Elsevier

Chemical Engineering Journal

Volume 239, 1 March 2014, Pages 274-283
Chemical Engineering Journal

Effects of temperature, pore dimensions and adsorbate on the transition from pore blocking to cavitation in an ink-bottle pore

https://doi.org/10.1016/j.cej.2013.11.019Get rights and content

Highlights

  • Evolution of hysteresis loop with temperature, pore dimensions and adsorbate.

  • Fused hysteresis loop of IUPAC and de Boer types.

  • Double maxima in excess isotherms under supercritical conditions.

Abstract

We have carried out comprehensive GCMC molecular simulations to study the effects of temperature (ranging from sub-critical to supercritical), cavity and neck dimensions, and adsorbate on the transition from pore blocking to cavitation in slit shaped ink-bottle pores. Varying these parameters affects, not only the position and the size of the hysteresis loop, but also its shape, which can change from H1, typical for pore blocking, to H1 or H2 combined with Type C (in the de Boer classification). The combined loops can either be fused loops or appear as two separate loops, one of which is of Type H1 and the other Type C. Another highlight of our simulation study is the double maxima in the adsorption isotherm at a supercritical temperature which results from the sigmoidal shape in the plot of bulk gas density versus pressure and the compression of the adsorbate in the confined space at high pressures.

Introduction

Hysteresis associated with capillary condensation and evaporation in porous materials has been the subject of immense interest for over 100 years [1], especially the search for the controlling mechanisms of adsorption and desorption. Adsorption in mesopores gives rise to hysteresis when the temperature is less than the critical hysteresis temperature and pore size is greater than a critical value [2], [3].

Materials such as activated carbon, porous glass and silica gel consist of interconnected networks of pores of various shape and size, and their experimental isotherms can exhibit single or double steps in the hysteresis loop [2], [4], [5], [6], [7], [8], [9], [10], [11], [12], [13], [14], [15], [16], [17], [18]. When hysteresis shows two distinct steps, the first, at lower pressure, is associated with condensation and evaporation in the smaller pores and the second with the same processes in wider pores. If the adsorbate–adsorbent system is wetting, adsorption proceeds by molecular layering, followed by condensation when both ends of a pore are exposed to the bulk surroundings, or by the advance of a meniscus from the closed end if one end of the pore is closed. Desorption, on the other hand, takes place by two processes which occur in conjunction: (1) the withdrawal of menisci from the pore mouth, and (2) the stretching of the condensed fluid in the interior region to a pressure where bubbles (cavities) appear in the adsorbate. When the first process dominates, the desorption mechanism is described as pore blocking, and as cavitation if stretching of the condensed fluid reaches the stability limit before the menisci have travelled to the pore interior. Both modes of evaporation can be illustrated by simulations using a simple ink-bottle pore model by tuning the neck size [19], [20], [21], [22], [23]. For a given adsorbate–adsorbent pair and temperature, the mechanism of desorption changes from pore-blocking to cavitation as the neck size decreases [20], [21], [22], [23], [24]. The cavitation pressure is a fluid property only when the cavity is large, typically greater than 7 nm for argon adsorption at 87 K, but is dependent on the cavity size for smaller cavities, because of the the overlap of adsorbent potential from closely spaced pore walls creates a stabilization effect [25], [26]. The neck length can affect evaporation in an interesting way: Even when the neck size is smaller than the value at which cavitation normally occurs, evaporation can switch to pore blocking when the neck is very short; the shorter the neck, the greater the desorption pressure. While the cavity size affects the cavitation pressure for small cavities and the neck dimensions (width and length) affect the governing mechanism for desorption, temperature can also affect the desorption mechanism, which changes from pore blocking to cavitation at high temperature, because stretching of the condensed fluid in the cavity, overrides the process of meniscus withdrawal. This has been observed both experimentally and theoretically [4], [7], [11]. The change of evaporation mechanism for a given adsorbent, can also switch from pore blocking to cavitation as the pressure is reduced, and this change depends strongly on the pore structure and temperature [11], [18], [27]. Finally, the adsorbate molecule can also affect the desorption mechanism; for example Reichenbach and co-worker [11] observed pore blocking for argon adsorption in porous glass at 77 K, but found cavitation to be the mechanism for nitrogen at the same temperature.

Despite numerous simulation studies, there is still no systematic investigation into the effects of pore dimensions, temperature and adsorbate on the switch in the mechanism of desorption, from pore blocking to cavitation, in ink-bottle pores. It is the objective of this paper to fill this gap.

Section snippets

Fluid–fluid potential model

Argon was modelled as a single Lennard Jones (LJ) site and the 2CLJ + 3q N2 (two LJ sites and three partial charges) model was chosen for nitrogen [28]. Their molecular parameters are listed in Table 1.

Fluid–solid potential model

A graphitic ink-bottle pore, with planar walls, connected to a gas reservoir via a neck is shown in Fig. 1. We used the Bojan-Steele equation [29], [30], [31] to calculate the fluid–solid potential energy with a 0.34 nm spacing between the two adjacent graphite layers, which are finite in the y

Ar adsorption at less than the critical hysteresis temperature

We first investigate argon adsorption in an ink-bottle pore whose cavity width is 3 nm and neck width is 2.3 nm. Note that in this small cavity and the force per unit area exerted by the cavitation pressure is less than the tensile strength of the bulk fluid because of the stabilization of the condensed fluid by the external field from the adsorbent. The lengths of the cavity and the neck were 10 nm. Fig. 2 shows the isotherms for temperatures from 60 K to 150 K. For ease of discussion, we define in

Conclusions

We have used grand canonical Monte Carlo simulation to investigate the transition from pore blocking to cavitation in the desorption branch of argon and nitrogen isotherms in carbonaceous ink-bottle pores. The transition temperature is found to increase with (i) increasing width of the pore neck, (ii) a decrease in neck length for very short necks (for longer necks, the transition temperature is insensitive to the length), (iii) a stronger holding potential for the adsorbate.

