Next Article in Journal
Triboelectric Nanogenerators for Energy Harvesting in Ocean: A Review on Application and Hybridization
Next Article in Special Issue
CO2 Capture by Virgin Ivy Plants Growing Up on the External Covers of Houses as a Rapid Complementary Route to Achieve Global GHG Reduction Targets
Previous Article in Journal
Comparative Analysis of Performance, Emission, and Combustion Characteristics of a Common Rail Direct Injection Diesel Engine Powered with Three Different Biodiesel Blends
Previous Article in Special Issue
Current Developments of Carbon Capture Storage and/or Utilization–Looking for Net-Zero Emissions Defined in the Paris Agreement
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Adsorption of Carbon Dioxide, Methane, and Nitrogen on Zn(dcpa) Metal-Organic Framework

by
Rui P. P. L. Ribeiro
*,
Isabel A. A. C. Esteves
and
José P. B. Mota
*
LAQV-REQUIMTE, Department of Chemistry, NOVA School of Science and Technology, NOVA University of Lisbon, 2829-516 Caparica, Portugal
*
Authors to whom correspondence should be addressed.
Energies 2021, 14(18), 5598; https://doi.org/10.3390/en14185598
Submission received: 11 August 2021 / Revised: 31 August 2021 / Accepted: 3 September 2021 / Published: 7 September 2021
(This article belongs to the Special Issue CO2 Capture and Renewable Energy)

Abstract

:
Adsorption-based processes using metal-organic frameworks (MOFs) are a promising option for carbon dioxide (CO2) capture from flue gases and biogas upgrading to biomethane. Here, the adsorption of CO2, methane (CH4), and nitrogen (N2) on Zn(dcpa) MOF (dcpa (2,6-dichlorophenylacetate)) is reported. The characterization of the MOF by powder X-ray diffraction (PXRD), thermogravimetric analysis (TGA), and N2 physisorption at 77 K shows that it is stable up to 650 K, and confirms previous observations suggesting framework flexibility upon exposure to guest molecules. The adsorption equilibrium isotherms of the pure components (CO2, CH4, and N2), measured at 273–323 K, and up to 35 bar, are Langmuirian, except for that of CO2 at 273 K, which exhibits a stepwise shape with hysteresis. The latter is accurately interpreted in terms of the osmotic thermodynamic theory, with further refinement by assuming that the free energy difference between the two metastable structures of Zn(dcpa) is a normally distributed variable due to the existence of different crystal sizes and defects in a real sample. The ideal selectivities of the equimolar mixtures of CO2/N2 and CO2/CH4 at 1 bar and 303 K are 12.8 and 2.9, respectively, which are large enough for Zn(dcpa) to be usable in pressure swing adsorption.

1. Introduction

Metal-organic frameworks (MOFs) are being touted as the next generation materials for several adsorptive separation and purification processes [1,2]. MOFs are porous crystalline materials consisting of metal centers connected by organic moieties [3]. An unlimited amount of MOF structures can be envisioned and perhaps synthesized; furthermore, the materials can be tailored for specific applications through pore size tuning and functionalization [4]. These features place MOFs as a very diverse class of materials with potential applications in nearly all fields of chemical engineering [5,6,7,8,9,10,11,12].
Among the available portfolio of MOFs, there are several structures that present structural flexibility, which can be triggered by exposure to specific guest species, changes in temperature or mechanical pressure, or interactions with light or electric fields [13,14]. Framework flexibility generally manifest itself through breathing or gate-opening effects [13]. A comprehensive review regarding MOF flexibility was published by Schneemann et al. [14], in which it is stated that, so far, less than a hundred MOFs have shown important breathing effects. The authors classified the type of flexibility into “breathing”, “swelling”, “linker rotation”, and “subnetwork displacement”. The most well-known cases of MOF flexibility are the breathing behavior of MIL-53 [15,16,17] and the gate-opening of ZIF-8 [18,19].
The MIL-53 family of MOFs consists of trivalent metal (e.g., Al [20], Cr [17], Fe [21], and Sc [22]) terephthalates, which can switch between large-pore (lp) and narrow-pore (np) forms [15,16,17], with unit cell volume variation of up to 40% [23]. The conformational change can be triggered by different stimuli, for example temperature changes [22,24], application of mechanical pressure [25,26], and adsorption of guest molecules (such as H2O, CO2, and other gases) [20,22,26,27]. Interestingly, the synthesis route and solvents employed critically impact the breathing properties of MIL-53 [28], and several authors have observed the absence of the breathing effect in the commercial MIL-53(Al) synthesized by BASF (Basolite©A100) when exposed to CO2 [29,30,31].
Another type of MOF flexibility is related to linker rotation [32], of which the most well-known example occurs in ZIF-8, which presents a gate-opening effect [18,19]. The linker rotation triggers a window opening that allows for the adsorption of larger molecules than expected [18].
Recently, MOFs with step-shaped isotherms typical of flexible MOFs have been considered as potential adsorbents for CO2 capture by temperature swing adsorption (TSA), as they permit decreasing the energy consumption of the process when compared with traditional zeolite 13X systems [33].
Zn(dcpa) is a poorly studied microporous MOF that reportedly exhibits dynamic behavior and stepwise adsorption. Zn(dcpa) is based on paddle-wheel Zn2 units and unsymmetrical pyridyl dicarboxylate, which give rise to a three-dimensional intersecting pore network with a pore opening of 6.3 × 12.2 Å2 [34]. The Zn(dcpa) framework structure is shown in Figure 1. Liu et al. [34] observed the dynamic behavior of Zn(dcpa) upon exposure to N2 and CO2 at 77 K and 195 K, respectively. However, they did not observe the MOF’s flexible behavior when adsorbing CO2 at 273 and 293 K and N2 at 293 K, up to 1 bar.
In this work, the potential of Zn(dcpa) for application in the separation/purification of gaseous streams containing CO2, CH4, and N2, namely post-combustion CO2 capture and biogas upgrading, is evaluated. For this purpose, the single-component adsorption equilibria of CO2, CH4, and N2 have been measured at 273−323 K up to 35 bar, and the isosteric heat of adsorption and ideal CO2/CH4 and CO2/N2 equilibrium selectivities evaluated. Furthermore, the MOF has been characterized regarding its textural properties and thermal stability. The uncommon stepwise adsorption and hysteretic desorption behavior for CO2 at 273 K has been interpreted in terms of the osmotic thermodynamic theory. The data reported here add important knowledge about the adsorption properties of Zn(dcpa), as prior studies about this MOF are limited to a few publications [34,35].

