Next Article in Journal
Layer-by-Layer Heparinization of the Cell Surface by Using Heparin-Binding Peptide Functionalized Human Serum Albumin
Next Article in Special Issue
Thermal Energy Storage of R1234yf, R1234ze, R134a and R32/MOF-74 Nanofluids: A Molecular Simulation Study
Previous Article in Journal
Band Structures and Transport Properties of High-Performance Half-Heusler Thermoelectric Materials by First Principles
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Energy Storage Analysis of a Mixed R161/MOF-5 Nanoparticle Nanofluid Based on Molecular Simulations

1
Key Laboratory of Low-grade Energy Utilization Technology & System, Ministry of Education, College of Power Engineering, Chongqing University, Chongqing 400044, China
2
College of Aerospace Engineering, Chongqing University, Chongqing 400044, China
3
Key Lab. of Opto-Electronic Technology & System, Ministry of Education, Chongqing University, Chongqing 400044, China
*
Author to whom correspondence should be addressed.
Materials 2018, 11(5), 848; https://doi.org/10.3390/ma11050848
Submission received: 29 April 2018 / Revised: 17 May 2018 / Accepted: 18 May 2018 / Published: 20 May 2018
(This article belongs to the Special Issue Nanotechnology in Renewable Energy)

Abstract

:
The thermal properties of refrigerants can be modified by adding porous nanoparticles into them. Here, molecular simulations, including molecular dynamics and grand canonical Monte Carlo, were employed to study the thermal energy storage properties of an R161/MOF-5 nanofluid. The results show that the thermodynamic energy change of MOF-5 nanoparticles is linear to the temperature. The adsorption heat calculated by grand canonical Monte Carlo is close to that calculated by the Clausius–Clapeyron equation. Additionally, a negative enhancement of the thermal energy storage capacity of the R161/MOF-5 nanofluid is found near the phase transition area.

1. Introduction

With the rapid development of modern society, serious problems, such as energy shortage and environmental pollution, have become increasingly prominent [1,2]. Energy conservation and emissions reduction are effective ways to alleviate the energy and environmental problems. The thermodynamic cycle is the main approach for energy conversion. Additionally, the working fluid is the energy carrier of the thermodynamic cycle. Therefore, using various means to improve the thermophysical properties of working fluid can improve the efficiency of energy conversion.
Dispersing nanoparticles into traditional working fluids, e.g., water, alcohol, oil, and refrigerants, to produce uniform and stable nanofluids can effectively improve the thermal conductivity of working fluids [3,4,5,6]. Nanofluids have great prospects for application in the fields of energy, chemical engineering, automobiles, building and construction, etc. The enhancement of the thermal conductivity of nanofluids is mainly due to the effective medium theory [3]. Also, the layering phenomenon, Brownian motion, clustering, ballistic phonon motion, thermal boundary resistance, and mass difference scattering will impact on the thermal conductivity of nanofluids [7,8,9,10].
Besides this, a nanofluid with porous materials can be used for thermal energy storage [11]. Carbon nanotubes, zeolites, and Metal-Organic Frameworks (MOFs) are common nanoporous materials. These materials have a very large specific surface area compared to conventional materials. Thus, extra energy can be stored and output by the fluid desorbing/adsorbing in the nanoporous materials. Chen et al. [12] performed comprehensive research on the energy storage of carbon nanotube nanofluids under the action of heat, force, and electric coupling. McGrail et al. [13] proposed to use metal-organic heat carrier nanofluids (MOHCs, a refrigerant to which has been added MOF nanoparticles to form the nanofluid) to improve the efficiency of the Organic Rankine cycle (ORC, using refrigerants as the working fluid for the Rankine cycle). They reported [14] noticeable adsorption and reversible desorption of refrigerant molecules in MOFs. Therefore, MOHCs have great potential in low-grade energy utilization and refrigeration cycles.
Theoretically, the heat adsorption energy of MOHCs (hMOHCs) is mainly composed of three parts: (i) the enthalpy change of the working fluid (∆hFluid); (ii) the thermodynamic energy change of MOF particles ((∫CpdT)MOFs); and (iii) the desorption heat of the fluid in MOFs (∆hdesorption), i.e., [11,13],
∆hMOHCs = (1 − x) ∙ ∆hFluid + x ∙ (∫CpdT)MOFs + x ∙ ∆hdesorption
where x is the mass fraction of MOFs in the MOHC. Additionally, Equation (1) can be written as
∆hMOHCs = ∆hFluid + x ∙ ((∫CpdT)MOFs + ∆hdesorption − ∆hFluid)
It can be inferred from Equation (2) that MOHCs are able to store more thermal energy than pure fluid when the sum of the thermodynamic energy change of MOF particles and the desorption heat of the fluid in MOFs is larger than the enthalpy change of the working fluid.
MOFs are a class of organic–inorganic hybrid materials constructed from organic ligands and inorganic metal units. They own a large specific surface area, large porosity, a diverse structure, and thermal stability, which have provided MOFs with a wide range of applications in the fields of energy, chemicals, materials, and so on [15]. In the numerous MOF structures, MOF-5 [16] is an MOF that consists of Zn4O clusters connected by 1,4-benzodicarboxylate linkers as shown in Figure 1. It has a specific surface area of 2000 m2/g and a good thermostability whose structure remains stable below 573 K [17]. Fluoroethane, also known as R161, is one of the most used organic refrigerants with zero ozone depression potential, low global warming potential, and good thermal properties. Additionally, the adsorption properties of R161 in MOF-5 have been rarely reported [14,18]. Thus, the thermal energy storage property of an MOHC based on R161 and MOF-5 is investigated in the present paper.
Besides this, due to the length scale of the nanopore structure in MOFs being small, it is difficult to investigate the thermal energy storage mechanism of the organic working fluid in MOFs by using conventional experimental and theoretical methods. Molecular simulation [19] is an ideal tool to reveal the micro-mechanisms of materials. Therefore, molecular simulation is employed to investigate the thermal energy storage of R161/MOF-5-based MOHCs. The work is also expected to provide useful insights on adsorption refrigeration and heat pumps [20].

