 63], understanding the phase of the fluid requires that the integral properties be considered. Therefore, in the exploration of phase transitions, emphasis is placed on global (i.e., integral) thermodynamic properties rather than local (i.e., differential) thermodynamic properties.

Relevant thermodynamic properties such as the grand potential Ω¯adssubscript¯Ωads\bar{\Omega}_{\text{ads}}over¯ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT, the expected number Nads¯¯delimited-⟨⟩subscript𝑁ads\overline{\langle N_{\text{ads}}\rangle}over¯ start_ARG ⟨ italic_N start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ⟩ end_ARG of molecules adsorbed, the pressure p^adssubscript^𝑝ads\hat{p}_{\text{ads}}over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT of the adsorbed phase, entropy S¯adssubscript¯𝑆ads\bar{S}_{\text{ads}}over¯ start_ARG italic_S end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT, enthalpy H¯adssubscript¯𝐻ads\bar{H}_{\text{ads}}over¯ start_ARG italic_H end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT, Helmholtz free energy F¯adssubscript¯𝐹ads\bar{F}_{\text{ads}}over¯ start_ARG italic_F end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT, and Gibbs free energy G¯adssubscript¯𝐺ads\bar{G}_{\text{ads}}over¯ start_ARG italic_G end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT can all be obtained from the grand partition function ΞadssubscriptΞads\Xi_{\text{ads}}roman_Ξ start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT, as shown in the following section.

II.3.1 Grand potential Ω¯adssubscript¯Ωads\bar{\Omega}_{\text{ads}}over¯ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT

The grand potential can be obtained from its definition

Ωads(x,y,z)subscriptΩads𝑥𝑦𝑧\displaystyle\Omega_{\text{ads}}(x,y,z)roman_Ω start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ( italic_x , italic_y , italic_z ) =τln[Ξads(x,y,z)]absent𝜏subscriptΞads𝑥𝑦𝑧\displaystyle=-\tau\ln[\Xi_{\text{ads}}(x,y,z)]= - italic_τ roman_ln [ roman_Ξ start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ( italic_x , italic_y , italic_z ) ]
=τln(Ξbulk)τln[1+ϕ(x,y,z)]absent𝜏subscriptΞbulk𝜏1italic-ϕ𝑥𝑦𝑧\displaystyle=-\tau\ln(\Xi_{\text{bulk}})-\tau\ln[1+\phi(x,y,z)]= - italic_τ roman_ln ( start_ARG roman_Ξ start_POSTSUBSCRIPT bulk end_POSTSUBSCRIPT end_ARG ) - italic_τ roman_ln [ 1 + italic_ϕ ( italic_x , italic_y , italic_z ) ]
=Ωbulkτln[1+ϕ(x,y,z)].absentsubscriptΩbulk𝜏1italic-ϕ𝑥𝑦𝑧\displaystyle=\Omega_{\text{bulk}}-\tau\ln[1+\phi(x,y,z)].= roman_Ω start_POSTSUBSCRIPT bulk end_POSTSUBSCRIPT - italic_τ roman_ln [ 1 + italic_ϕ ( italic_x , italic_y , italic_z ) ] . (27)

where ΩbulksubscriptΩbulk\Omega_{\text{bulk}}roman_Ω start_POSTSUBSCRIPT bulk end_POSTSUBSCRIPT is the grand potential of the bulk fluid. Since we are interested in the integral properties inside the MOF pore, we spatially average the grand potential:

Ω¯adssubscript¯Ωads\displaystyle\bar{\Omega}_{\text{ads}}over¯ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT =1VVΩads(x,y,z)𝑑x𝑑y𝑑zabsent1𝑉subscript𝑉subscriptΩads𝑥𝑦𝑧differential-d𝑥differential-d𝑦differential-d𝑧\displaystyle=\frac{1}{V}\int_{V}\Omega_{\text{ads}}(x,y,z)dxdydz= divide start_ARG 1 end_ARG start_ARG italic_V end_ARG ∫ start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT roman_Ω start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ( italic_x , italic_y , italic_z ) italic_d italic_x italic_d italic_y italic_d italic_z
=p^Vξ.absent^𝑝𝑉𝜉\displaystyle=\hat{p}V\equiv\xi.= over^ start_ARG italic_p end_ARG italic_V ≡ italic_ξ . (28)

II.3.2 Pressure p^adssubscript^𝑝ads\hat{p}_{\text{ads}}over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT of adsorbed phase

For any macroscopic system, we can write

EtTStμNt=τlnΞt.subscript𝐸𝑡𝑇subscript𝑆𝑡𝜇subscript𝑁𝑡𝜏subscriptΞ𝑡\displaystyle E_{t}-TS_{t}-\mu N_{t}=-\tau\ln\Xi_{t}.italic_E start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT - italic_T italic_S start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT - italic_μ italic_N start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = - italic_τ roman_ln roman_Ξ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT . (29)

Combining Eqs. (3), (4), and (29), we obtain the integral pressure p^adssubscript^𝑝ads\hat{p}_{\text{ads}}over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT of the adsorbed fluid in terms of the grand partition function:

ξp^adsV=τlnΞads=Ω¯ads.𝜉subscript^𝑝ads𝑉𝜏subscriptΞadssubscript¯Ωads\displaystyle\xi\equiv\hat{p}_{\text{ads}}V=\tau\ln\Xi_{\text{ads}}=-\bar{% \Omega}_{\text{ads}}.italic_ξ ≡ over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT italic_V = italic_τ roman_ln roman_Ξ start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT = - over¯ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT . (30)

Therefore,

p^ads=(Ω¯ads)μ,TV.subscript^𝑝adssubscriptsubscript¯Ωads𝜇𝑇𝑉\displaystyle\hat{p}_{\text{ads}}=-\frac{(\bar{\Omega}_{\text{ads}})_{\mu,T}}{% V}.over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT = - divide start_ARG ( over¯ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_μ , italic_T end_POSTSUBSCRIPT end_ARG start_ARG italic_V end_ARG . (31)

II.3.3 Expected number Nads¯¯delimited-⟨⟩subscript𝑁ads\overline{\langle N_{\text{ads}}\rangle}over¯ start_ARG ⟨ italic_N start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ⟩ end_ARG of molecules adsorbed

The expected number Nads¯¯delimited-⟨⟩subscript𝑁ads\overline{\langle N_{\text{ads}}\rangle}over¯ start_ARG ⟨ italic_N start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ⟩ end_ARG of molecules adsorbed can be calculated as follows:

Nads(x,y,z)delimited-⟨⟩subscript𝑁ads𝑥𝑦𝑧\displaystyle\langle N_{\text{ads}}(x,y,z)\rangle⟨ italic_N start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ( italic_x , italic_y , italic_z ) ⟩ =NNadseμadsN/τΞadsabsentsubscript𝑁𝑁subscriptadssuperscript𝑒superscript𝜇ads𝑁𝜏subscriptΞads\displaystyle=\frac{\sum_{N}N\mathbb{Z}_{\text{ads}}e^{\mu^{\text{ads}}N/\tau}% }{\Xi_{\text{ads}}}= divide start_ARG ∑ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT italic_N blackboard_Z start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT italic_N / italic_τ end_POSTSUPERSCRIPT end_ARG start_ARG roman_Ξ start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT end_ARG (32a)
=Nbulkμads+N2bulkμadsϕ(x,y,z)1+Nbulkμadsϕ(x,y,z),absentsubscriptsuperscriptdelimited-⟨⟩𝑁superscript𝜇adsbulksubscriptsuperscriptdelimited-⟨⟩superscript𝑁2superscript𝜇adsbulkitalic-ϕ𝑥𝑦𝑧1subscriptsuperscriptdelimited-⟨⟩𝑁superscript𝜇adsbulkitalic-ϕ𝑥𝑦𝑧\displaystyle=\frac{\langle N\rangle^{\mu^{\text{ads}}}_{\text{bulk}}+\langle N% ^{2}\rangle^{\mu^{\text{ads}}}_{\text{bulk}}\phi(x,y,z)}{1+\langle N\rangle^{% \mu^{\text{ads}}}_{\text{bulk}}\phi(x,y,z)},= divide start_ARG ⟨ italic_N ⟩ start_POSTSUPERSCRIPT italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bulk end_POSTSUBSCRIPT + ⟨ italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ start_POSTSUPERSCRIPT italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bulk end_POSTSUBSCRIPT italic_ϕ ( italic_x , italic_y , italic_z ) end_ARG start_ARG 1 + ⟨ italic_N ⟩ start_POSTSUPERSCRIPT italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bulk end_POSTSUBSCRIPT italic_ϕ ( italic_x , italic_y , italic_z ) end_ARG , (32b)