2. Materials and Methods

2.1. Materials

The Zn(dcpa) MOF sample employed was synthesized at the Materials Center at Technical University Dresden (Germany). After its synthesis, the sample was washed with DMF and, subsequently, activated at 453 K under vacuum for 24 h. The gases employed in the measurements were provided by Air Liquide and Praxair (Portugal) with purities of 99.998% (CO2), 99.95% (CH4), 99.99% (N2), and 99.999% (He).

2.2. Zn(dcpa) Characterization

The sample was characterized using powder X-ray diffraction (PXRD), thermogravimetric analysis (TGA), N2 physisorption at 77 K, and helium porosimetry. The N2 adsorption isotherm at 77 K and PXRD were determined by the supplier upon request. TGA analysis was performed using a LABSYS Evo TGA-DTA/DSC from SETARAM Instrumentation, under an argon flow at a heating rate of 3 K/min (up to 1130 K). Helium picnometry was performed at 323 K, in a gravimetric apparatus (described in the next section), to determine the skeletal density of the MOF (ρs).

2.3. Single-Component Adsorption Equilibrium

Single-component adsorption equilibrium isotherms of CO2, CH4, and N2 at 273 K, 303 K, and 323 K, between 0 and 35 bar, were determined using the standard static gravimetric method [31,36,37]. The measurements were performed in a high-pressure magnetic-suspension balance ISOSORP 2000 (Rubotherm GmbH, Germany) using approximately 600 mg of Zn(dcpa) powder. Both the adsorption and desorption data were recorded to evaluate the hysteretic effects. The sample was received from the supplier already activated and stored in an argon atmosphere, which is why the pre-treatment performed before measuring the adsorption equilibrium isotherms was limited to overnight vacuum. The experimental setup and procedure are detailed elsewhere [31,37].
The excess amount adsorbed, q exc , is determined as follows
q exc = w m s m h + V h ρ g m s + v s ρ g  
where w is the apparent mass weighted; m s is the mass of MOF; V h and m h correspond to the volume and mass of the measuring cell, respectively, which contribute to the buoyancy effects; ρ g is the density of the bulk gas at the experimental conditions; and v s is the specific volume of the solid matrix of the MOF ( v s = 1 / ρ s , where ρ s is the skeletal density of the adsorbent). v s was determined by helium pycnometry performed at 323 K in the gravimetric apparatus. This was determined assuming that He penetrates the MOF pore volume without being adsorbed.
The absolute amount adsorbed, q , can be determined from the excess amount adsorbed, using the following
q   = q exc ( ρ l ρ l ρ g )
assuming that the adsorbed phase density corresponds to the density of the liquid at its boiling point at 1 atm ( ρ l ) [38].

3. Results and Discussion

3.1. Zn(dcpa) Characterization

The PXRD patterns obtained are displayed in Figure 2, showing that the position of the reflexes changed during activation. This effect was also observed by Liu et al. [34], who associated it with framework shrinkage upon removal of the guest molecules. The authors also observed the reversibility of this phenomenon when readsorbing the solvent.
The thermal stability of the sample was also characterized by TGA; the recorded sample mass as a function of the heating temperature is shown in Figure 3. The results show an initial mass decrease (~10%) from room temperature up to around 400 K, due to the removal of pre-adsorbed impurities and humidity, due to MOF exposure to the indoor atmosphere just prior to the analysis. MOF was stable up to 650 K, after which a steep decrease in the mass was observed, reaching a mass decrease of 50%, similar to the behavior observed by Liu et al. [34]. Above 750 K, the Zn(dcpa) mass decreased more smoothly until reaching the remaining experimental amount of ca. 25% at 1130 K.
The Zn(dcpa) porosity was evaluated by N2 adsorption at 77 K. The obtained isotherm is plotted in Figure 4, showing an initial step followed by a smoother increase until p / p 0 = 0.06 ; then, another steep increase is observed before reaching a nearly constant plateau with only a small increase between p / p 0 = 0.2     (274 cm3/g) and p / p 0 = 0.97     (303 cm3/g). Note that p and p 0 are the equilibrium and saturation pressures of the adsorbate at 77 K, respectively. The same behavior was observed by Liu et al. [34] who attributed the first step of the isotherm to the Zn(dcpa) structure with shrunken pores, and the second step to an expanded structure. In our work, the expanded structure had a specific pore volume of 0.47 cm3/g, determined at a relative pressure of p / p 0 = 0.97 , assuming the pores were filled with condensed liquid N2 at its normal boiling point. The desorption branch showed hysteresis at p / p 0 < 0.2 , which is also in accordance with a previous report, although in our case, the hysteresis loop seemed to close at lower pressures—which corresponds to a return to the shrunken pore conformation—as opposed to the observation of Liu et al. [34].