2. Method and Computational Details

According to the thermodynamic principles discussed in Equation (2), the thermodynamic parameters of a pure working fluid are classic, and can be obtained by an experimental or a theoretical method. Since MOFs are a novel material, the thermal properties of MOFs and the adsorption properties of the refrigerant in an MOF need further research. Therefore, here, the thermodynamic parameters of R161 were obtained from the National Institute of Standards and Technology (NIST [21]). The thermodynamic energy change of MOF-5 can be obtained by molecular dynamics (MD) simulation [22]. Additionally, the desorption heat of R161 in MOF-5 can be easily calculated from grand canonical Monte Carlo (GCMC) simulations [23].

2.1. Simulation Models

The calculation model of MOF-5 comprising 2 × 2 × 2 unit cells (X: 51.788 Å, Y: 51.788 Å, Z: 51.788 Å) is shown in Figure 1 with 1536 carbon atoms, 832 oxygen atoms, 768 hydrogen atoms, and 256 Zn atoms. The structure of R161 (CH3CH2F) is shown in Figure 2. The MD simulations and GCMC simulations were performed using Materials Studio [24]. In the simulations, the COMPASS force field [25] was used to describe the intra and inter molecular interactions. The Ewald method [26] was used to deal with the long-range Coulombic interactions. Periodic boundary conditions were applied in the X, Y, and Z directions in all of our simulations.

2.2. MD Simulation Details

The MD simulations were performed in the Forcite module of Materials Studio. The timestep was set as 1 fs and the system was equilibrated for 100 ps to calculate the thermodynamic energies of MOF-5 nanoparticles at 280, 300, 320, 340, 360, 380, 400, and 420 K. The simulations were computed in the canonical ensemble (NVT) ensemble with the Berendsen method [27] to control the temperature.

2.3. GCMC Simulation Details

The GCMC simulations were performed in the Sorption module of Materials Studio. The adsorption isotherms (280, 300, 320, 340, 360, 380, 400, and 420 K) of the adsorption of R161 in MOF-5 nanoparticles were calculated from 0 to 6000 kPa. The fugacity was calculated by the Peng–Robinson equation. For each point of the adsorption isotherms, the equilibration time was 10,000 steps with another 100,000 steps for statistics.