where Nbulkμadssubscriptsuperscriptdelimited-⟨⟩𝑁superscript𝜇adsbulk\langle N\rangle^{\mu^{\text{ads}}}_{\text{bulk}}⟨ italic_N ⟩ start_POSTSUPERSCRIPT italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bulk end_POSTSUBSCRIPT is the average number of molecules that would be present in the bulk if the chemical potential were μadssuperscript𝜇ads\mu^{\text{ads}}italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT. Note that the term Nads(x,y,z)delimited-⟨⟩subscript𝑁ads𝑥𝑦𝑧\langle N_{\text{ads}}(x,y,z)\rangle⟨ italic_N start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ( italic_x , italic_y , italic_z ) ⟩ is not the actual number of molecules at a given position but the expected number. Therefore, the expected total number of molecules in the unit cell is

Nads¯=1VVNads(x,y,z)𝑑x𝑑y𝑑z.¯delimited-⟨⟩subscript𝑁ads1𝑉subscript𝑉delimited-⟨⟩subscript𝑁ads𝑥𝑦𝑧differential-d𝑥differential-d𝑦differential-d𝑧\displaystyle\overline{\langle N_{\text{ads}}\rangle}=\frac{1}{V}\int_{V}% \langle N_{\text{ads}}(x,y,z)\rangle dxdydz.over¯ start_ARG ⟨ italic_N start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ⟩ end_ARG = divide start_ARG 1 end_ARG start_ARG italic_V end_ARG ∫ start_POSTSUBSCRIPT italic_V end_POSTSUBSCRIPT ⟨ italic_N start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ( italic_x , italic_y , italic_z ) ⟩ italic_d italic_x italic_d italic_y italic_d italic_z . (33)

The number of molecules confined within a system is influenced by two factors: intermolecular interactions and heterogeneous interactions. The maximum number of molecules is capped by the volume of the unit cell. These interactions collectively contribute to the effective potential experienced by the adsorbate molecules. Equation (26) uses a decoupling approach to separate these two interactions. The intermolecular interaction among the confined molecules is considered independently, following which the heterogeneous interactions are incorporated as an additional potential term. The volume constraint is accounted for by capping the summation in Eq. (II.3.3) at Nmaxsubscript𝑁maxN_{\text{max}}italic_N start_POSTSUBSCRIPT max end_POSTSUBSCRIPT such that

Nmax=VcellVmVlb,subscript𝑁maxsubscript𝑉cellsubscript𝑉msubscript𝑉l𝑏\displaystyle N_{\text{max}}=\left\lfloor\frac{V_{\text{cell}}-V_{\text{m}}-V_% {\text{l}}}{b}\right\rfloor,italic_N start_POSTSUBSCRIPT max end_POSTSUBSCRIPT = ⌊ divide start_ARG italic_V start_POSTSUBSCRIPT cell end_POSTSUBSCRIPT - italic_V start_POSTSUBSCRIPT m end_POSTSUBSCRIPT - italic_V start_POSTSUBSCRIPT l end_POSTSUBSCRIPT end_ARG start_ARG italic_b end_ARG ⌋ , (34)

where \lfloor\cdot\rfloor⌊ ⋅ ⌋ is the floor function, Vcellsubscript𝑉cellV_{\text{cell}}italic_V start_POSTSUBSCRIPT cell end_POSTSUBSCRIPT is the unit-cell volume, Vmsubscript𝑉mV_{\text{m}}italic_V start_POSTSUBSCRIPT m end_POSTSUBSCRIPT and Vlsubscript𝑉lV_{\text{l}}italic_V start_POSTSUBSCRIPT l end_POSTSUBSCRIPT are the volume of the unit cell occupied by the metal and ligand, respectively, b=2σa3𝑏2superscriptsubscript𝜎𝑎3b=\sqrt{2}\sigma_{a}^{3}italic_b = square-root start_ARG 2 end_ARG italic_σ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT is the volume of each molecule according to van der Waals theory, and σasubscript𝜎𝑎\sigma_{a}italic_σ start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT is the Lennard–Jones size parameter for adsorbates.

II.3.4 Other thermodynamic functions

The following standard thermodynamic relations still apply to the integral values:

Entropy:(S¯ads)Entropy:subscript¯𝑆ads\displaystyle\text{Entropy:}~{}(\bar{S}_{\text{ads}})Entropy: ( over¯ start_ARG italic_S end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ) =(Ω¯adsT)p,μads,absentsubscriptpartial-derivative𝑇subscript¯Ωads𝑝superscript𝜇ads\displaystyle=\Bigl{(}\partialderivative{\bar{\Omega}_{\text{ads}}}{T}\Bigr{)}% _{p,\mu^{\text{ads}}},= ( divide start_ARG ∂ start_ARG over¯ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT end_ARG end_ARG start_ARG ∂ start_ARG italic_T end_ARG end_ARG ) start_POSTSUBSCRIPT italic_p , italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT end_POSTSUBSCRIPT , (35)
Gibbs free energy:(G¯ads)Gibbs free energy:subscript¯𝐺ads\displaystyle\text{Gibbs free energy:}~{}(\bar{G}_{\text{ads}})Gibbs free energy: ( over¯ start_ARG italic_G end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ) =μadsNads¯,absentsuperscript𝜇ads¯delimited-⟨⟩subscript𝑁ads\displaystyle=\mu^{\text{ads}}\overline{\langle N_{\text{ads}}\rangle},= italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT over¯ start_ARG ⟨ italic_N start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ⟩ end_ARG , (36)
Helmholtz free energy:(F¯ads)Helmholtz free energy:subscript¯𝐹ads\displaystyle\text{Helmholtz free energy:}~{}(\bar{F}_{\text{ads}})Helmholtz free energy: ( over¯ start_ARG italic_F end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ) =Ω¯ads+μadsNads¯,absentsubscript¯Ωadssuperscript𝜇ads¯delimited-⟨⟩subscript𝑁ads\displaystyle=\bar{\Omega}_{\text{ads}}+\mu^{\text{ads}}\overline{\langle N_{% \text{ads}}\rangle},= over¯ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT + italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT over¯ start_ARG ⟨ italic_N start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ⟩ end_ARG , (37)
Enthalpy:(H¯ads)Enthalpy:subscript¯𝐻ads\displaystyle\text{Enthalpy:}~{}(\bar{H}_{\text{ads}})Enthalpy: ( over¯ start_ARG italic_H end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ) =G¯ads+TS¯ads.absentsubscript¯𝐺ads𝑇subscript¯𝑆ads\displaystyle=\bar{G}_{\text{ads}}+T\bar{S}_{\text{ads}}.= over¯ start_ARG italic_G end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT + italic_T over¯ start_ARG italic_S end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT . (38)

III Results and Discussions

III.1 Benchmarking

Refer to caption
Figure 2: Benchmarking of the model with GCMC simulations. (a) Model metal-organic framework (MOF) with metal oxides and ligands; a𝑎aitalic_a is the unit-cell length. (b) Potential-energy distribution inside the unit cell of this MOF. (c) Adsorption isotherm for argon obtained from the model (green line) and the GCMC simulations (pink diamonds). Capillary condensation occurs at relative pressure p/p00.6𝑝subscript𝑝00.6p/p_{\text{0}}\approx 0.6italic_p / italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ 0.6. The insets (i)–(iii) show the distribution of argon atoms inside the unit cell obtained from the GCMC simulation and the expected relative density distribution (normalized with the maximum local density for better contrast) obtained from the proposed model presented at the corresponding locations in the isotherm. The benchmarking is done for argon adsorption in a model MOF with unit-cell length a=24𝑎24a=24italic_a = 24 Å, LJ parameters σm=5subscript𝜎𝑚5\sigma_{m}=5italic_σ start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = 5 Å, εm=120subscript𝜀𝑚120\varepsilon_{m}=120italic_ε start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = 120 K, and temperature T=120𝑇120T=120italic_T = 120 K.