3.2. Single-Component Adsorption Equilibrium

Prior to the adsorption of CO2, CH4, and N2, the skeletal density of Zn(dcpa) was determined by helium picnometry at 323 K, obtaining a ρ s = 1.74   g / cm 3 ( v s = 1 / ρ s = 0.575 cm3/g). For a purely crystalline porous material with a regular lattice, v μ + v s is equal to the specific volume of the unit cell of the lattice. The particle density determined, ρ p = 1 / ( v s + v μ ) = 0.957 g/cm3, is in excellent agreement with the value obtained from the crystallographic data ( ρ p = 0.961 g/cm3) by Liu et al. [34].
The adsorption equilibria of CO2, CH4, and N2 on the Zn(dcpa) MOF were measured at 273, 303, and 323 K over the pressure range of 0 to 35 bar. The CO2, CH4, and N2 absolute adsorption equilibrium isotherms obtained are reported in Figure 5, Figure 6 and Figure 7, respectively. The CO2 adsorption isotherms were quite steep in the Henry region, showing a high adsorption capacity at a low pressure, an important feature for use in post-combustion CO2 capture applications. On the other hand, the N2 adsorption isotherms were much more linear and had lower adsorption capacity; the CH4 adsorption isotherms were intermediate between those of CO2 and N2.
An interesting feature of the CO2 adsorption equilibrium isotherms can be observed in Figure 5. At 273 K, the adsorption branch of the isotherm follows a typical Langmuir-type shape up to approximately 22 bar, where a step in the isotherm is observed. This behaviour is similar to that observed by Liu et al. [34] for CO2 adsorption at 195 K, which the authors related to the transition between a shrunken-pore phase and an expanded-pore one. The desorption branch then follows a different path than the adsorption one, showing a hysteresis loop that closes at 7 bar. The reproducibility of this behaviour was checked by repeating the measurements. The stepwise CO2 adsorption observed is an interesting feature of Zn(dcpa) that can be explored for gas separation or storage applications. It can enhance the working capacity of the solid material upon mild pressure or temperature swings [33]. Despite the observation of the MOF flexibility for CO2 adsorption at 273 K, the same behaviour was not observed for any of the other temperatures nor for the adsorbate species tested (CH4 and N2).

3.3. Osmotic Thermodynamic Theory

Some materials present clear transitions between different metastable framework structures. Zn(dcpa) is an example of such materials. In these cases, an “osmotic subensemble” [15,39,40,41,42] can be employed to describe the equilibrium between host structures when exposed to gaseous adsorbates. Alternatively, Ghysels et al. [42] proposed another free energy model able to describe the thermodynamics of breathing phenomena in flexible materials. In this work, we interpreted our data using the former approach.
The osmotic potential [43] of the solid−adsorbate system for the ith structure of Zn(dcpa), either shrunken-pore (SP) or expanded-pore (EP), is
Ω os ( i ) ( P , T ) = F host ( i ) ( T ) + P v p ( i ) 0 P q ( i ) ( P , T ) v g ( P , T )   d P = F host ( i ) ( T ) + P v p ( i ) R T 0 P q ( i ) ( P , T ) Z ( P , T )   d ln P
where F host ( i ) ( T ) corresponds to the empty structure’s free energy at temperature T and v p ( i ) to its apparent specific volume (i.e., to the sum of its skeletal, v s , and porous, v μ ( i ) , volumes); q ( i ) ( P , T ) corresponds to the adsorption isotherm considering a rigid framework in its ith structural form; and v g = 1 / n g and P are the molar volume and compressibility factor of the adsorptive. The difference in the osmotic potential between both of the structures considered (EP and SP), Ω os ( P , T ) =   Ω os ( EP ) ( P , T ) Ω os ( SP ) ( P , T ) , is thus
Ω os ( P , T ) = F host ( T ) + P v p R T 0 P q ( P , T ) Z ( P , T )   d ln P
where ϕ ϕ ( EP ) ϕ ( SP ) is the difference in the value of property ϕ between the EP and SP structures at temperature T . If Ω os > 0 , the SP structure will be more stable than EP; if Ω os < 0 , the reverse will be true. For Zn(dcpa), v μ ( EP ) = 0.47 cm3/g and v μ ( SP ) = 0.14 cm3/g; hence v p = ( v s + v μ ) ( EP ) ( v s + v μ ) ( SP ) v μ ( EP ) v μ ( SP ) = 0.33 cm3/g. Assuming ideal gas behavior ( Z 1 ), the previous equation can be simplified to
Ω os ( P , T ) F host ( T ) + P v p R T 0 P q ( P , T )   d ln P .
If F host and q ( P ) are known at a given temperature T , putting Ω os = 0 in Equation (4) (or Equation (5)) and solving it for P gives the pressure at which the phase transition occurs at T . However, in real scenarios, the MOF crystals have defects and the sample has a distribution of crystal sizes, both contributing to smoothing the structural transitions between the two metastable framework structures. Here, we extend the osmotic thermodynamic theory to account for this diffuse effect. It is assumed that for a real sample, F host is normally (Gaussian) distributed around the corresponding value for a perfect crystal with a probability density function
f ( F host ) = 1 σ F 2 π exp ( 1 2 ( F host F μ ) 2 σ F 2 ) ,
where F μ is the mean or expectation of the distribution (i.e., the value of F host for a perfect crystal) and σ F is its standard deviation. Given that in the case under study, the structural transition is triggered by exposure to a specific guest species, the previous hypothesis is almost equivalent to considering that the adsorptive pressure, P , that triggers the structural transition at a fixed temperature is also a normally distribute variable and, therefore, that its cumulative distribution function, Φ ( P ) , at a fixed temperature is
Φ ( P ) f ( P ) d P = 1 2 [ 1 + erf ( P P μ σ P 2 ) ]
where P μ is the mean or expectation of the distribution (i.e., the transition pressure for a perfect crystal) and σ P is its standard deviation. Note that Φ ( x ) = prob ( P     x ) , where the right-hand side represents the probability that P takes on a value less than or equal to x . Therefore, the macroscopically observed adsorption branch of the isotherm for a real sample is given by
q ads ( P ) = [ 1 Φ ads ( P ) ] q ( SP ) ( P ) + Φ ads ( P ) q ( EP ) ( P ) q des ( P ) = [ 1 Φ des ( P ) ] q ( SP ) ( P ) + Φ des ( P ) q ( EP ) ( P ) ( fixed   T )
where q ( i ) ( P , T ) is the adsorption isotherm considering a rigid framework constrained to its ith form.
This model was fitted to our experimental data, assuming the adsorption isotherms for the metastable forms of the framework are Langmuirian, i.e.,
q ( i ) ( P ) = q ( i ) b ( i ) P 1 + b ( i ) P
where q ( i ) and b ( i ) are the saturation capacity and equilibrium constant for the ith form, respectively, in which case Equation (5) reduces to
Ω os ( P , T ) = F host ( T ) + P v p R T [ q ( EP ) ln ( 1 + b ( EP ) P ) q ( SP ) ln ( 1 + b ( SP ) P ) ]
Table 1 lists the parameter values resulting from the model fitting to the experimental data, saturation capacity and Langmuir equilibrium constant ( q ( i ) and b ( i ) ) for CO2 adsorption at 273 K in the SP and EP metastable structures of Zn(dcpa), mean and standard deviation ( P μ and σ μ ,) of the Gaussian distribution of adsorptive pressure that triggers the phase transition along the adsorption and desorption branches of the isotherm, and the corresponding free energy changes of the empty structure ( F host ).
Figure 8 compares the experimental adsorption data and the fitted osmotic thermodynamic model, and shows that excellent agreement with the experimental results has been reached using the proposed procedure, substantiated by the fact that the plotted adsorption and desorption curves accurately reproduce the experimental data.
The first two numeric columns of Table 2 list the parameter values of the Langmuir adsorption isotherm model that best fit the CH4 and N2 experimental adsorption data at 273 K (Figure 9); these values apply when the framework is in its SP form. The last two columns of Table 2 list the corresponding values if the framework were hypothetically in EP form; given the absence of experimental data, the Langmuir parameters were estimated using the following scaling rules:
q i , ( EP ) = q i , ( SP ) q CO 2 , ( EP ) q CO 2 , ( SP )   and   b i ( EP ) = b i ( SP ) b CO 2 ( EP ) b CO 2 ( SP )
Although these rules are rather crude and their application cannot be considered quantitatively precise, they allow us to explain why the adsorption isotherms of CH4 and N2 do not point to a phase transition between the SP and EP structures (and no hysteresis) in the pressure range tested experimentally by us. The reason is that in the case of CH4 or N2 adsorption, the pressure must be increased considerably, well above the maximum experimental value tested by us, for the term R T 0 P q ( P , T )   d ln P to change the sign of Ω os . A similar reasoning explains why the CO2 adsorption isotherms at 303 K and 323 K do not hint at a phase transition between the SP and EP structures.