3. Results and Discussion

3.1. Thermodynamic Energy of MOF-5

The thermodynamic energies of MOF-5 at different temperatures are plotted in Figure 3. It shows that the thermodynamic energy of MOF-5 increases linearly with the temperature rise. The increment of internal energy per unit of temperature is Cp of MOF. Here, the ρcp = 1.24(J/cm3·K), which is close to the value of a typical MOF, 1.44 J/cm3·K [13,28].

3.2. Adsorption Isotherms and Enthalpy of Desorption

The adsorption isotherms of R161 adsorbed in MOF-5 are shown in Figure 4. The adsorption isotherms are also fitted by the Langmuir equation, details of which are shown in Appendix A. The adsorption snapshot in Figure 4 shows that R161 mainly distributes around the O and Zn atoms of MOF-5 because these atoms are more attractive than the benzene ligand. Also, the adsorption isotherms decrease as the temperature rises. This is because the raised temperature results in the increased motion of molecules. Thus, the kinetic energy of the fluid molecules can overcome the adsorption energy of the MOF structure. This is the basis for thermal energy storage in MOHCs.
The enthalpy of desorption can be calculated by the Clausius–Clapeyron (C–C) equation based on the adsorption isotherm,
( δ ln P δ T ) θ = q s t R T 2
It can be rewritten as:
ln P = q s t R 1 T + C
Thus, the desorption heat qst can be obtained by calculating the slope relationship between lnP and 1/T, which is shown in Figure 5. Additionally, the results of the desorption heat calculated by the C–C equation and GCMC simulations are listed in Table 1. The desorption heat calculated by GCMC simulations is close to that calculated by the C–C equation. With consideration of the calculation error of the C–C equation, the desorption heat calculated by GCMC will be used to calculate the thermal energy storage in the following section.

3.3. Thermal Energy Storage

Then, the thermal energy storage capacity of MOHCs can be computed according to Equation (1) or Equation (2). As aforementioned, the enthalpy difference of MOHCs at a different temperature is the thermal energy storage capacity. Thus, the thermal energy storage capacity of MOHCs with a different mass ratio of MOF-5 is presented as Figure 6. The temperature of the cold source was set as 280 K. Here, the enthalpy difference of pure R161 was considered to be the reference working fluid. Also, the thermodynamic energy change of MOF particles and the desorption heat of the fluid in MOFs are calculated and plotted in Figure 7. Note that the R161 experiences a phase transition when the temperature difference is about 100 K. The enthalpy difference of pure R161 experiences a sharp increase near this temperature range. The thermodynamic energy change of MOF particles and the desorption heat of the fluid in MOFs linearly increase as the temperature difference rises. It can be concluded that the thermal energy storage property of MOHCs is enhanced when the R161 is in a liquid state. Additionally, the more MOF-5 nanoparticles are added, the more thermal energy MOHCs can store. However, the MOHCs performed a negative enhancement of their thermal energy storage properties near the phase transition point of R161. This is because the enthalpy difference of the phase transition of R161 is large, which is more than the sum of the thermodynamic energy change of MOF-5 and the desorption heat of R161 in MOF-5 at this temperature range. Finally, the thermal energy storage property of MOHCs is enhanced when the temperature difference is over 140 K. Obviously, the sum of the thermodynamic energy change of MOF-5 and the desorption heat of R161 in MOF-5 is larger than the enthalpy difference of R161 when the temperature difference is over 140 K. This indicates that the MOHCs cannot be used for enhancing the thermal energy storage of a working fluid in certain conditions.

4. Conclusions

In this paper, molecular simulations (MD and GCMC methods), as well as thermodynamic calculation, are employed to investigate thermal energy storage by the adsorption of R161 in MOF-5. The R161/MOF-5 mixed MOHCs noticeably enhanced the thermal energy storage property of R161 in the liquid state. The thermal energy storage capacity of the studied MOHCs is less than that of pure R161 near the phase transition range. The MOHCs can be used for thermal energy storage when the sum of the thermodynamic energy change of an MOF and the desorption heat of the fluid in the MOF’s structure is larger than the enthalpy difference of the working fluid.