We assume that bulk argon behaves as a van der Waals fluid, so we can write the canonical partition function of the bulk fluid as

bulk=1N!(VNbλT3)Nexp(aN2Vτ),subscriptbulk1𝑁superscript𝑉𝑁𝑏subscriptsuperscript𝜆3𝑇𝑁𝑎superscript𝑁2𝑉𝜏\displaystyle\mathbb{Z}_{\text{bulk}}=\frac{1}{N!}\left(\frac{V-Nb}{\lambda^{3% }_{T}}\right)^{N}\exp\left(\frac{-aN^{2}}{V\tau}\right),blackboard_Z start_POSTSUBSCRIPT bulk end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_N ! end_ARG ( divide start_ARG italic_V - italic_N italic_b end_ARG start_ARG italic_λ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT roman_exp ( divide start_ARG - italic_a italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_V italic_τ end_ARG ) , (39)

where a𝑎aitalic_a and b𝑏bitalic_b are the van-der Waals coefficients for argon and λTsubscript𝜆𝑇\lambda_{T}italic_λ start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT is the thermal de Broglie wavelength [64]. Based on this assumption various cases with different pore sizes and temperatures were analyzed using an in-house GPU-accelerated Python code (available upon request).

This proposed model is benchmarked using a GCMC simulation for a model MOF with a cubic unit cell. Metal oxides occupy the vertices of the cube and ligands are located on the edges of the cube, as shown in Fig. 2(a). To maintain the consistency with the statistical model, we used averaged LJ parameters for metal-oxide and aromatic rings to model this framework [65]. The GCMC simulations were performed using the RASPA [66] simulation software package. All simulations included a 50 000-cycle equilibration period and a 100 000-cycle production run. In these simulations, the structure of all the frameworks is considered rigid; that is, all species of the framework are held fixed at their crystallographic positions. The argon atoms can move in three different ways in the GCMC simulation: translation, rotation, and swap. The interaction between MOF and argon and between argon atoms was modeled using the LJ potential function and the Lorentz–Berthelot mixing rule. Sample crystallographic information (.cif) files and the LJ parameters are listed in the electronic supporting information (ESI) [67] Sec. S1 A.

Figure 2 shows the benchmarking results for the 24 Å unit cell. Figures 2(a) and 2(b) show the unit-cell structure and the potential-energy distribution, respectively. Figure 2(c) compares the adsorption isotherm derived from the GCMC simulation with that produced by the proposed model, showing that the two curves are consistent. Additionally, Fig. 2(c) shows in the three lower panels the molecular distribution obtained through the GCMC simulation along with the density distribution within the pore at positions (i), (ii), and (iii). The proposed model correctly captures the previously observed trend of layered adsorption [14, 15, 16, 17, 18]. Specifically, at lower relative pressures, adsorption predominantly occurs near the heterogeneity, as shown in Fig. 2(c)(i). As the relative pressure increases but before capillary condensation, a distinct layering of adsorbed molecules is evident in Fig. 2(c)(ii). Finally, beyond capillary condensation, the pore becomes saturated with a density distribution resembling that of the bulk liquid [Fig. 2(c)(iii)]. Additional benchmarking results are presented in Sec. S1. Moreover, Sec. S2 of the ESI [67] provides a brief parametric investigation that supports the hypothesis made in our previous paper [68] that the thermodynamic properties of confined fluids are a function of the confinement parameter Ψσma/aΨsubscript𝜎𝑚𝑎𝑎\Psi\equiv\sigma_{ma}/aroman_Ψ ≡ italic_σ start_POSTSUBSCRIPT italic_m italic_a end_POSTSUBSCRIPT / italic_a.

Figure S1 of the ESI [67] shows that the isotherms for ultrasmall pore size (10 Å) and larger pore sizes (24 Å) are consistent with the results obtained from GCMC simulations. This result is attributed to the assumption of a uniform field generated by the fluid molecules adsorbed in the cavity. In ultrasmall pores, the variability in the field resulting from argon adsorption at different positions is effectively equivalent to a uniform distribution. Similarly, for larger pore sizes, the distribution of molecules near adsorption sites remains independent at lower concentrations, thereby creating a uniform field inside the cavity, as depicted in Fig. 2(c)(i). At higher concentrations (that is, post capillary condensation), argon molecules densely occupy the cavity, resulting in a uniform field, as shown in Fig. 2(c)(iii). For concentrations between these extremes, the adsorption isotherm slightly deviates from the GCMC values, as shown in Fig. 2(c)(ii).

Likewise, for the medium pore sizes shown in Figs. S1(b)–S1(e) of the ESI [67], the isotherm produced by the proposed model deviates from the GCMC isotherm. This discrepancy arises when the assumption of a uniform field in the cavity is no longer valid. This problem could be addressed through an iterative process, where the density distribution obtained from the proposed model serves as the initial guess. However, this problem is beyond the scope of present paper.

III.2 Phase transition in confinement

III.2.1 Types of phase transitions

Since the inception of fullerenes and 3D carbon nanotubes with cylindrical pores, the phase-transitions inside these structures have been discussed [69]. Multiple studies show that freezing of water confined in these structures may occur continuously or discontinuously [70, 71]. The results in this section show that this transition depends on the pore size. Smaller pores produce continuous phase transitions, whereas larger pores produce discontinuous (first-order) phase transitions.

The Helmholtz free energy can be expressed in terms of the canonical partition function [Eq. (22)]:

Fads(N,V,T)subscript𝐹ads𝑁𝑉𝑇\displaystyle F_{\text{ads}}(N,V,T)italic_F start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT ( italic_N , italic_V , italic_T ) =τln[bulk(1+Nϕ)]absent𝜏subscriptbulk1𝑁italic-ϕ\displaystyle=-\tau\ln[\mathbb{Z}_{\text{bulk}}(1+N\phi)\bigr{]}= - italic_τ roman_ln [ blackboard_Z start_POSTSUBSCRIPT bulk end_POSTSUBSCRIPT ( 1 + italic_N italic_ϕ ) ]
=Nτ{ln[(VNb)λT3N]+1}N2aVabsent𝑁𝜏𝑉𝑁𝑏subscriptsuperscript𝜆3𝑇𝑁1superscript𝑁2𝑎𝑉\displaystyle=-N\tau\left\{\ln\left[\frac{(V-Nb)}{\lambda^{3}_{T}N}\right]+1% \right\}-\frac{N^{2}a}{V}= - italic_N italic_τ { roman_ln [ divide start_ARG ( italic_V - italic_N italic_b ) end_ARG start_ARG italic_λ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_T end_POSTSUBSCRIPT italic_N end_ARG ] + 1 } - divide start_ARG italic_N start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_a end_ARG start_ARG italic_V end_ARG
τln[1+Nϕ(x,y,z)].𝜏1𝑁italic-ϕ𝑥𝑦𝑧\displaystyle-\tau\ln[1+N\phi(x,y,z)\bigr{]}.- italic_τ roman_ln [ 1 + italic_N italic_ϕ ( italic_x , italic_y , italic_z ) ] . (40)

Therefore, we define the differential chemical potential μadssuperscript𝜇ads\mu^{\text{ads}}italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT and integral chemical potential μ^adssuperscript^𝜇ads\hat{\mu}^{\text{ads}}over^ start_ARG italic_μ end_ARG start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT as follows:

μads=(FadsN)V,T,superscript𝜇adssubscriptpartial-derivative𝑁subscript𝐹ads𝑉𝑇\displaystyle\mu^{\text{ads}}=\left(\partialderivative{F_{\text{ads}}}{N}% \right)_{V,T},italic_μ start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT = ( divide start_ARG ∂ start_ARG italic_F start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT end_ARG end_ARG start_ARG ∂ start_ARG italic_N end_ARG end_ARG ) start_POSTSUBSCRIPT italic_V , italic_T end_POSTSUBSCRIPT ,
μ^ads=(F¯adsN)V,T=(F¯adsN)V,T.superscript^𝜇adssubscriptpartial-derivative𝑁subscript¯𝐹ads𝑉𝑇subscriptsubscript¯𝐹ads𝑁𝑉𝑇\displaystyle\hat{\mu}^{\text{ads}}=\left(\partialderivative{\bar{F}_{\text{% ads}}}{N}\right)_{V,T}=\left(\frac{\bar{F}_{\text{ads}}}{N}\right)_{V,T}.over^ start_ARG italic_μ end_ARG start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT = ( divide start_ARG ∂ start_ARG over¯ start_ARG italic_F end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT end_ARG end_ARG start_ARG ∂ start_ARG italic_N end_ARG end_ARG ) start_POSTSUBSCRIPT italic_V , italic_T end_POSTSUBSCRIPT = ( divide start_ARG over¯ start_ARG italic_F end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT end_ARG start_ARG italic_N end_ARG ) start_POSTSUBSCRIPT italic_V , italic_T end_POSTSUBSCRIPT . (41)
Refer to caption
Figure 3: Types of phase transitions. Potential wells for fluid confined in (a) 11 Å pores and (b) 24 Å pores. Entropy variation is plotted as a function of relative bulk pressure confined in (c) 11 Å pores and (d) 24 Å pores at temperatures ranging from 110 K - 150 K.

Using the formulation in the canonical ensemble, we calculate the grand potential ω𝜔\omegaitalic_ω as follows:

ω=F¯adsNμ^satads.𝜔subscript¯𝐹ads𝑁subscriptsuperscript^𝜇adssat\displaystyle\omega=\bar{F}_{\text{ads}}-N\hat{\mu}^{\text{ads}}_{\text{sat}}.italic_ω = over¯ start_ARG italic_F end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT - italic_N over^ start_ARG italic_μ end_ARG start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT start_POSTSUBSCRIPT sat end_POSTSUBSCRIPT . (42)

where μ^satadssubscriptsuperscript^𝜇adssat\hat{\mu}^{\text{ads}}_{\text{sat}}over^ start_ARG italic_μ end_ARG start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT start_POSTSUBSCRIPT sat end_POSTSUBSCRIPT is the chemical potential at which capillary condensation occurs for a given temperature.

Figures 3(a) and 3(b) show the grand potential thus obtained plotted as a function of density. For the small pore size (11 Å), two phenomena occur. First, the double well vanishes at a much lower temperature than for the large pore size, indicating a lower critical temperature for fluids confined in small pores. Consequently, the entropy variation in Fig. 3(c) shows that a continuous phase transition occurs beyond 130 K. Second, the energy barrier to cross the well is less than the thermal noise (ΔEa0.015kBTΔsubscript𝐸𝑎0.015subscript𝑘B𝑇\Delta E_{a}\approx 0.015{k}_{\text{B}}Troman_Δ italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ≈ 0.015 italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T for 110 K), implying that the system spontaneously jumps between the wells. As a result, a minute step occurs in the entropy variation at 110 K at a lower relative pressure, as shown in Fig. 3(c). This implies that two different phases exist. However,

PA1/PA2=exp(ΔEa/kBT)1,subscript𝑃A1subscript𝑃A2Δsubscript𝐸𝑎subscript𝑘B𝑇1\displaystyle P_{\text{A}1}/P_{\text{A}2}=\exp(\Delta E_{a}/k_{\text{B}}T)% \approx 1,italic_P start_POSTSUBSCRIPT A 1 end_POSTSUBSCRIPT / italic_P start_POSTSUBSCRIPT A 2 end_POSTSUBSCRIPT = roman_exp ( start_ARG roman_Δ italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT / italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T end_ARG ) ≈ 1 , (43)

implying that the probability of the system being in state A1 or A3 (PA1subscript𝑃A1P_{\text{A}1}italic_P start_POSTSUBSCRIPT A 1 end_POSTSUBSCRIPT) approximately equals the probability of the system being in state A2 (PA2subscript𝑃A2P_{\text{A}2}italic_P start_POSTSUBSCRIPT A 2 end_POSTSUBSCRIPT). Therefore, these two phases are practically indistinguishable. In addition, we hypothesize that the absence of hysteresis during the adsorption-desorption loop is because the required activation energy is negligible. This hypothesis is consistent with published data that show that the Type-I adsorption isotherm for H22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT adsorption in IRMOF-1 (generally observed for small pores) has no hysteresis [72].

In contrast, for the large pore size (24 Å), the barrier height is significantly larger (ΔEa15kBTΔsubscript𝐸𝑎15subscript𝑘B𝑇\Delta E_{a}\approx 15k_{\text{B}}Troman_Δ italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT ≈ 15 italic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T), highlighting a clear distinction between the “gas-like adsorbed phase” (i.e., state B1) and the “capillary condensed phase” (i.e., state B3), as shown in Figs. 3(b) and 3(d). However, at 150 K, the double well completely vanishes, and the adsorbed fluid exhibits a single well, which is characteristic of the supercritical bulk fluid, suggesting that the confined fluid has a critical point. Furthermore, based on the activation energy required for capillary condensation, we hypothesize that hysteresis occurs during the adsorption-desorption process for large pores. As seen for the adsorption of H22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT in MOF-253 with a Type-V adsorption isotherm (generally observed for large pores), significant hysteresis occurs in the adsorption-desorption loop [73].

III.2.2 Capillary condensation

Figure 4(a) shows that the phase transition for the adsorbed fluid occurs at a lower relative pressure than for the bulk phase transition. To elucidate this, we compare the nucleation of droplets in a bulk fluid with that in a confined fluid.

As with macroscale condensation, capillary condensation starts with the heterogeneous nucleation of drop clusters that subsequently grow to form droplets [74, 75, 76]. Consistent with classical nucleation theory [77], the nucleation rate depends on the nucleation barrier ΔG*Δsuperscript𝐺\Delta G^{*}roman_Δ italic_G start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT, which is the difference between the interface free energy and the fluid free energy. A lower nucleation barrier corresponds to a higher nucleation rate.

Refer to caption
Figure 4: Energy barrier for capillary condensation. (a) Isotherm of relative number density vs relative pressure for adsorbed argon and the bulk argon. Adsorbed argon condenses at a lower pressure than bulk argon (p/p00.6𝑝subscript𝑝00.6p/p_{\text{0}}\approx 0.6italic_p / italic_p start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ 0.6). Nmaxsubscript𝑁maxN_{\text{max}}italic_N start_POSTSUBSCRIPT max end_POSTSUBSCRIPT is the maximum number of molecules that can be adsorbed. (b) The energy barrier for the phase transition of bulk argon is higher than that for confined argon (ΔEb>ΔEaΔsubscript𝐸𝑏Δsubscript𝐸𝑎\Delta E_{b}>\Delta E_{a}roman_Δ italic_E start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT > roman_Δ italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT).

Figure 4(b) compares the energy barrier ΔEbΔsubscript𝐸𝑏\Delta E_{b}roman_Δ italic_E start_POSTSUBSCRIPT italic_b end_POSTSUBSCRIPT for the bulk fluid with that for the adsorbed fluid (ΔEaΔsubscript𝐸𝑎\Delta E_{a}roman_Δ italic_E start_POSTSUBSCRIPT italic_a end_POSTSUBSCRIPT). The result shows that the free energy for the adsorbed fluid is notably lower than that for the bulk fluid, implying that the free-energy barrier ΔG*Δsuperscript𝐺\Delta G^{*}roman_Δ italic_G start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT for the confined drop nucleation is less than that for nonconfined drop nucleation. Consequently, at a given temperature, the condensation pressure for the adsorbed fluid is less than that for the bulk fluid, as shown in Fig. 4(a).

III.3 Phase diagram of adsorbed fluid

Refer to caption
Figure 5: Phase diagram for argon adsorbed in a metal-organic framework (MOF). The integral properties are plotted in the form of a phase diagram for the adsorbed fluid. Different perspectives of the (a) 3D N-p-h𝑁-𝑝-N\text{-}p\text{-}hitalic_N - italic_p - italic_h diagram are plotted in the insets: (b) p-h𝑝-p\text{-}hitalic_p - italic_h, (c) N-p𝑁-𝑝N\text{-}pitalic_N - italic_p, and (d) N-h𝑁-N\text{-}hitalic_N - italic_h. The red highlighted areas are the phase-coexistence region constructed using the boundaries of discontinuities on either end of each isotherm. Using this coexistence region, a critical point for capillary condensation is depicted with a red dot. The isotherms are plotted every 6 K and one isotherm in each inset is highlighted with green, to show the shape of the isotherm. Each isotherm follows the direction A\rightarrowB\rightarrowC\rightarrowD\rightarrowE\rightarrowF\rightarrowG.

From an application perspective, the phase diagram is an essential tool for designing engineering processes. For example, Lilley and Prasher [78] presented a qualitative phase diagram for crystallization of salts in an ionocaloric refrigeration cycle. In a previous paper [79], we introduced the concept of a 3D phase diagram as a valuable tool for the design of hybrid compression-adsorption heat-pump systems. The statistical model discussed earlier provides a framework for constructing such a phase diagram. Previous research has also aimed to construct phase diagrams for confined fluids. For example, numerous attempts have been made to construct such phase diagrams by using Monte Carlo simulations [80, 81, 37, 82]. However, such simulations are computationally demanding, limiting the number of adsorption isotherms that can be generated. Consequently, acquiring the requisite thermodynamic properties to create a phase diagram for adsorbed fluids is a significant challenge.

Radhakrishnan et al. [37] solved this problem by applying umbrella sampling and bias potentials to compute the system’s free energy. However, such an approach necessitates a priori knowledge of the process, and the convergence of their method depends heavily on the selected collective variables. Lum and Chandler [83] addressed this issue within the framework of statistical mechanics, albeit for a specific scenario, by deriving a phase diagram for vapor confined within a cylindrical pore. Unfortunately, this approach overlooks the nonuniform characteristics inherent in the external field at this length scale, thus lacking generalizability. In contrast, Travalloni et al. [53] constructed a phase diagram employing square-well potentials to account for external heterogeneity. Nevertheless, this particular study does not consider the excess chemical potential induced by external interactions, leading to unrealistically high pressures during capillary condensation.

Figure 5(a) illustrates the 3D phase diagram of adsorbed argon within a 24 Å model of an MOF. The three axes correspond to distinct state variables: the external pressure is denoted pbulksubscript𝑝bulkp_{\text{bulk}}italic_p start_POSTSUBSCRIPT bulk end_POSTSUBSCRIPT, the enthalpy per unit mass of adsorbed argon is denoted hadssubscriptadsh_{\text{ads}}italic_h start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT, and the number of argon molecules adsorbed per unit cell of the model MOF is denoted Nadssubscript𝑁adsN_{\text{ads}}italic_N start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT. Figures 5(b)–5(d) portray the projections of this phase diagram from three distinct orientations, generating the p-h𝑝-p\text{-}hitalic_p - italic_h, N-p𝑁-𝑝N\text{-}pitalic_N - italic_p, and N-h𝑁-N\text{-}hitalic_N - italic_h diagrams. Notably, the isotherms displayed contain discontinuities up to a certain temperature, indicating that two distinct phases coexist. The connection of these discontinuous points yields saturation lines, which are accentuated in red. The highlighted area signifies the coexistence region corresponding to the capillary condensed phase and the gas-like adsorbed phase. In this phase diagram, A\rightarrowB adsorption occurs near the heterogeneity-forming layered structure inside the pore, whereas B\rightarrowC\rightarrowD\rightarrowE depicts the coexistence region. E\rightarrowF shows that the capillary condensed liquid density increases with pressure. Finally, beyond the bulk saturation line F\rightarrowG, the pore is completely filled.

Furthermore, a detailed examination of the p-h𝑝-p\text{-}hitalic_p - italic_h diagram in Fig. 5(b) shows that, at low pressure, the enthalpy of the adsorbed fluid contrasts with that of the bulk fluid. As the pressure decreases, the magnitude of the enthalpy of the adsorbed fluid increases. This phenomenon can be understood by considering the occurrence of layered adsorption near the metallic heterogeneous site. The heightened cohesive interaction with the adsorption site liberates additional energy, increasing the enthalpy at lower pressures. ESI Fig. S8 shows the bulk argon p-h𝑝-p\text{-}hitalic_p - italic_h diagram.

An important observation from the phase diagram is the absence of discontinuities in the isotherms beyond a specific temperature, resembling the behavior of bulk fluids. This temperature is denoted the critical point for capillary condensation. Beyond this critical point, capillary condensation for the adsorbed fluid ceases, resulting in a lack of stepwise behavior in the adsorption isotherm. Notably, the critical point for capillary condensation is positioned at a lower temperature compared with the bulk critical point of argon (151 K, 48.5 bar). This difference is attributed to the excess chemical potential of the adsorbed fluid vis-á-vis the bulk fluid, a consequence of heterogeneous interactions. Similar results of reduction in critical pressure of the liquid-liquid phase transition have been observed for water in a salt solution, where the salt ions act as the heterogeneity [84, 85].

This phase diagram provides a basis to understand the phase transition of confined fluids. Note that a critical order parameter analysis for the “gas-like” adsorbed phase to “capillary-condesed” liquid phase and the analytical construction of the co-existence region still remains to be addressed.

IV Outlook and Conclusions

IV.1 Outlook

This paper proposes a method to predict the integral pressure inside MOF pores (Sec. II). Figure S3(d) of the ESI [67] shows the calculated integral pressure. p^adssubscript^𝑝ads\hat{p}_{\text{ads}}over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT is significantly greater than the bulk pressure p𝑝pitalic_p and jumps discontinuously upon crossing the pore boundary. Therefore, we define a disjoining pressure ΠdsubscriptΠ𝑑\Pi_{d}roman_Π start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT [86, 50] such that

Πdp^adspbulk.subscriptΠ𝑑subscript^𝑝adssubscript𝑝bulk\Pi_{d}\equiv\hat{p}_{\text{ads}}-p_{\text{bulk}}.roman_Π start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ≡ over^ start_ARG italic_p end_ARG start_POSTSUBSCRIPT ads end_POSTSUBSCRIPT - italic_p start_POSTSUBSCRIPT bulk end_POSTSUBSCRIPT . (44)

In a similar way, when considering a constant-pressure ensemble (NPT𝑁𝑃𝑇NPTitalic_N italic_P italic_T), we establish the integral chemical potential such that ξμ^adsN𝜉superscript^𝜇ads𝑁\xi\equiv\hat{\mu}^{\text{ads}}Nitalic_ξ ≡ over^ start_ARG italic_μ end_ARG start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT italic_N. Subsequently, a disjoining chemical potential Mdsubscript𝑀𝑑M_{d}italic_M start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT [87] is defined as

Mdμ^adsμbulk.subscript𝑀𝑑superscript^𝜇adssuperscript𝜇bulkM_{d}\equiv\hat{\mu}^{\text{ads}}-\mu^{\text{bulk}}.italic_M start_POSTSUBSCRIPT italic_d end_POSTSUBSCRIPT ≡ over^ start_ARG italic_μ end_ARG start_POSTSUPERSCRIPT ads end_POSTSUPERSCRIPT - italic_μ start_POSTSUPERSCRIPT bulk end_POSTSUPERSCRIPT . (45)

The proposed formulation can then be extended in other contexts where wall effects and nonuniform external fields are significant, enabling us to calculate these disjoining quantities to explain phenomena other than adsorption.

For example, numerous attempts have been made to explain the stability of surface nanobubbles using the Young–Laplace equation and considering external conditions such as surface charge [88, 89] and contact-line pinning due to hydrophilic and hydrophobic heterogeneities [90, 89]. However, a consensus regarding the exact factors contributing to surface-bubble stability remains elusive. The effect of surface heterogeneity clearly explains the excessive pressure inside these bubbles in terms of a disjoining pressure, offering a plausible explanation for their stability [91].

Similarly, experimental observations reveal that the nucleation temperature for heterogeneous boiling is consistently lower than predicted by classical nucleation theory [92, 93, 94]. Numerous models [95, 96, 97] and complex simulations [98, 36, 99] have been proposed to explain this disparity between experimental and theoretical results. Many such models consider liquid-vapor surface tension through the Young–Laplace equation and liquid-solid surface tension through contact-angle variation. However, to the best of our knowledge, none of these models account for the wall effect, leading to a disjoining pressure between vapor clusters and the liquid, thereby lowering the free-energy barrier for phase transition, as illustrated in Fig. 4(b). The lower free-energy barrier in turn explain the lower the nucleation temperature of heterogeneous boiling [100].

Additionally, in the context of water transport through nanoscale osmotic membranes, a recent study [101] argued that no chemical potential gradient exists for transport within the membrane. However, the nanometer-scale membrane porosity allows surface effects to trigger a disjoining chemical potential, thereby creating a chemical-potential gradient that drives the transport process. Considering how surface effects have a heightened impact on the nanoscale will allow a better understanding of water transport in osmotic membranes.

IV.2 Conclusions

In conclusion, our investigation offers a statistical approach to address the 3D Ising model of phase transitions in confined fluids, producing reasonably accurate results. By applying the proposed model and integrating Hill’s theory of nanothermodynamics, we derive both the differential (local) and integral (global) thermodynamic properties of the adsorbed fluid. The proposed model has practical utility for predicting the behavior of adsorbed fluids within porous structures, facilitating the design of materials tailored to specific requirements. The key insights derived from this model are outlined as follows:

First, the nature of the phase transition in the confined fluid is determined by the extent of confinement, specifically the pore size. In small pores, the activation-energy barrier for phase transition is approximately 0.01 kBTsubscript𝑘B𝑇k_{\text{B}}Titalic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T, significantly less than the thermal noise. Consequently, the phase transition occurs spontaneously. Additionally, the unstable and metastable states, while theoretically existent, are practically indistinguishable from the stable state. Conversely, in the case of large pores, the activation-energy barrier for phase transition is of the order of 10 kBTsubscript𝑘B𝑇k_{\text{B}}Titalic_k start_POSTSUBSCRIPT B end_POSTSUBSCRIPT italic_T, clearly distinguishing between the gas-like adsorbed phase and the capillary condensed phase.

Second, owing to additional interactions with the surface, the free-energy barrier for phase transitions in confined fluids is lower than in bulk fluids. This reduced energy barrier implies that condensation inside MOF pores occurs at a lower pressure for a given temperature, explaining the lower capillary condensation pressure.

Finally, the model proposed in this paper consolidates the integral thermodynamic properties in the form of a phase diagram for confined fluids. The phase diagram resembles the bulk fluid phase diagram except for the higher enthalpy released at lower pressure and the lower critical temperature and pressure due heterogenous interactions.

Acknowledgements.
We acknowledge funding by JST, CREST (Grant No. JPMJCR17I3).

References

  • Taverna et al. [2008] D. Taverna, M. Kociak, O. Stéphan, A. Fabre, E. Finot, B. Décamps, and C. Colliex, Probing physical properties of confined fluids within individual nanobubbles, Phys. Rev. Lett. 100, 035301 (2008).
  • Sircar and Myers [1970] S. E. Sircar and A. L. Myers, Statistical thermodynamics of adsorption from liquid mixtures on solids. i. ideal adsorbed phase, J. Phys. Chem. 74, 2828 (1970).
  • Evans [1990] R. Evans, Fluids adsorbed in narrow pores: phase equilibria and structure, J. Phys. Condens. Matter. 2, 8989 (1990).
  • Morris and Wheatley [2008] R. E. Morris and P. S. Wheatley, Gas storage in nanoporous materials, Angew. Chem. Int. Ed. 47, 4966 (2008).
  • Jew et al. [2022] A. D. Jew, J. L. Druhan, M. Ihme, A. R. Kovscek, I. Battiato, J. P. Kaszuba, J. R. Bargar, and G. E. Brown Jr, Chemical and reactive transport processes associated with hydraulic fracturing of unconventional oil/gas shales, Chem. Rev. 122, 9198 (2022).
  • Krishnamurthy and Cornell [2012] V. Krishnamurthy and B. Cornell, Engineering aspects of biological ion channels—from biosensors to computational models for permeation, Protoplasma 249, 3 (2012).
  • Alabarse et al. [2012] F. Alabarse, J. Haines, O. Cambon, C. Levelut, D. Bourgogne, A. Haidoux, D. Granier, and B. Coasne, Freezing of water confined at the nanoscale, Phys. Rev. Lett. 109, 035701 (2012).
  • Algara-Siller et al. [2015] G. Algara-Siller, O. Lehtinen, F. Wang, R. R. Nair, U. Kaiser, H. Wu, A. K. Geim, and I. V. Grigorieva, Square ice in graphene nanocapillaries, Nature 519, 443 (2015).
  • Gimondi and Salvalaglio [2018] I. Gimondi and M. Salvalaglio, CO2𝐶subscript𝑂2CO_{2}italic_C italic_O start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT packing polymorphism under confinement in cylindrical nanopores, Mol. Syst. Des. Eng. 3, 243 (2018).
  • Kapil et al. [2022] V. Kapil, C. Schran, A. Zen, J. Chen, C. J. Pickard, and A. Michaelides, The first-principles phase diagram of monolayer nanoconfined water, Nature 609, 512 (2022).
  • Hamada et al. [2007] Y. Hamada, K. Koga, and H. Tanaka, Phase equilibria and interfacial tension of fluids confined in narrow pores, J. Chem. Phys. 127, 084908 (2007).
  • Fumagalli et al. [2018] L. Fumagalli, A. Esfandiar, R. Fabregas, S. Hu, P. Ares, A. Janardanan, Q. Yang, B. Radha, T. Taniguchi, K. Watanabe, et al., Anomalously low dielectric constant of confined water, Science 360, 1339 (2018).
  • Kavokine et al. [2022] N. Kavokine, M.-L. Bocquet, and L. Bocquet, Fluctuation-induced quantum friction in nanoscale water flows, Nature 602, 84 (2022).
  • De Gennes [1981] P. G. De Gennes, Polymer solutions near an interface. adsorption and depletion layers, Macromolecules 14, 1637 (1981).
  • Fukui et al. [1999] K. Fukui, H. Onishi, and Y. Iwasawa, Imaging of atomic-scale structure of oxide surfaces and adsorbed molecules by noncontact atomic force microscopy, App. Surf. Sci. 140, 259 (1999).
  • Fukuma et al. [2010] T. Fukuma, Y. Ueda, S. Yoshioka, and H. Asakawa, Atomic-scale distribution of water molecules at the mica-water interface visualized by three-dimensional scanning force microscopy, Phys. Rev. Lett. 104, 016101 (2010).
  • Page et al. [2014] A. J. Page, A. Elbourne, R. Stefanovic, M. A. Addicoat, G. G. Warr, K. Voïtchovsky, and R. Atkin, 3-dimensional atomic scale structure of the ionic liquid–graphite interface elucidated by AM-AFM and quantum chemical simulations, Nanoscale 6, 8100 (2014).
  • Wang and Hadjiconstantinou [2018] G. J. Wang and N. G. Hadjiconstantinou, Layered fluid structure and anomalous diffusion under nanoconfinement, Langmuir 34, 6976 (2018).
  • Gibbs [1928] J. W. Gibbs, The collected works of JW Gibbs (Longmans, Green, 1928).
  • Hill [1962] T. L. Hill, Thermodynamics of small systems, J. Chem. Phys. 36, 3182 (1962).
  • Ono and Kondo [1960] S. Ono and S. Kondo, Molecular theory of surface tension in liquids, in Structure of Liquids / Struktur der Flüssigkeiten (Springer Berlin Heidelberg, Berlin, Heidelberg, 1960) pp. 134–280.
  • Steele and Ross [1960] W. A. Steele and M. Ross, Distribution Functions of a Fluid in an External Potential Field: Application to Physical Adsorption, J. Chem. Phys. 33, 464 (1960).
  • Wertheim [1984a] M. S. Wertheim, Fluids with highly directional attractive forces. I. Statistical thermodynamics, J. Stat. Phys. 35, 19 (1984a).
  • Wertheim [1984b] M. S. Wertheim, Fluids with highly directional attractive forces. II. Thermodynamic perturbation theory and integral equations, J. Stat. Phys. 35, 35 (1984b).
  • Wertheim [1986a] M. Wertheim, Fluids with highly directional attractive forces. III. Multiple attraction sites, J. Stat. Phys. 42, 459 (1986a).
  • Wertheim [1986b] M. Wertheim, Fluids with highly directional attractive forces. IV. Equilibrium polymerization, J. Stat. Phys. 42, 477 (1986b).
  • Biesheuvel et al. [2006] P. M. Biesheuvel, F. A. Leermakers, and M. A. C. Stuart, Self-consistent field theory of protein adsorption in a non-gaussian polyelectrolyte brush, Phys. Rev. E 73, 011802 (2006).
  • Poluektov [2015] Y. M. Poluektov, Thermodynamic perturbation theory for classical systems based on self-consistent field model, Ukr. J. Phys. 60, 553 (2015).
  • Laktionov et al. [2022] M. Y. Laktionov, O. V. Shavykin, F. A. Leermakers, E. B. Zhulina, and O. V. Borisov, Colloidal particles interacting with a polymer brush: A self-consistent field theory, Phys. Chem. Chem. Phys. 24, 8463 (2022).
  • Sircar [1985] S. Sircar, Excess properties and thermodynamics of multicomponent gas adsorption, J. Chem. Soc., Faraday Trans. 1 81, 1527 (1985).
  • Nicholson [1975] D. Nicholson, Molecular theory of adsorption in pore spaces. Part 1.—Isotherms for simple lattice models, J. Chem. Soc., Faraday Trans. 1 71, 238 (1975).
  • Nicholson [1976] D. Nicholson, Molecular theory of adsorption in pore spaces. part 2.—thermodynamic and molecular lattice model descriptions of capillary condensation, J. Chem. Soc., Faraday Trans. 1 72, 29 (1976).
  • Martinez et al. [2007] A. Martinez, M. Castro, C. McCabe, and A. Gil-Villegas, Predicting adsorption isotherms using a two-dimensional statistical associating fluid theory, J. Chem. Phys. 126 (2007).
  • Evans et al. [1986] R. Evans, U. M. B. Marconi, and P. Tarazona, Fluids in narrow pores: Adsorption, capillary condensation, and critical points, J. Chem. Phys. 84, 2376 (1986).
  • Schmidt and Löwen [1997] M. Schmidt and H. Löwen, Phase diagram of hard spheres confined between two parallel plates, Phys. Rev. E 55, 7228 (1997).
  • Kimura and Maruyama [2002] T. Kimura and S. Maruyama, Molecular dynamics simulation of heterogeneous nucleation of a liquid droplet on a solid surface, Micr. Thermo. Eng. 6, 3 (2002).
  • Radhakrishnan et al. [2002] R. Radhakrishnan, K. E. Gubbins, and M. Sliwinska-Bartkowiak, Global phase diagrams for freezing in porous media, J. Chem. Phys. 116, 1147 (2002).
  • Takaiwa et al. [2008] D. Takaiwa, I. Hatano, K. Koga, and H. Tanaka, Phase diagram of water in carbon nanotubes, Proc. Natl. Acad. Sci. 105, 39 (2008).
  • Zhou et al. [2019] Y. Zhou, M. Scheffler, and L. M. Ghiringhelli, Determining surface phase diagrams including anharmonic effects, Phys. Rev. B 100, 174106 (2019).
  • Zhang et al. [2019] Z. Zhang, J. A. Schott, M. Liu, H. Chen, X. Lu, B. G. Sumpter, J. Fu, and S. Dai, Prediction of carbon dioxide adsorption via deep learning, Angew. Chem. Int. Ed. 131, 265 (2019).
  • Gurnani et al. [2021] R. Gurnani, Z. Yu, C. Kim, D. S. Sholl, and R. Ramprasad, Interpretable machine learning-based predictions of methane uptake isotherms in metal–organic frameworks, Chem. Mater. 33, 3543 (2021).
  • Fung et al. [2021] V. Fung, G. Hu, P. Ganesh, and B. G. Sumpter, Machine learned features from density of states for accurate adsorption energy prediction, Nat. Comm. 12, 88 (2021).
  • Cui et al. [2023] J. Cui, F. Wu, W. Zhang, L. Yang, J. Hu, Y. Fang, P. Ye, Q. Zhang, X. Suo, Y. Mo, et al., Direct prediction of gas adsorption via spatial atom interaction learning, Nat. Comm. 14, 7043 (2023).
  • Castelvecchi [2016] D. Castelvecchi, Can we open the black box of AI?, Nat. News 538, 20 (2016).
  • Rudin [2019] C. Rudin, Stop explaining black box machine learning models for high stakes decisions and use interpretable models instead, Nat. Mach. Intell. 1, 206 (2019).
  • Lin et al. [2017] H. W. Lin, M. Tegmark, and D. Rolnick, Why does deep and cheap learning work so well?, J. Stat. Phys. 168, 1223 (2017).
  • Košmrlj and Nelson [2017] A. Košmrlj and D. R. Nelson, Statistical mechanics of thin spherical shells, Phys. Rev. X 7, 011002 (2017).
  • Goodrich et al. [2021] C. P. Goodrich, E. M. King, S. S. Schoenholz, E. D. Cubuk, and M. P. Brenner, Designing self-assembling kinetics with differentiable statistical physics models, Proc. Natl. Acad. Sci. 118, e2024083118 (2021).
  • Molina et al. [2021] A. Molina, S. Kumar, S. Karpitschka, and M. Prakash, Droplet tilings for rapid exploration of spatially constrained many-body systems, Proc. Natl. Acad. Sci. 118, e2020014118 (2021).
  • Israelachvili [2011] J. N. Israelachvili, Intermolecular and surface forces (Academic press, 2011).
  • Hill [2001] T. L. Hill, A different approach to nanothermodynamics, Nano Lett. 1, 273 (2001).
  • Hill [1994] T. L. Hill, Thermodynamics of small systems (Courier Corporation, 1994).
  • Travalloni et al. [2010] L. Travalloni, M. Castier, F. W. Tavares, and S. I. Sandler, Critical behavior of pure confined fluids from an extension of the van der waals equation of state, J. Supercrit. Fluids 55, 455 (2010).
  • Simon et al. [2014] C. M. Simon, J. Kim, L.-C. Lin, R. L. Martin, M. Haranczyk, and B. Smit, Optimizing nanoporous materials for gas storage, Phys. Chem. Chem. Phys. 16, 5499 (2014).
  • Simon et al. [2017] C. M. Simon, E. Braun, C. Carraro, and B. Smit, Statistical mechanical model of gas adsorption in porous crystals with dynamic moieties, Proc. Natl. Acad. Sci. 114, E287 (2017).
  • Singh et al. [2022] N. Singh, F. Simeski, and M. Ihme, Computing thermodynamic properties of fluids augmented by nanoconfinement: Application to pressurized methane, J. Phys. Chem. B 126, 8623 (2022).
  • Dong et al. [2023] W. Dong, T. Franosch, and R. Schilling, Thermodynamics, statistical mechanics and the vanishing pore width limit of confined fluids, Comm. Phys. 6, 161 (2023).
  • Mayer [1937] J. E. Mayer, The statistical mechanics of condensing systems. i, J. Chem. Phys. 5, 67 (1937).
  • Mayer and Montroll [1941] J. E. Mayer and E. Montroll, Molecular distribution, J. Chem. Phys. 9, 2 (1941).
  • Widom [1963] B. Widom, Some topics in the theory of fluids, J. Chem. Phys. 39, 2808 (1963).
  • Hill and Greenschlag [1961] T. L. Hill and S. Greenschlag, Statistical mechanics of monatomic systems in an external periodic potential field. I. Introduction, virial expansion, and classical second virial coefficient, J. Chem. Phys. 34, 1538 (1961).
  • Stanley [1971] H. E. Stanley, Phase transitions and critical phenomena, Vol. 7 (Clarendon Press, Oxford, 1971).
  • Nicolis and Prigogine [1989] G. Nicolis and I. Prigogine, Exploring complexity: An introduction (1989).
  • Kittel and Kroemer [1980] C. Kittel and H. Kroemer, Thermal physics freeman, San Francisco  (1980).
  • You et al. [2022] X. You, Y. Li, H. Mo, and Y. Gui, Theoretical studies on lennard-jones parameters of benzene and polycyclic aromatic hydrocarbons, Faraday Discuss. 238, 103 (2022).
  • Dubbeldam et al. [2016] D. Dubbeldam, S. Calero, D. E. Ellis, and R. Q. Snurr, Raspa: molecular simulation software for adsorption and diffusion in flexible nanoporous materials, Mol. Simul. 42, 81 (2016).
  • [67] Supplementary information, see Supplemental Material at [URL will be inserted by publisher] for further information.
  • Auti et al. [2023] G. Auti, Y. Kametani, H. Kimura, S. Paul, W.-L. Hsu, S. Kusaka, R. Matsuda, T. Uemura, S. Chiashi, and H. Daiguji, Effect of pore size on heat release from CO22{}_{2}start_FLOATSUBSCRIPT 2 end_FLOATSUBSCRIPT adsorption in MIL-101, MOF-177, and UiO-66, J. Mat. Chem. A 11, 20043 (2023).
  • Ball [1993] P. Ball, New horizons in inner space, Nature 361, 297 (1993).
  • Koga et al. [2001] K. Koga, G. Gao, H. Tanaka, and X. C. Zeng, Formation of ordered ice nanotubes inside carbon nanotubes, Nature 412, 802 (2001).
  • Han et al. [2010] S. Han, M. Choi, P. Kumar, and H. E. Stanley, Phase transitions in confined water nanofilms, Nat. Phys. 6, 685 (2010).
  • Rowsell et al. [2004] J. L. Rowsell, A. R. Millward, K. S. Park, and O. M. Yaghi, Hydrogen sorption in functionalized metal- organic frameworks, J. Am. Chem. Soc. 126, 5666 (2004).
  • Yaghi [2012] O. M. Yaghi, Hydrogen storage in metal-organic frameworks, Tech. Rep. (University of California-Los Angeles, Los Angeles, CA, 2012).
  • Bocquet et al. [1998] L. Bocquet, E. Charlaix, S. Ciliberto, and J. Crassous, Moisture-induced ageing in granular media and the kinetics of capillary condensation, Nature 396, 735 (1998).
  • Restagno et al. [2000] F. Restagno, L. Bocquet, and T. Biben, Metastability and nucleation in capillary condensation, Phys. Rev. Lett. 84, 2433 (2000).
  • Maeda and Israelachvili [2002] N. Maeda and J. N. Israelachvili, Nanoscale mechanisms of evaporation, condensation and nucleation in confined geometries, J. Phys. Chem. B 106, 3534 (2002).
  • Debenedetti [1996] P. G. Debenedetti, Metastable liquids: concepts and principles (Princeton university press, 1996).
  • Lilley and Prasher [2022] D. Lilley and R. Prasher, Ionocaloric refrigeration cycle, Science 378, 1344 (2022).
  • Shamim et al. [2022] J. A. Shamim, G. Auti, H. Kimura, S. Fei, W.-L. Hsu, H. Daiguji, and A. Majumdar, Concept of a hybrid compression-adsorption heat pump cycle, Cell Rep. Phys. Sci. , 101131 (2022).
  • Page and Monson [1996] K. Page and P. Monson, Monte carlo calculations of phase diagrams for a fluid confined in a disordered porous material, Phys. Rev. E 54, 6557 (1996).
  • Radhakrishnan et al. [2000] R. Radhakrishnan, K. E. Gubbins, and M. Sliwinska-Bartkowiak, Effect of the fluid-wall interaction on freezing of confined fluids: Toward the development of a global phase diagram, J. Chem. Phys. 112, 11048 (2000).
  • De Grandis et al. [2007] V. De Grandis, P. Gallo, and M. Rovere, The phase diagram of confined fluids, J. Mol. Liq. 134, 90 (2007).
  • Lum and Chandler [1998] K. Lum and D. Chandler, Phase diagram and free energies of vapor films and tubes for a confined fluid, Int. J. Thermophys. 19, 845 (1998).
  • Gallo et al. [2016] P. Gallo, K. Amann-Winkel, C. A. Angell, M. A. Anisimov, F. Caupin, C. Chakravarty, E. Lascaris, T. Loerting, A. Z. Panagiotopoulos, J. Russo, et al., Water: A tale of two liquids, Chem. Rev. 116, 7463 (2016).
  • Perin and Gallo [2023] L. Perin and P. Gallo, Phase diagram of aqueous solutions of LiCl: A study of concentration effects on the anomalies of water, J. Phys. Chem. B 127, 4613 (2023).
  • Derjaguin and Kusakov [1936] B. Derjaguin and M. Kusakov, The properties of thin layers of liquids, Proc. Acad. Sci. USSR Chem. Series 5, 741 (1936).
  • Dong [2023] W. Dong, Nanoscale thermodynamics needs the concept of a disjoining chemical potential, Nat. Comm. 14, 1824 (2023).
  • Shi et al. [2021] X. Shi, S. Xue, T. Marhaba, and W. Zhang, Probing internal pressures and long-term stability of nanobubbles in water, Langmuir 37, 2514 (2021).
  • Tan et al. [2021] B. H. Tan, H. An, and C.-D. Ohl, Stability of surface and bulk nanobubbles, Curr. Opin. Colloid Interface Sci 53, 101428 (2021).
  • Maheshwari et al. [2016] S. Maheshwari, M. van der Hoef, X. Zhang, and D. Lohse, Stability of surface nanobubbles: A molecular dynamics study, Langmuir 32, 11116 (2016).
  • Svetovoy et al. [2016] V. B. Svetovoy, I. Devic, J. H. Snoeijer, and D. Lohse, Effect of disjoining pressure on surface nanobubbles, Langmuir 32, 11188 (2016).
  • Witharana et al. [2012] S. Witharana, B. Phillips, S. Strobel, H. Kim, T. McKrell, J.-B. Chang, J. Buongiorno, K. Berggren, L. Chen, and Y. Ding, Bubble nucleation on nano-to micro-size cavities and posts: An experimental validation of classical theory, J. App. Phys. 112 (2012).
  • Chen et al. [2019] J. Chen, K. Zhou, Y. Wang, J. Gao, T. Yuan, J. Pang, S. Tang, H.-Y. Chen, and W. Wang, Measuring the activation energy barrier for the nucleation of single nanosized vapor bubbles, Proc. Natl. Acad. Sci. 116, 12678 (2019).
  • Paul et al. [2020] S. Paul, W.-L. Hsu, M. Magnini, L. Mason, Y. Ho, O. K. Matar, and H. Daiguji, Single-bubble dynamics in nanopores: Transition between homogeneous and heterogeneous nucleation, Phys. Rev. Res. 2, 043400 (2020).
  • Lutsko [2018] J. F. Lutsko, Systematically extending classical nucleation theory, New J. Phys. 20, 103015 (2018).
  • Abyzov and Schmelzer [2014] A. S. Abyzov and J. W. Schmelzer, Heterogeneous nucleation in solutions: Generalized gibbs’ approach, J. Chem. Phys. 140 (2014).
  • Magaletti et al. [2020] F. Magaletti, A. Georgoulas, and M. Marengo, Unraveling low nucleation temperatures in pool boiling through fluctuating hydrodynamics simulations, Int. J. Multiph. Flow 130, 103356 (2020).
  • Shen and Debenedetti [2001] V. K. Shen and P. G. Debenedetti, Density-functional study of homogeneous bubble nucleation in the stretched Lennard-Jones fluid, J. Chem. Phys. 114, 4149 (2001).
  • Lutsko [2011] J. F. Lutsko, Density functional theory of inhomogeneous liquids. iv. squared-gradient approximation and classical nucleation theory, J. Chem. Phys. 134 (2011).
  • Majumdar and Mezic [1999] A. Majumdar and I. Mezic, Instability of ultra-thin water films and the mechanism of droplet formation on hydrophilic surfaces, J. Heat Transfer 121, 964 (1999).
  • Song et al. [2021] L. Song, M. Heiranian, and M. Elimelech, True driving force and characteristics of water transport in osmotic membranes, Desalination 520, 115360 (2021).