3.4. Potential of Zn(dcpa) for CO2/N2 and CO2/CH4 Separation

To assess the potential use of Zn(dcpa) for the adsorptive separation of CO2/N2 and CO2/CH4, we first compared the single-component equilibrium isotherms and the obtained selectivities. Figure 10a compares the isotherms obtained at 303 K for the three gases. The ideal selectivity for an equimolar mixture, α A / B = q A / q B , was calculated using the single-component adsorption capacity ratio; the obtained results are reported in Figure 10b,c for the CO2/N2 and CO2/CH4 selectivities, respectively. Both the CO2/N2 and CO2/CH4 selectivities decreased with the increasing pressure, ranging from 12.8 (at 1 bar) to 6.7 (6 bar) for CO2/N2, and from 2.9 (at 1 bar) to 2.1 (6 bar) for CO2/CH4. The CO2/N2 selectivity was significantly higher than that of CO2/CH4, especially at a lower pressure. Although the reported equilibrium selectivity did not take into account the influence of adsorption kinetics or the impact of a real gas mixture, these results serve as a first evaluation of the good potential of Zn(dcpa) for CO2 separation from CO2/N2 and CO2/CH4 mixtures. Comparing the selectivity of Zn(dcpa) for CO2/N2 with those of the commercial MOFs MIL-53(Al) [31], ZIF-8, [44], and Fe-BTC [45] (Figure 10b,c), it is concluded that the former outperformed the others at a lower pressure. For example, at 1 bar, the order of selectivities was 12.8 (Zn(dcpa)) > 10.5 (Fe-BTC) > 9.6 (MIL-53(Al)) > 5.9 (ZIF-8). The same can be said about the CO2/CH4 selectivity trend: 2.9 (Zn(dcpa)) > 2.7 (MIL-53(Al)) > 2.6 (ZIF-8), although in this case, ZIF-8 surpassed Zn(dcpa) at pressures above 2.4 bar.
The isosteric heat of adsorption, Q st , was determined from the experimental data via the Clausius−Clapeyron equation: ( log   P ) q = const Q st / R T [46] Using this approach, the plot of log   P versus 1 / R T , at constant loading, should give a straight line from which the slope of Q st can be obtained. The plots in Figure 11 of Q st for CO2, CH4, and N2 on Zn(dcpa) as a function of loading show that for CH4, their values were approximately constant within the loading range studied, while for CO2 and N2, a linear increase was observed. It should be noted that in the case of CO2, Q st was plotted for loadings lower than the MOF’s conformational change observed at 273 K, i.e., less than 4 mol/kg. The isosteric heat of the adsorption was higher for CO2 (23–28 kJ/mol), which is in accordance with the values reported in the literature [34], followed by that for CH4 (~19.5 kJ/mol) and N2 (13–15 kJ/mol).

4. Conclusions

Zn(dcpa) MOF was characterized through PXRD, thermogravimetric analysis, and N2 adsorption at 77 K. In line with the XRD and N2 data reported by Liu et al. [34], our results indicate that the framework conformation changes (pore shrinkage/pore expansion) upon removal/loading of guest molecules. The TGA results demonstrate that the MOF is stable up to 650 K.
The adsorption equilibrium isotherms of CO2, CH4, and N2 on Zn(dcpa) at 273 K, 303 K, and 323 K are reported up to 35 bar. The obtained data highlight the interesting behavior of CO2 adsorption at 273 K, which exhibits a stepped isotherm related to the transition between shrunken- and expanded-pore phases at around 22 bar. When observing the desorption branch of the isotherm, a hysteresis loop is present, closing at 7 bar. Although the same behavior is observed for CO2 adsorption at 195 K [34], this is the first report of this effect at 273 K. None of the remaining CO2 (303 K and 323 K), CH4, or N2 isotherms show the same behavior.
The CO2 adsorption equilibrium at 273 K is accurately interpreted using the osmotic thermodynamic theory, which is further refined by considering that the free energy difference between the two metastable structures of Zn(dcpa) is a normally distributed variable due to the distribution of crystal sizes and the defects in a real MOF sample.
Regarding the uptake amounts, Zn(dcpa) can adsorb higher amounts of CO2, followed by CH4 and N2. The CO2/N2 and CO2/CH4 ideal equilibrium selectivities of Zn(dcpa) at 303 K are evaluated for equimolar mixtures, resulting in 12.8 and 2.9, respectively, for a total pressure of 1 bar.
The reported data are essential for the modelling of adsorption-based processes, namely pressure swing adsorption (PSA) and temperature swing adsorption (TSA), for the separation of mixtures containing the studied gases, e.g., biogas upgrading and CO2 capture from flue gases.