Author Contributions

Conceptualization, Leilei Li; Methodology, Qiang Wang and Leilei Li; Software, Shengli Tang; Validation, Qiang Wang and Shengli Tang; Formal Analysis, Qiang Wang; Investigation, Shengli Tang; Resources, Shengli Tang; Data Curation, Qiang Wang; Writing-Original Draft Preparation, Qiang Wang; Writing-Review & Editing, Leilei Li; Visualization, Shengli Tang; Supervision, Leilei Li; Project Administration, Leilei Li; Funding Acquisition, Leilei Li.

Funding

Please add: This research was funded by [National Natural Science Foundation of China] grant number [51506013], [the Fundamental Research Funds for the Central Universities of China] grant number [106112016CDJXY320002, 106112017CDJPT120003], and [the Chongqing Research Program of Basic Research and Frontier Technology] grant number [No. cstc2016jcyjA0300].

Acknowledgments

The authors wish to acknowledge useful discussions with Dr. Qibin Li (Chongqing University).

Conflicts of Interest

The authors declare no conflict of interest.

Nomenclature

ΔhMOHCsThe enthalpy of MOHCs (kJ/kg)
ΔhFluidThe enthalpy of pure organic fluid (kJ/kg)
ΔhdesorptionThe enthalpy of desorption (kJ/mol)
TTemperature (K)
CpThe heat capacity of MOFs (kcal/(mol·K))
qstThe adsorption heat (kcal/(mol·K))
xThe mass fraction of MOF in MOHCs
ρDensity (g/cm3)

Appendix A. Adsorption Isotherms Fitted by the Langmuir Equation

Langmuir equation: M = a·b·p/(1 + a·p), where a and b are adsorption constants.
Table A1. The Langmuir equation fitting for R161 adsorption in MOF-5.
Table A1. The Langmuir equation fitting for R161 adsorption in MOF-5.
Temperature (K)Constant aConstant b
2800.0069021.027
3000.0058420.798
3200.0056420.567
3400.0043520.187
3600.0036419.670
3800.0032519.008
4000.0029918.340
4200.0024517.641