Argon adsorption in

Acknowledgement

This project was supported by the Australian Research Council.

References (36)

  • M. Kruk et al.

    Argon adsorption at 77 K as a useful tool for the elucidation of pore connectivity in ordered materials with large cagelike mesopores

    Chemistry of Materials

    (2003)
  • S.P. Rigby et al.

    Experimental evidence for pore blocking as the mechanism for nitrogen sorption hysteresis in a mesoporous material

    The Journal of Physical Chemistry B

    (2004)
  • K. Morishige et al.

    Change in desorption mechanism from pore blocking to cavitation with temperature for nitrogen in ordered silica with cagelike pores

    Langmuir

    (2006)
  • A. Grosman et al.

    Capillary condensation in porous materials. Hysteresis and interaction mechanism without pore blocking/percolation process

    Langmuir

    (2008)
  • K. Morishige et al.

    Neck size of ordered cage-type mesoporous silica FDU-12 and origin of gradual desorption

    The Journal of Physical Chemistry C

    (2010)
  • K. Morishige et al.

    Large-pore cagelike silica with necks of molecular dimensions

    The Journal of Physical Chemistry C

    (2010)
  • C. Reichenbach et al.

    Cavitation and pore blocking in nanoporous glasses

    Langmuir

    (2011)
  • C.J. Rasmussen et al.

    Cavitation in metastable liquid nitrogen confined to nanoscale pores

    Langmuir

    (2010)
  • Cited by (20)

    • Preparation of uniform and highly dispersed magnetic copper ferrite sub-micron sized particles regulated by short-chain surfactant with catechol structure: Dual-functional materials for supercapacitor and dye degradation

      2020, Journal of Electroanalytical Chemistry
      Citation Excerpt :

      The pore size distribution obtained from the adsorption curve and nitrogen isotherm calculated by the BJH method showed that the relatively narrow pore size distribution of CF-1 is concentrated at about 14.4 nm, the pore volume is 0.04 cm2·g−1. The pore size distribution of CF-2 and CF/F were concentrated around 9.1 nm and 10.5 nm and the pore volume is 0.05 and 0.03 cm2·g−1, respectively [45,46]. The result indicated that the amount of short-chain surfactant (DA) had a greater impact on the specific surface area and pore structure of the material.

    • Effects of nitrogen and oxygen functional groups and pore width of activated carbon on carbon dioxide capture: Temperature dependence

      2020, Chemical Engineering Journal
      Citation Excerpt :

      The hysteresis loop of porous carbon is typically closed at relative pressure, P/P0 ~ 0.4, which is the cavitation pressure of N2 at −196 °C. This pressure is low enough to break the N2-N2 stretching in the cavity pore [54]. Pore size distributions obtained from N2 adsorption at −196 °C in Fig. 2b point out that the micropore volume of all prepared samples fluctuated in the range of 0.7 to 1.5 nm.

    • Monitoring the structural and textural changes of Ni-Zn-Al hydrotalcites under heating

      2020, Thermochimica Acta
      Citation Excerpt :

      Additionally, the N0.50Z0.25A0.25 sample showed type H1 hysteresis, which is typical of materials with porous formed by agglomerates of uniform spheres with a regular arrangement that originate an average pore diameter of 4.45 nm in this solid. For N0.75A0.25 (4.45 nm; pore diameter) and N0.66A0.33 (4.17 nm; pore diameter) samples, the H2 type hysteresis loops were noted, that are typical of systems of interconnected pores or of ink-bottle type pores [46,47]. These materials additionally presented a narrower and more symmetrical distribution of pores than that recorded by zinc-containing materials, as shown in Fig. 11.

    • A general strategy to fabricate soft magnetic CuFe<inf>2</inf>O<inf>4</inf>@SiO<inf>2</inf> nanofibrous membranes as efficient and recyclable Fenton-like catalysts

      2019, Journal of Colloid and Interface Science
      Citation Excerpt :

      Consequently, the porous structure of relevant CuFe2O4@SNM was further detected through the surface area analyzer. As the N2 adsorption-desorption curves of the relevant samples shown in Fig. 4a, the N2 adsorption capacity increased slowly with the increasing relative pressure, and subsequently dramatically increased to the maximum when P/P0 reached 0.9, exhibiting a type IV isotherm on the basis of classification methods of International Union of Pure and Applied Chemistry (IUPAC) [29,30]. Furthermore, a narrow and reproducible H1 hydrolysis loop located in the P/P0 range of 0.4–0.9 could be obviously observed, representing the filling and emptying of open mesopores through capillary evaporation and N2 condensation [31].

    • Hysteresis and scanning curves in linear arrays of mesopores with two cavities and three necks. Classification of the scanning curves

      2016, Colloids and Surfaces A: Physicochemical and Engineering Aspects
      Citation Excerpt :

      These scanning curves are observed when some parts of the adsorbent are filled with adsorbate (Region I), while the other regions (Region II) are still empty, except for molecular layers on the pore walls. The descending scanning curve is associated with stretching of the condensate in Region I, followed by evaporation, and the ascending scanning curve is associated with further layering of molecules on the pore walls in Region II, followed by condensation [15–17]. Recognizing the potential gain of a deeper insight into pore structure, and the growing power of modern adsorption equipment, scanning has been receiving increasing attention in both experimental and theoretical studies [18–20].

    View all citing articles on Scopus
    View full text