Author Contributions

Conceptualization, R.P.P.L.R. and J.P.B.M.; formal analysis, R.P.P.L.R. and J.P.B.M.; investigation, R.P.P.L.R. and I.A.A.C.E.; writing—original draft preparation, R.P.P.L.R.; writing—review and editing, R.P.P.L.R., I.A.A.C.E. and J.P.B.M.; visualization, R.P.P.L.R. and J.P.B.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Associate Laboratory for Green Chemistry (AQV), which is financed by national funds from FCT/MCTES (UIDB/50006/2020 and UIDP/50006/2020). Rui Ribeiro and Isabel Esteves acknowledge financial support from FCT/MCTES through the Norma Transitória DL 57/2016 Program Contract and project IF/01016/2014, respectively. The authors also acknowledge support from the ERANet LAC initiative through project ELAC2014/BEE0367.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Herm, Z.R.; Swisher, J.A.; Smit, B.; Krishna, R.; Long, J.R. Metal-organic frameworks as adsorbents for hydrogen purification and precombustion carbon dioxide capture. J. Am. Chem. Soc. 2011, 133, 5664–5667. [Google Scholar] [CrossRef] [PubMed]
  2. Grande, C.A.; Blom, R.; Andreassen, K.A.; Stensrød, R.E. Experimental results of pressure swing adsorption (PSA) for pre-combustion CO2 capture with metal organic frameworks. Energy Procedia 2017, 114, 2265–2270. [Google Scholar] [CrossRef]
  3. Rowsell, J.L.C.; Yaghi, O.M. Metal–organic frameworks: A new class of porous materials. Micropor. Mesopor. Mat. 2004, 73, 3–14. [Google Scholar] [CrossRef]
  4. Ferey, G. Hybrid porous solids: Past, present, future. Chem. Soc. Rev. 2008, 37, 191–214. [Google Scholar] [CrossRef] [PubMed]
  5. Millward, A.R.; Yaghi, O.M. Metal-organic frameworks with exceptionally high capacity for storage of carbon dioxide at room temperature. J. Am. Chem. Soc. 2005, 127, 17998–17999. [Google Scholar] [CrossRef]
  6. Furukawa, H.; Cordova, K.E.; O’Keeffe, M.; Yaghi, O.M. The chemistry and applications of metal-organic frameworks. Science 2013, 341. [Google Scholar] [CrossRef] [Green Version]
  7. Czaja, A.U.; Trukhan, N.; Muller, U. Industrial applications of metal-organic frameworks. Chem. Soc. Rev. 2009, 38, 1284–1293. [Google Scholar] [CrossRef] [PubMed]
  8. Meek, S.T.; Greathouse, J.A.; Allendorf, M.D. Metal-organic frameworks: A rapidly growing class of versatile nanoporous materials. Adv. Mater. 2011, 23, 249–267. [Google Scholar] [CrossRef]
  9. Mueller, U.; Schubert, M.; Teich, F.; Puetter, H.; Schierle-Arndt, K.; Pastre, J. Metal-organic frameworks—Pospective industrial applications. J. Mater. Chem. 2006, 16, 626–636. [Google Scholar] [CrossRef]
  10. Ribeiro, R.P.P.L.; Antunes, C.L.; Garate, A.U.; Portela, A.F.; Plaza, M.G.; Mota, J.P.B.; Esteves, I.A.A.C. Binderless shaped metal-organic framework particles: Impact on carbon dioxide adsorption. Microporous Mesoporous Mater. 2019, 275, 111–121. [Google Scholar] [CrossRef]
  11. An, Y.; Tian, Y.; Li, Y.; Wei, C.; Tao, Y.; Liu, Y.; Xi, B.; Xiong, S.; Feng, J.; Qian, Y. Heteroatom-doped 3D porous carbon architectures for highly stable aqueous zinc metal batteries and non-aqueous lithium metal batteries. Chem. Eng. J. 2020, 400, 125843. [Google Scholar] [CrossRef]
  12. An, Y.; Tian, Y.; Li, Y.; Xiong, S.; Zhao, G.; Feng, J.; Qian, Y. Green and tunable fabrication of graphene-like N-doped carbon on a 3D metal substrate as a binder-free anode for high-performance potassium-ion batteries. J. Mater. Chem. A 2019, 7, 21966–21975. [Google Scholar] [CrossRef]
  13. Alhamami, M.; Doan, H.; Cheng, C.-H. A review on breathing behaviors of metal-organic-frameworks (MOFs) for gas adsorption. Materials 2014, 7, 3198–3250. [Google Scholar] [CrossRef]
  14. Schneemann, A.; Bon, V.; Schwedler, I.; Senkovska, I.; Kaskel, S.; Fischer, R.A. Flexible metal-organic frameworks. Chem. Soc. Rev. 2014, 43, 6062–6096. [Google Scholar] [CrossRef] [Green Version]
  15. Coudert, F.-X.; Mellot-Draznieks, C.; Fuchs, A.H.; Boutin, A. Double structural transition in hybrid material MIL-53 upon hydrocarbon adsorption: The termodynamics behind the scenes. J. Am. Chem. Soc. 2009, 131, 3442–3443. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Mishra, P.; Edubilli, S.; Uppara, H.P.; Mandal, B.; Gumma, S. Effect of adsorbent history on adsorption characteristics of MIL-53 (Al) metal organic framework. Langmuir 2013, 29, 12162–12167. [Google Scholar] [CrossRef] [PubMed]
  17. Serre, C.; Millange, F.; Thouvenot, C.; Noguès, M.; Marsolier, G.; Louër, D.; Férey, G. Very large breathing effect in the first nanoporous chromium (III)-based solids: MIL-53 or CrIII (OH)·{OC−C6H4−CO2}·{HO2C−C6H4−CO2H}x·H2Oy. J. Am. Chem. Soc. 2002, 124, 13519–13526. [Google Scholar] [CrossRef]