References

  1. Zhang, C.; Liu, C.; Wang, S.; Xu, X.; Li, Q. Thermo-economic comparison of subcritical organic Rankine cycle based on different heat exchanger configurations. Energy 2017, 123, 728–741. [Google Scholar] [CrossRef]
  2. Chan, C.W.; Ling-Chin, J.; Roskilly, A.P. A review of chemical heat pumps, thermodynamic cycles and thermal energy storage technologies for low grade heat utilisation. Appl. Therm. Eng. 2013, 50, 1257–1273. [Google Scholar] [CrossRef]
  3. Xuan, Y.; Li, Q. Heat Transfer Enhancement of Nanofluids. Int. J. Heat Fluid Flow 2000, 21, 58–64. [Google Scholar] [CrossRef]
  4. Choi, S.U.S.; Zhang, Z.G.; Yu, W.; Lockwood, F.E.; Grulke, E.A. Anomalous thermal conductivity enhancement in nanotube suspensions. Appl. Phys. Lett. 2001, 79, 2252–2254. [Google Scholar] [CrossRef]
  5. Li, Q.; Yu, Y.; Liu, Y.; Liu, C.; Lin, L. Thermal properties of the mixed n-octadecane/Cu nanoparticle nanofluids during phase transition: A molecular dynamics study. Materials 2017, 10, 38. [Google Scholar] [CrossRef] [PubMed]
  6. Celen, A.; Cebi, A.; Aktas, M.; Mahian, O.; Dalkilic, A.S.; Wongwises, S. A review of nanorefrigerants: Flow characteristics and applications. Int. J. Refrig. 2014, 44, 125–140. [Google Scholar] [CrossRef]
  7. Milanese, M.; Iacobazzi, F.; Colangelo, G.; Risi, A.D. An investigation of layering phenomenon at the liquid-solid interface in Cu and CuO based nanofluids. Int. J. Heat Mass Transf. 2016, 103, 564–571. [Google Scholar] [CrossRef]
  8. Iacobazzi, F.; Milanese, M.; Colangelo, G.; Lomascolo, M.; Risi, A.D. An explanation of the Al2O3 nanofluid thermal conductivity based on the phonon theory of liquid. Energy 2016, 116, 786–794. [Google Scholar] [CrossRef]
  9. Li, Q.; Liu, C. Molecular dynamics simulation of heat transfer with effects of fluid–lattice interactions. Int. J. Heat Mass Transf. 2012, 55, 8088–8092. [Google Scholar] [CrossRef]
  10. Li, Q.; Liu, C.; Chen, X. Molecular dynamics simulation of sulphur nucleation in S–H2S system. Mol. Phys. 2014, 112, 947–955. [Google Scholar] [CrossRef]
  11. Li, Q.; Xiao, Y.; Shi, X.; Song, S. Rapid evaporation of water on graphene/graphene-oxide: A molecular dynamics study. Nanomaterials 2017, 7, 265. [Google Scholar] [CrossRef] [PubMed]
  12. Chen, X.; Xu, B.; Liu, L. Nanoscale fluid mechanics and energy conversion. Appl. Mech. Rev. 2014, 66, 050803. [Google Scholar] [CrossRef]
  13. McGrail, B.P.; Thallapally, P.K.; Blanchard, J.; Nune, S.K.; Jenks, J.J.; Dang, L.X. Metal-organic heat carrier nanofluids. Nano Energy 2013, 2, 845–855. [Google Scholar] [CrossRef]
  14. Annapureddy, H.V.R.; Motkuri, R.K.; Nguyen, P.T.M.; Truong, T.B.; Thallapally, P.K.; McGrail, B.P.; Dang, L.X. Computational studies of adsorption in metal organic frameworks and interaction of nanoparticles in condensed phases. Mol. Simul. 2014, 40, 571–584. [Google Scholar] [CrossRef]
  15. James, S.L. Metal-organic frameworks. Chem. Soc. Rev. 2003, 32, 276–288. [Google Scholar] [CrossRef] [PubMed]
  16. Kaye, S.S.; Daily, A.; Yaghi, O.M.; Long, J.R. Impact of preparation and handling on the hydrogen storage properties of Zn4O (1, 4-benzenedicarboxylate) 3 (MOF-5). J. Am. Chem. Soc. 2007, 129, 14176–14177. [Google Scholar] [CrossRef] [PubMed]
  17. Lu, C.M.; Liu, J.; Xiao, K.; Harris, A.T. Microwave enhanced synthesis of MOF-5 and its CO2 capture ability at moderate temperatures across multiple capture and release cycles. Chem. Eng. J. 2010, 156, 465–470. [Google Scholar] [CrossRef]
  18. Getman, R.B.; Bae, Y.S.; Wilmer, C.E.; Snurr, R.Q. Review and analysis of molecular simulations of methane, hydrogen, and acetylene storage in metal-organic frameworks. Chem. Rev. 2012, 112, 703–723. [Google Scholar] [CrossRef] [PubMed]
  19. Frenkel, D.; Smit, B. Understanding Molecular Simulation: From Algorithms to Applications, 2nd ed.; Academic Press: New York, NY, USA, 2002. [Google Scholar]
  20. Henninger, S.K.; Habib, H.A.; Janiak, C. MOFs as adsorbents for low temperature heating and cooling applications. J. Am. Chem. Soc. 2009, 131, 2776–2777. [Google Scholar] [CrossRef] [PubMed]
  21. National Institute of Standards and Technology (NIST). Available online: http://webbook.nist.gov/chemistry/fluid/ (accessed on 15 February 2018).
  22. Li, Q.; Wang, M.; Liang, Y.; Lin, L.; Fu, T.; Wei, P.; Peng, T. Molecular dynamics simulations of aggregation of copper nanoparticles with different heating rates. Phys. E Low-Dimens. Syst. Nanostruct. 2017, 90, 137–142. [Google Scholar] [CrossRef]
  23. Lei, G.; Liu, C.; Li, Q.; Xu, X. Graphyne nanostructure as a potential adsorbent for separation of H2S/CH4 mixture: Combining grand canonical Monte Carlo simulations with ideal adsorbed solution theory. Fuel 2016, 182, 210–219. [Google Scholar] [CrossRef]
  24. Accelrys, I. Materials Studio; Accelrys Software Inc.: San Diego, CA, USA, 2010. [Google Scholar]
  25. Sun, H. COMPASS:  An ab Initio force-field optimized for condensed-phase applicationsoverview with details on alkane and benzene compounds. J. Phys. Chem. B 1998, 102, 7338–7364. [Google Scholar] [CrossRef]
  26. Ewald, P.P. Die berechnung optischer und elektrostatischer gitterpotentiale. Ann. Phys. 1921, 369, 253–287. [Google Scholar] [CrossRef]
  27. Berendsen, H.J.C.; Postma, J.P.M.; Gunsteren, W.F.V.; Dinola, A.; Haak, J.R. Molecular dynamics with coupling to an external bath. J. Chem. Phys. 1984, 81, 3684–3690. [Google Scholar] [CrossRef]
  28. Mu, B.; Walton, K.S. Thermal analysis and heat capacity study of metal–organic frameworks. J. Phys. Chem. C 2011, 115, 22748–22754. [Google Scholar] [CrossRef]
Figure 1. 2 × 2 × 2 unit cells of MOF-5.
Figure 1. 2 × 2 × 2 unit cells of MOF-5.
Materials 11 00848 g001
Figure 2. Molecular configuration of R161.
Figure 2. Molecular configuration of R161.
Materials 11 00848 g002
Figure 3. The thermodynamic energy of MOF-5.
Figure 3. The thermodynamic energy of MOF-5.
Materials 11 00848 g003
Figure 4. Adsorption isotherms of R161 adsorbed in MOF-5.
Figure 4. Adsorption isotherms of R161 adsorbed in MOF-5.
Materials 11 00848 g004
Figure 5. The relationship between lnP and 1/T of R161 adsorption in MOF-5.
Figure 5. The relationship between lnP and 1/T of R161 adsorption in MOF-5.
Materials 11 00848 g005
Figure 6. The relative energy change of different working fluids as the temperature difference is increased at 5 MPa.
Figure 6. The relative energy change of different working fluids as the temperature difference is increased at 5 MPa.
Materials 11 00848 g006
Figure 7. The relationship between enthalpy difference and temperature difference at 5 MPa.
Figure 7. The relationship between enthalpy difference and temperature difference at 5 MPa.
Materials 11 00848 g007
Table 1. The desorption heat (kJ/mol) calculated by the Clausius–Clapeyron (C–C) equation and grand canonical Monte Carlo (GCMC).
Table 1. The desorption heat (kJ/mol) calculated by the Clausius–Clapeyron (C–C) equation and grand canonical Monte Carlo (GCMC).
C–CGCMC280KGCMC300KGCMC320KGCMC340KGCMC360KGCMC380KGCMC400KGCMC420K
41.7448.5848.3448.1047.6247.5147.1647.0146.85

Share and Cite

MDPI and ACS Style

Wang, Q.; Tang, S.; Li, L. Energy Storage Analysis of a Mixed R161/MOF-5 Nanoparticle Nanofluid Based on Molecular Simulations. Materials 2018, 11, 848. https://doi.org/10.3390/ma11050848

AMA Style

Wang Q, Tang S, Li L. Energy Storage Analysis of a Mixed R161/MOF-5 Nanoparticle Nanofluid Based on Molecular Simulations. Materials. 2018; 11(5):848. https://doi.org/10.3390/ma11050848

Chicago/Turabian Style

Wang, Qiang, Shengli Tang, and Leilei Li. 2018. "Energy Storage Analysis of a Mixed R161/MOF-5 Nanoparticle Nanofluid Based on Molecular Simulations" Materials 11, no. 5: 848. https://doi.org/10.3390/ma11050848

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