  18. Fairen-Jimenez, D.; Galvelis, R.; Torrisi, A.; Gellan, A.D.; Wharmby, M.T.; Wright, P.A.; Mellot-Draznieks, C.; Duren, T. Flexibility and swing effect on the adsorption of energy-related gases on ZIF-8: Combined experimental and simulation study. Dalton Trans. 2012, 41, 10752–10762. [Google Scholar] [CrossRef] [PubMed]
  19. Fairen-Jimenez, D.; Moggach, S.A.; Wharmby, M.T.; Wright, P.A.; Parsons, S.; Düren, T. Opening the gate: Framework flexibility in ZIF-8 explored by experiments and simulations. J. Am. Chem. Soc. 2011, 133, 8900–8902. [Google Scholar] [CrossRef] [Green Version]
  20. Loiseau, T.; Serre, C.; Huguenard, C.; Fink, G.; Taulelle, F.; Henry, M.; Bataille, T.; Ferey, G. A rationale for the large breathing of the porous aluminum terephthalate (MIL-53) upon hydration. Chem. Eur. J. 2004, 10, 1373–1382. [Google Scholar] [CrossRef] [PubMed]
  21. Hamon, L.; Serre, C.; Devic, T.; Loiseau, T.; Millange, F.; Férey, G.; Weireld, G.D. Comparative study of hydrogen sulfide adsorption in the MIL-53 (Al, Cr, Fe), MIL-47 (V), MIL-100 (Cr), and MIL-101 (Cr) metal-rganic frameworks at room temperature. J. Am. Chem. Soc. 2009, 131, 8775–8777. [Google Scholar] [CrossRef]
  22. Chen, L.; Mowat, J.P.S.; Fairen-Jimenez, D.; Morrison, C.A.; Thompson, S.P.; Wright, P.A.; Düren, T. Elucidating the breathing of the metal–organic framework MIL-53 (Sc) with ab initio molecular dynamics simulations and in situ X-ray powder diffraction experiments. J. Am. Chem. Soc. 2013, 135, 15763–15773. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Boutin, A.; Coudert, F.X.; Springuel-Huet, M.A.; Neimark, A.V.; Férey, G.; Fuchs, A.H. The behavior of flexible MIL-53 (Al) upon CH4 and CO2 adsorption. J. Phys. Chem. C 2010, 114, 22237–22244. [Google Scholar] [CrossRef] [Green Version]
  24. Liu, Y.; Her, J.-H.; Dailly, A.; Ramirez-Cuesta, A.J.; Neumann, D.A.; Brown, C.M. Reversible structural transition in MIL-53 with large temperature hysteresis. J. Am. Chem. Soc. 2008, 130, 11813–11818. [Google Scholar] [CrossRef] [PubMed]
  25. Beurroies, I.; Boulhout, M.; Llewellyn, P.L.; Kuchta, B.; Férey, G.; Serre, C.; Denoyel, R. Using pressure to provoke the structural transition of metal-rganic frameworks. Angew. Chem. Int. Ed. 2010, 49, 7526–7529. [Google Scholar] [CrossRef] [Green Version]
  26. Neimark, A.V.; Coudert, F.-X.; Triguero, C.; Boutin, A.; Fuchs, A.H.; Beurroies, I.; Denoyel, R. Structural transitions in MIL-53 (Cr): View from outside and inside. Langmuir 2011, 27, 4734–4741. [Google Scholar] [CrossRef] [Green Version]
  27. Serre, C.; Bourrelly, S.; Vimont, A.; Ramsahye, N.A.; Maurin, G.; Llewellyn, P.L.; Daturi, M.; Filinchuk, Y.; Leynaud, O.; Barnes, P.; et al. An explanation for the very large breathing effect of a metal–organic framework during CO2 adsorption. Adv. Mater. 2007, 19, 2246–2251. [Google Scholar] [CrossRef]
  28. Mounfield, W.P., III; Walton, K.S. Effect of synthesis solvent on the breathing behavior of MIL-53 (Al). J. Colloid Interface Sci. 2015, 447, 33–39. [Google Scholar] [CrossRef]
  29. Heymans, N.; Vaesen, S.; De Weireld, G. A complete procedure for acidic gas separation by adsorption on MIL-53 (Al). Microporous Mesoporous Mater. 2012, 154, 93–99. [Google Scholar] [CrossRef]
  30. Deniz, E.; Karadas, F.; Patel, H.A.; Aparicio, S.; Yavuz, C.T.; Atilhan, M. A combined computational and experimental study of high pressure and supercritical CO2 adsorption on Basolite MOFs. Microporous Mesoporous Mater. 2013, 175, 34–42. [Google Scholar] [CrossRef]
  31. Camacho, B.C.R.; Ribeiro, R.P.P.L.; Esteves, I.A.A.C.; Mota, J.P.B. Adsorption equilibrium of carbon dioxide and nitrogen on the MIL-53 (Al) metal organic framework. Sep. Purif. Technol. 2015, 141, 150–159. [Google Scholar] [CrossRef]
  32. Gonzalez-Nelson, A.; Coudert, F.X.; van der Veen, M.A. Rotational dynamics of linkers in metal-organic frameworks. Nanomaterials 2019, 9, 330. [Google Scholar] [CrossRef] [Green Version]
  33. Hefti, M.; Joss, L.; Bjelobrk, Z.; Mazzotti, M. On the potential of phase-change adsorbents for CO2 capture by temperature swing adsorption. Faraday Discuss. 2016, 192, 153–179. [Google Scholar] [CrossRef] [Green Version]
  34. Liu, B.; Li, Y.; Hou, L.; Yang, G.; Wang, Y.Y.; Shi, Q.-Z. Dynamic Zn-based metal-organic framework: Step wise adsorption, hysteretic desorption and selective carbon dioxide uptake. J. Mater. Chem. 2013, 1, 6535–6538. [Google Scholar] [CrossRef]
  35. Ribeiro, R.P.P.L.; Barreto, J.; GrossoXavier, M.D.; Martins, D.; Esteves, I.A.A.C.; Branco, M.; Tirolien, T.; Mota, J.P.B.; Bonfait, G. Cryogenic neon adsorption on Co3 (ndc) 3 (dabco) metal-organic framework. Microporousand Mesoporous Mater. 2020, 298, 110055. [Google Scholar] [CrossRef]
  36. Lyubchyk, A.; Esteves, I.A.A.C.; Cruz, F.J.A.L.; Mota, J.P.B. Experimental and theoretical studies of supercritical methane adsorption in the MIL-53 (Al) metal organic framework. J. Phys. Chem. 2011, 115, 20628–20638. [Google Scholar] [CrossRef]
  37. Ribeiro, R.P.P.L.; Camacho, B.C.R.; Lyubchyk, A.; Esteves, I.A.A.C.; Cruz, F.J.A.L.; Mota, J.P.B. Experimental and computational study of ethane and ethylene adsorption in the MIL-53 (Al) metal organic framework. Microporousand Mesoporous Mater. 2016, 230, 154–165. [Google Scholar] [CrossRef]
  38. Dreisbach, F.; Staudt, R.; Keller, J.U. High pressure adsorption data of methane, nitrogen, carbon dioxide and their binary and ternary mixtures on activated carbon. Adsorption 1999, 5, 215–227. [Google Scholar] [CrossRef]
  39. Coudert, F.-X.; Jeffroy, M.; Fuchs, A.H.; Boutin, A.; Mellot-Draznieks, C. Thermodynamics of guest-induced structural transitions in hybrid organic−inorganic frameworks. J. Am. Chem. Soc. 2008, 130, 14294–14302. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  40. Mota, J.P.B.; Martins, D.; Lopes, D.; Catarino, I.; Bonfait, G. Structural transitions in the MIL-53 (Al) metal–organic framework upon cryogenic hydrogen adsorption. J. Phys. Chem. 2017, 121, 24252–24263. [Google Scholar] [CrossRef]
  41. Hiraide, S.; Sakanaka, Y.; Kajiro, H.; Kawaguchi, S.; Miyahara, M.T.; Tanaka, H. High-through put gas separation by flexible metal–organic frameworks with fast gating and thermal management capabilities. Nat. Commun. 2020, 11, 3867. [Google Scholar] [CrossRef]
  42. Ghysels, A.; Vanduyfhuys, L.; Vandichel, M.; Waroquier, M.; Van Speybroeck, V.; Smit, B. On the thermodynamics of framework breathing: A free energy model for gas adsorption in MIL-53. J. Phys. Chem. 2013, 117, 11540–11554. [Google Scholar] [CrossRef] [Green Version]
  43. Jeffroy, M.; Fuchs, A.H.; Boutin, A. Structural changes in nanoporous solids due to fluid adsorption: Thermodynamic analysis and Monte Carlo simulations. Chem. Commun. 2008, 3275–3277. [Google Scholar] [CrossRef] [PubMed]
  44. Ferreira, T.J.; Vera, A.T.; De Moura, B.A.; Esteves, L.M.; Tariq, M.; Esperança, J.M.S.S.; Esteves, I.A.A.C. Paramagnetic ionic liquid/metal organic framework composites for CO2/CH4 and CO2/N2 separations. Front. Chem. 2020, 8, 590191. [Google Scholar] [CrossRef] [PubMed]
  45. Nabais, A.R.; Ribeiro, R.P.P.L.; Mota, J.P.B.; Alves, V.D.; Esteves, I.A.A.C.; Neves, L.A. CO2/N2 gas separation using Fe (BTC)-based mixed matrix membranes: A view on the adsorptive and filler properties of metal-organic frameworks. Sep. Purif. Technol. 2018, 202, 174–184. [Google Scholar] [CrossRef]
  46. Poling, B.E.; Prausnitz, J.M.; O’Connell, J.P. The Properties of Gases and Liquids; McGraw-Hill: New York, NY, USA, 2001. [Google Scholar]
Figure 1. View of the Zn(dcpa) framework along the c-axis (Zn: blue, O: red, C: grey, and H: white). Data generated from the CIF file reported by Liu et al. [34].
Figure 1. View of the Zn(dcpa) framework along the c-axis (Zn: blue, O: red, C: grey, and H: white). Data generated from the CIF file reported by Liu et al. [34].
Energies 14 05598 g001
Figure 2. PXRD diffraction pattern for activated, as made, and theoretical Zn(dcpa) samples.
Figure 2. PXRD diffraction pattern for activated, as made, and theoretical Zn(dcpa) samples.
Energies 14 05598 g002
Figure 3. TGA results for Zn(dcpa) as a function of temperature (heating rate of 3 K/min).
Figure 3. TGA results for Zn(dcpa) as a function of temperature (heating rate of 3 K/min).
Energies 14 05598 g003
Figure 4. N2 adsorption equilibrium isotherm at 77 K on Zn(dcpa) in linear (a) and log (b) scales. Filled and empty symbols represent the adsorption and desorption data, respectively.
Figure 4. N2 adsorption equilibrium isotherm at 77 K on Zn(dcpa) in linear (a) and log (b) scales. Filled and empty symbols represent the adsorption and desorption data, respectively.
Energies 14 05598 g004
Figure 5. Absolute adsorption equilibrium isotherms of CO2 on Zn(dcpa) at 273 K (), 303 K (), and 323 K (●). Filled and empty symbols represent adsorption and desorption experimental data, respectively.
Figure 5. Absolute adsorption equilibrium isotherms of CO2 on Zn(dcpa) at 273 K (), 303 K (), and 323 K (●). Filled and empty symbols represent adsorption and desorption experimental data, respectively.
Energies 14 05598 g005
Figure 6. Absolute adsorption equilibrium isotherms of CH4 on Zn(dcpa) at 273 K (), 303 K (), and 323 K (●). Filled and empty symbols represent adsorption and desorption experimental data, respectively.
Figure 6. Absolute adsorption equilibrium isotherms of CH4 on Zn(dcpa) at 273 K (), 303 K (), and 323 K (●). Filled and empty symbols represent adsorption and desorption experimental data, respectively.
Energies 14 05598 g006
Figure 7. Absolute adsorption equilibrium isotherms of N2 on Zn(dcpa) at 273 K (), 303 K (), and 323 K (●). Filled and empty symbols represent adsorption and desorption experimental data, respectively.
Figure 7. Absolute adsorption equilibrium isotherms of N2 on Zn(dcpa) at 273 K (), 303 K (), and 323 K (●). Filled and empty symbols represent adsorption and desorption experimental data, respectively.
Energies 14 05598 g007
Figure 8. Fitting of the osmotic thermodynamic theory to the adsorption (●) and desorption (○) branches of the experimental CO2 isotherm at 273 K. Dashed lines: fitted Langmuir adsorption isotherms if the framework were rigid and constrained to its EP (– – –) and SP (– – –) forms; solid lines: predicted adsorption (–––) and desorption (–––) branches of the isotherm at 273 K.
Figure 8. Fitting of the osmotic thermodynamic theory to the adsorption (●) and desorption (○) branches of the experimental CO2 isotherm at 273 K. Dashed lines: fitted Langmuir adsorption isotherms if the framework were rigid and constrained to its EP (– – –) and SP (– – –) forms; solid lines: predicted adsorption (–––) and desorption (–––) branches of the isotherm at 273 K.
Energies 14 05598 g008
Figure 9. Fitting of the Langmuir isotherm model to the CH4 () and N2 (●) experimental adsorption equilibrium isotherms at 273 K. Filled and empty symbols denote experimental adsorption and desorption data, respectively. Solid lines: fitted CH4 Langmuir isotherm for rigid framework constrained to its EP (red) and SP (blue) forms; dashed lines: fitted N2 Langmuir isotherm for rigid framework constrained to its EP (red) and SP (blue) forms.
Figure 9. Fitting of the Langmuir isotherm model to the CH4 () and N2 (●) experimental adsorption equilibrium isotherms at 273 K. Filled and empty symbols denote experimental adsorption and desorption data, respectively. Solid lines: fitted CH4 Langmuir isotherm for rigid framework constrained to its EP (red) and SP (blue) forms; dashed lines: fitted N2 Langmuir isotherm for rigid framework constrained to its EP (red) and SP (blue) forms.
Energies 14 05598 g009
Figure 10. (a) Adsorption equilibrium isotherms of CO2 (), CH4 (), and N2 (●) on Zn(dcpa) at 303 K. Filled and empty symbols represent adsorption and desorption data, respectively; (b) CO2/N2 and (c) CO2/CH4 equilibrium selectivity at 303 K as a function of pressure for Zn(dcpa) and commercial MOFs MIL-53(Al) [31], ZIF-8 [44], and Fe-BTC [45].
Figure 10. (a) Adsorption equilibrium isotherms of CO2 (), CH4 (), and N2 (●) on Zn(dcpa) at 303 K. Filled and empty symbols represent adsorption and desorption data, respectively; (b) CO2/N2 and (c) CO2/CH4 equilibrium selectivity at 303 K as a function of pressure for Zn(dcpa) and commercial MOFs MIL-53(Al) [31], ZIF-8 [44], and Fe-BTC [45].
Energies 14 05598 g010
Figure 11. CO2 (blue), CH4 (black), and N2 (red) isosteric heats of adsorption, Q s t , as a function of loading, q.
Figure 11. CO2 (blue), CH4 (black), and N2 (red) isosteric heats of adsorption, Q s t , as a function of loading, q.
Energies 14 05598 g011
Table 1. Fitted parameters of the osmotic thermodynamic theory to the adsorption (Ads) and desorption (Des) branches of the experimental CO2 isotherm at 273 K. F host F host ( EP ) F host ( SP ) .
Table 1. Fitted parameters of the osmotic thermodynamic theory to the adsorption (Ads) and desorption (Des) branches of the experimental CO2 isotherm at 273 K. F host F host ( EP ) F host ( SP ) .
q b P μ σ P F h o s t
(mol/kg)(atm −1) (atm)(atm)(J/g)
SP5.651.30Ads23.01.3−4.45
EP12.20.143Des7.00.5−10.7
Table 2. Fitting of the Langmuir adsorption isotherm model to the CH4 and N2 experimental adsorption data at 273 K. The parameters for the EP form are estimated as q i , ( EP ) = q i , ( SP ) q CO 2 , ( EP ) / q CO 2 , ( SP ) and b i ( EP ) = b i ( SP ) b CO 2 ( EP ) / b CO 2 ( SP ) .
Table 2. Fitting of the Langmuir adsorption isotherm model to the CH4 and N2 experimental adsorption data at 273 K. The parameters for the EP form are estimated as q i , ( EP ) = q i , ( SP ) q CO 2 , ( EP ) / q CO 2 , ( SP ) and b i ( EP ) = b i ( SP ) b CO 2 ( EP ) / b CO 2 ( SP ) .
q ( SP ) b ( SP ) q ( EP ) b ( EP )
(mol/kg)(atm −1)(mol/kg)(atm −1)
N 2 3.6810.0897.9480.010
CH 4 4.3280.4119.3460.045
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ribeiro, R.P.P.L.; Esteves, I.A.A.C.; Mota, J.P.B. Adsorption of Carbon Dioxide, Methane, and Nitrogen on Zn(dcpa) Metal-Organic Framework. Energies 2021, 14, 5598. https://doi.org/10.3390/en14185598

AMA Style

Ribeiro RPPL, Esteves IAAC, Mota JPB. Adsorption of Carbon Dioxide, Methane, and Nitrogen on Zn(dcpa) Metal-Organic Framework. Energies. 2021; 14(18):5598. https://doi.org/10.3390/en14185598

Chicago/Turabian Style

Ribeiro, Rui P. P. L., Isabel A. A. C. Esteves, and José P. B. Mota. 2021. "Adsorption of Carbon Dioxide, Methane, and Nitrogen on Zn(dcpa) Metal-Organic Framework" Energies 14, no. 18: 5598. https://doi.org/10.3390/en14185598

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop