Next Article in Journal
The Optimized Conformation Dynamics of the KcsA Filter as a Probe for Lateral Membrane Effects: A First Principle Based Femto-Sec Resolution MD Study
Next Article in Special Issue
Role of Titanium Dioxide-Immobilized PES Beads in a Combined Water Treatment System of Tubular Alumina Microfiltration and PES Beads
Previous Article in Journal
Composite GO/Ceramic Membranes Prepared via Chemical Attachment: Characterisation and Gas Permeance Properties
Previous Article in Special Issue
Biorefinery of Tomato Leaves by Integrated Extraction and Membrane Processes to Obtain Fractions That Enhance Induced Resistance against Pseudomonas syringae Infection
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Experimental Performance of a Membrane Desorber with a H2O/LiCl Mixture for Absorption Chiller Applications

by
Jonathan Ibarra-Bahena
1,
Ulises Dehesa-Carrasco
1,
Yuridiana Rocio Galindo-Luna
2,
Iván Leonardo Medina-Caballero
3 and
Wilfrido Rivera
3,*
1
Subcoordinación de Conservación de Cuencas y Servicios Ambientales, Instituto Mexicano de Tecnología del Agua, Jiutepec 62550, Mexico
2
Departamento de Ingeniería de Procesos e Hidráulica, Universidad Autónoma Metropolitana-Iztapalapa, Mexico City 09340, Mexico
3
Instituto de Energías Renovables, Universidad Nacional Autónoma de México, Temixco 62580, Mexico
*
Author to whom correspondence should be addressed.
Membranes 2022, 12(12), 1184; https://doi.org/10.3390/membranes12121184
Submission received: 19 October 2022 / Revised: 11 November 2022 / Accepted: 17 November 2022 / Published: 24 November 2022
(This article belongs to the Special Issue Advances in Integrated Membrane Processes and Systems)

Abstract

:
For absorption cooling cycles using water as a refrigerant, H2O/LiCl mixtures are suitable for replacing conventional H2O/LiBr mixtures. In addition, membrane devices can be used to develop compact and lighter absorption systems, and they can operate with H2O/LiCl mixtures. The present paper describes an experimental evaluation of a membrane desorber/condenser operating at atmospheric pressure. Two operation modes were analyzed: continuous cycle operation and intermittent operation. For the first operation mode, the maximum desorption rate was 3.49 kg/h·m2, with a solution temperature of 90.3 °C and a condensation temperature of 25.1 °C. The lowest desorption rate value was 0.26 kg/h·m2, with a solution temperature of 75.4 °C and a condensation temperature of 40.1 °C. In the second mode, after three operating hours, the refrigerant fluid produced, per 1 m2 of membrane area, 7.7, 5.6, 4.3, and 2.2 kg, at solution temperatures of 90.3, 85.3, 80.4, and 75.4 °C, respectively. A one-dimension heat and mass transfer model is presented. The calculated values of desorption rate and outlet temperatures were compared with the experimental data; a square correlation coefficient of 0.9929 was reached for the desorption rate; meanwhile, for the outlet solution temperatures and the outlet cooling-water temperatures, a square correlation coefficient up to 0.9991 was achieved. The membrane desorber has the advantages of operating at atmospheric-pressure conditions, high condensation temperature, the ability to use different saline solution working mixtures, and different operation methods. These advantages can lead to new absorption systems.

1. Introduction

H2O/LiBr is the most used mixture in absorption chillers for air-conditioning applications. Water, as a refrigerant, shows convenient features: it has a high latent heat of evaporation; is a natural refrigerant; and is abundant, low cost, and harmless to the environment [1]. Absorption cycles using water as a refrigerant perform better than ammonia-based cycles, and a rectifier is unnecessary for them [2]. Among the water-based mixtures, H2O/LiCl mixtures are an interesting alternative showing some advantages compared with H2O/LiBr mixtures: triple-state point (solid, liquid, and vapor forms), long-term stability, and lower cost [3]. In addition, H2O/LiCl mixtures show better hygroscopic properties than other aqueous solutions, such as H2O/CaCl2 [4], although the smaller chemical potential of LiCl relative to LiBr leads to a larger absorber area [5]. On the other hand, Ahamad et al. [6], through a simulation analysis, found that at similar operating system temperatures, the solution concentrations in the desorber and absorber were considerably lower for the H2O/LiCl mixture than for the H2O/LiBr mixture, considerably reducing the risk of crystallization in these components. Additionally, they found that a slightly higher maximum Coefficient Of Performance (COP) could be reached with the H2O/LiCl mixture. Furthermore, the desorber temperature is lower with H2O/LiCl mixtures because the vapor pressure is higher than that in conventional H2O/LiBr mixtures [7]. The convenience of H2O/LiCl mixtures has already been demonstrated: higher COP [8,9], higher exergy efficiency [10,11], and a high potential to be integrated into thermal solar energy systems [12,13].
Additionally, using the correct working mixture allows the development of compact components for small-scale duties, such as single-family houses and automotive applications [14,15]. Light and small devices must replace conventional heat exchangers used as desorbers. In this regard, membrane modules constitute a novel technology that provides a high heat and mass transfer rate with a reduced size [16]. Studies on using aqueous lithium chloride mixtures with membrane components for absorption chiller systems are scarce: De Vega et al. [17] used microchannel membrane absorber simulation analysis to develop compact absorption chillers. The authors analyzed H2O/LiBr, NH3/LiNO3, and H2O/LiCl mixtures, and a membrane absorber with 13 microchannels 0.15 mm in height, 1.5 mm in width, and 50 mm in length was simulated. A membrane with 60 μm thickness, 80% porosity, and a pore diameter of 1 μm was assumed. According to the authors’ results, the highest ratio between cooling capacity and absorber volume was reached with the NH3/LiNO3 mixture. The H2O/LiCl mixture had the lowest desorption temperature; however, the COP was also the lowest. Asfand et al. [18] carried out a 2D simulation of a plate-and-frame membrane absorber that operates with H2O/(LiBr + LiI + LiNO3 + LiCl) and H2O/(LiNO3 + KNO3 + NaNO3) mixtures. The membrane absorber was 0.5 mm in width and 200 mm in length, and a membrane with 40 μm thickness, 85% porosity, and a pore diameter of 1 μm was considered. According to the authors’ results, the absorption rate increased by 25% with the H2O/(LiBr + LiI + LiNO3 + LiCl) mixture compared with H2O/LiBr under air-cooling thermal conditions. In addition, an absorption rate of 0.00523 kg/s × m2 was reached when the H2O/(LiNO3 + KNO3 + NaNO3) mixture was used in the membrane-based absorber of the third stage of a triple-effect absorption cooling cycle.
In addition, there are several studies on membrane absorption heat pumps using H2O/LiCl mixtures. These devices can be used for air-dehumidification, water-heating, and absorption cooling applications. Yang et al. [19] carried out a 3D simulation on a counter-flow parallel-plate membrane absorption heat pump (PMAHP) to recover low-temperature waste heat from used cooling water at 40 °C with a H2O/LiCl mixture. The authors concluded that the optimum operating conditions for the PMAHP were as follows: lower solution flow rate, higher water flow rate, lower inlet solution mass fraction, and higher inlet water temperature. Huang [20] analyzed the heat and mass transfer in a cross-flow hollow-fiber membrane absorption heat pump (HFMAHP) using a H2O-LiCl mixture. The solution temperature lift and the overall heat and mass transfer coefficients increased as the skin layer thickness increased. According to the author, the skin layer effect was higher than the porous layer. In another paper, Huang [21] analyzed a quasi-counter-flow parallel-plate membrane-absorption heat pump (QPMAHP) for fluid heating with a H2O/LiCl mixture. The author concluded that the sensible and latent heat transfer across the membranes was due to the large overall heat and mass transfer resistances. Nonetheless, the latent heat flux was dominant. The solution temperature lift and efficiency increased until it reached 9.1%, when the entrance and aspect ratios were equal to 0.1. Woods et al. [22] designed and tested a membrane absorption heat pump. The system included a membrane absorber with two hollow-fiber rows; the aqueous salt solution flowed inside one row, whereas the water flowed in the other. H2O/CaCl2 and H2O/LiCl mixtures were used. The authors mentioned that the cost of LiCl was higher than that of CaCl2. Performance comparison of a hollow-fiber membrane module with Oxyphan membranes and Accurel membranes was carried out assuming that they had the same membrane physical characteristics. The authors found that the mass transfer coefficient with Oxyphan membranes was 4.2 times higher than that with the second type. The water vapor mass flux was 2.1 times higher with the Accurel membranes compared with the Oxyphan ones.
The literature reviewed demonstrates the suitability of H2O/LiCl mixtures with membrane devices, especially for absorption cooling applications. The aim of the present study was to demonstrate the suitability of the H2O/LiCl mixture for the operation of a membrane desorber for absorption-cooling-system applications. This paper presents an experimental evaluation of a membrane desorber with a H2O/LiCl mixture at atmospheric-pressure conditions. The membrane device uses the air gap membrane distillation (AGMD) configuration. Two operation modes were analyzed: (i) continuous cycle operation and (ii) intermittent operation. Furthermore, the effects of the solution temperature and cooling-water temperature on the desorption rate were analyzed, and a one-dimension heat and mass transfer model was constructed.

2. Air Gap Membrane Distillation (AGMD) Configuration

Membrane distillation is a thermal separation process using a hydrophobic membrane as an interphase contactor between two fluids at different temperatures, where the separation occurs by the vapor mass transfer. At relatively low operating pressures, the hydrophobicity of the membrane prevents the liquid phase from wetting the membrane pores, and vapor is the only phase to cross the membrane. The temperature difference between the two sides of the membrane produces the driving force for the vapor mass transfer from the fluid at a higher temperature (hot side) to the one at a lower temperature (cold side) [23]. A schematic diagram of the AGMD configuration is shown in Figure 1. Some of the advantages of the AGMD configuration are as follows: the air gap serves as a thermal insulation layer and reduces the heat loss from the membrane; latent heat is recovered without an external heat exchanger; thermal efficiency is higher; it is suitable for pilot testing plants; and the AGMD modules can be scaled up and made with polymeric corrosion-resistant materials [24].

3. Methodology

The membrane desorber/condenser assessment was performed by varying the main system’s operating parameters, such as LiCl mass flow rate (LiCl), inlet solution temperature (TLiCl,in), and inlet cooling-water temperature (Tcw,in). Additionally, two operation modes were analyzed. In the first, a continuous operation absorption cycle was assumed, which means that the initial LiCl concentration was constant. In the second case, intermittent operation mode was assumed, meaning that the desorption/condensation and the evaporation/absorption processes in an absorption system separately occurred and that the LiCl concentration increased with operation time.

3.1. Experimental Setup

The experimental setup was mainly integrated by a membrane desorber/condenser unit, heating system, and cooling system, as shown in Figure 2. The membrane desorber/condenser unit functioned to separate part of the water contained in the LiCl solution as previously described. The unit comprised two Nylamid support plates, neoprene gaskets, a metallic mesh, an aluminum cooling plate, and a PTFE membrane, as shown in Figure 3.
The characteristics of the hydrophobic membrane are shown in Table 1.
The heating system comprises a heating bath with temperature control, a stainless-steel plate heat exchanger (PHE), a pump, and a Coriolis mass flowmeter. This system is in charge of supplying the necessary heat to the LiCl solution in the PHE before the solution enters the desorber unit.
The cooling system comprises a circulating cooling bath with temperature control and an integrated pump. This system provides the cooling to the desorber/condenser unit to condense the water in a vapor phase passing through the membrane.
A detailed description of the membrane desorber/condenser unit was already reported by Ibarra et al. [25]. Figure 4 shows a photograph of the experimental setup.
For the continuous operation mode, 16 experimental test runs were carried out; meanwhile, 4 test runs in intermittent operation mode were completed. The operating temperatures were selected on the basis of cooling absorption systems operating with renewable thermal energies [12,13] or low-grade heat sources [26]. The uncertainties of the measured variables and instruments used in the experimental test runs are shown in Table 2, and Table 3 and Table 4 report the experimental operating conditions for each operation mode.

3.2. Heat and Mass Transfer Model

A conceptual diagram of the desorber/condenser unit is shown in Figure 5. The hot saline solution is in direct contact with one side of the porous membrane. The refrigerant fluid evaporates at the entrance of the pores because of the hydrophobic nature of the membrane while a concentration profile is created: the salt concentration of the bulk fluid (Xbulk) increases until a higher concentration at the membrane interface (Xmem). The working fluid phase change occurs according to the vapor–liquid equilibrium of the H2O/LiCl mixture. The vapor permeates the membrane through the pores by the Poiseuille–Knudsen mechanism, crosses the air gap by diffusion, and condenses on the refrigerated plate. The heat transfer through the system is driven by the temperature difference between the hot saline solution channel and the cooling-water channel. The vapor mass transfer simultaneously occurs in the same direction with the heat transfer. The coupling of these phenomena is called the Soret effect.
In order to estimate the global performance parameters to scale and design a membrane desorber/condenser, a previously described one-dimension heat and mass transfer model was used [27] and adapted; it includes the H2O/LiCl mixture equilibrium conditions to describe the vapor water desorption. The following assumptions were made:
(1)
The desorber/condenser unit operates at steady state conditions.
(2)
Thermophysical properties are constant.
(3)
The heat and mass transfer processes occur in one dimension.
(4)
Natural convection is neglected in the air gap region.
(5)
Liquid–vapor equilibrium exists at the evaporation and condensation interfaces.
Energy conservation was considered in the different regions inside the experimental device [28,29].
The desorption heat flux (QD) transferred from the hot saline solution to the membrane is given by:
Q D = h 1 ( T 1 T 2 )
The heat transferred to the membrane:
h 1 ( T 1 T 2 ) = J w λ w + k m e m δ m e m ( T 2 T 3 )
Inside the air gap:
h 1 ( T 1 T 2 ) = J w λ w + k g a p δ g a p ( T 3 T 4 )
In the permeate film:
h 1 ( T 1 T 2 ) = h 4 ( T 4 T 5 )
At the condensing plate:
h 1 ( T 1 T 2 ) = k p δ p ( T 5 T 6 )
In the cooling stream:
h 1 ( T 1 T 2 ) = h 6 ( T 6 T 7 )
where h is the convective heat transfer coefficient, Jw is the desorption rate, λw is the latent heat of vaporization of water, k is the thermal conductivity, and δ refers to the thickness. Subscripts mem, gap, and p refer to the membrane, air gap, and condensation plate, respectively. Subscripts 1 to 7 are the numerical notation in Figure 5.
QD can be simplified in a mathematical expression as a function of the H2O/LiCl mixture temperature and the cooling-water stream temperature (T1 and T7, respectively). The equation system, integrated by Equations (2)–(6), was solved to dismiss the intermediate temperatures (T2 to T6). Thus, Equation (7) is proposed:
Q D = ψ ω ψ + k e q [ N A λ A + k e q ( T 1 T 7 ) ]
where the ψ, ω, keq, and heq are defined, respectively, as follows:
ψ = h 6 k p δ p h 6 + k p
ω = h e q k e q + h e q
k e q = k m e m k g a p δ g a p k m e m + δ m e m k g a p
h e q = h 1 h 4 h 1 + h 4
The convective heat transfer coefficients h1 and h6 were estimated from empirical correlations. In the H2O/LiCl mixture channel, the correlation proposed by Shah and London [30] was used to calculate the Nusselt number, which is valid for a fully developed flow in thin rectangular channels. On the other hand, in the cooling-water channel, the correlation reported by Khayet [31] was used. The correlation proposed by Bird et al. [32] was used to calculate the convective heat transfer coefficient for the condensate film (h4).
The temperatures for each zone inside the desorber/condenser device (T2 to T6) were calculated from Equations (1)–(6). However, T1 and T7, considered as the bulk temperatures of H2O/LiCl solution and cooling-water stream, respectively, were assumed to be the experimental inlet and outlet temperatures, respectively. Stream inlet temperatures (TLiCl,in and Tcw,in) are input parameters for the mathematical model, and the outlet temperatures were calculated by a numerical procedure, as a function of the heat flux in both streams:
T L i C l , o u t = T L i C l , i n Q D + Q l C p , L i C l   m ˙ L i C l
T c w , o u t = T c w , i n Q D + Q l C p , c w   m ˙ c w
where Cp is the heat capacity, is the mass flow rate, and Ql is the heat loss to the environment.
The ratio between the mass fraction at the liquid–membrane interface (Xm) and the mass fraction in the bulk flow (Xb) is defined as follows [22,33]:
X m e m X b = exp ( J w φ b l ρ L i C l )
where φbl is the convective mass transfer coefficient at the boundary layer, which was calculated by the correlation proposed by Woods et al. [33]; and ρLiCl is the H2O/LiCl solution density.
The driving force of the desorption rate (Jw) is the partial pressure difference between the interface de evaporation (p1) and condensate film (p4). However, the mass transfer flux is restricted by a global mass transfer coefficient (Kov):
J w = K o v ( p 1 p 4 )
The global mass transfer coefficient involves the boundary layer mass transfer coefficient at the liquid–membrane interface (Kbl), the mass transfer coefficient on the membrane (Kmem), and the mass transfer coefficient for the air gap (Kgap). It is expressed as:
K o v = [ 1 K b l + 1 K m e m + 1 K g a p ] 1
The mass transfer coefficient for the membrane is expressed as:
K m e m = M w R T m e m δ m e m [ ( 2 3 r p ε τ { 8 R T m e m π M w } 1 2 ) 1 + ( D v a p τ P p l m ) 1 ] 1
The mass transfer coefficient for the air gap is defined as:
K g a p = M w R T g a p D v a p δ g a p P p l m
The boundary layer mass transfer coefficient is defined as:
K b l = J w p b p b l
In these equations, Mw is the molecular mass of water; R is the universal gas constant; P and plm are the total operating pressure and the logarithmic-mean pressure of air, respectively; δmem, ε, and τ are the thickness, porosity, and tortuosity of the membrane, respectively; Dvap is the water vapor/air binary mass diffusion coefficient; rp is the membrane mean pore radius; δgap is the air gap thickness; and pb and pbl are the vapor partial pressure at the bulk layer and the boundary layer, respectively. The thermodynamic properties of the H2O/LiCl solution were calculated by using the correlations reported by Chaudhari and Patel [34] and Wimby and Berntsson [35].
In the experimental device, the channel length is small; therefore, the temperature difference between the inlet and the outlet is smaller than the temperature difference across the hot saline solution channel and cooling-water channel. Thus, a one-dimensional model assumption allows a general description of the system for engineering design purposes.

4. Results

4.1. Continuous Cycle Operation

The experimental desorption rate (Jw) was measured on the basis of the mass of distillate water produced by the membrane desorber after a defined time period, under different operating conditions. As demonstrated in previous reports with LiBr aqueous solutions [25,27], the desorption rate is mainly affected by the solution temperature rather than by the condensation temperature. For instance, at a solution temperature of 90.3 °C at the lowest condensation temperature, the desorption rate was 2.6 times higher than at 75.4 °C for the same condensation temperature. Meanwhile, the desorption rate at the highest condensation temperature was 9.4 times higher at the solution temperature of 90.3 °C than at 75.4 °C. As can be seen in Figure 6, the influence of the condenser temperature in the desorption rate was greater at the lowest solution temperature because at 75.4 °C the desorption rate was 5.2 times higher at the lowest condenser temperature than at the highest temperature; meanwhile, at the solution temperature of 90.3 °C, the desorption rate was just 1.4 times higher at the same condenser temperatures. The maximum desorption rate was 3.49 kg/h·m2 at a solution temperature of 90.3 °C and a condensation temperature of 25.1 °C. On the other hand, the lowest desorption rate value was 0.26 kg/h·m2 at a solution temperature of 75.4 °C and a condensation temperature of 40.1 °C. Because the vapor mass transfer driving force depends on the temperature difference between both sides of the membrane, as this temperature difference increases, the desorption rate also increases.
In Table 5, a comparison of the experimental results of this work with what was reported in the literature is presented. The desorption rates with the H2O/LiCl mixture are slightly lower than those reported with the H2O/LiBr mixture under similar operating conditions and with the same membrane configuration; however, the desorption rates are in a similar range to those reported with the H2O/LiBr mixture. These results show the suitability of using the H2O/LiCl mixture in a desorption process with membrane devices for absorption cooling systems.

4.2. Intermittent Operation

For this analysis, the refrigerant production (or water distilled) was defined as the quantity, in kilograms, of the refrigerant fluid produced per 1 m2 of membrane area. After three operating hours, at a constant condensation temperature of 30.1 °C, the refrigerant fluid produced was 7.7, 5.6, 4.3, and 2.2 kg, at solution temperatures of 90.3, 85.3, 80.4, and 75.4 °C, respectively. Figure 7 shows that the refrigerant produced decreased with operation time and showed a nonlinear dependence, particularly with high solution temperatures. As previously mentioned, the mass transfer driving force increases at higher temperature differences; thus, the refrigerant production was higher at the highest solution temperature. On the other hand, as a function of operation time, the amount of refrigerant contained in the solution decreased and the LiCl concentration increased, as is shown in Figure 8, so the refrigerant production decreased. Moreover, this behavior was observed in a previous study [46].

4.3. Model Validation

On the basis of the one-dimensional heat and mass transfer model, the desorption rate (JW), outlet solution temperatures (TLiCL,out), and outlet cooling-water temperatures (Tcw,out) were calculated. Figure 9 shows the comparison of theoretical and experimental refrigerant desorption rates. A square correlation coefficient (R2) of 0.9929 was reached, which means that the mathematical model provided values close to the experimental data. Figure 10 and Figure 11 show the theoretical TLiCl,out as a function of the experimental TLiCl,out and the theoretical Tcw,out as a function of the experimental Tcw,out, respectively. In both cases, an R2 value up to 0.9991 was achieved. This validation is essential for design heat networks to improve the heat efficiency of the absorption chillers.
One way to improve the mathematical model is to divide the channels into vertical control volumes, where each one has a uniform temperature. This means considering several vertical one-dimensional heat transfer paths in parallel.

5. Conclusions

A membrane desorber/condenser unit was experimentally evaluated, under laboratory conditions, with a H2O/LiCl solution. Two operation modes were analyzed: continuous cycle operation and intermittent operation. In the first case, four solution temperatures (90.3, 85.3, 80.4, and 75.4 °C) and four condensation temperatures (25.1, 30.1, 35.1, and 40.1 °C) were evaluated. According to the results, the desorption rate was principally affected by the solution temperature rather than by the condensation temperature; however, the condensation temperature effect on the desorption rate was notable at the lowest solution temperature. The maximum desorption rate was 3.49 kg/h·m2 at a solution temperature of 90.3 °C and condensation temperature of 25.1 °C.
On the other hand, the lowest desorption rate value was 0.26 kg/h·m2 at a solution temperature of 75.4 °C and condensation temperature of 40.1 °C. The same four solution temperatures as the first operation case were analyzed in the second mode at a constant condensation temperature of 30.1 °C. Mathematical model validation was carried out, and a square correlation coefficient of 0.9929 was reached for the desorption rate; meanwhile, for the outlet solution temperatures and the outlet cooling-water temperatures, a square correlation coefficient up to 0.9991 was achieved. After three operating hours, the refrigerant fluid produced, assuming a 1 m2 of membrane area, was 7.7, 5.6, 4.3, and 2.2 kg with solution temperatures of 90.3, 85.3, 80.4, and 75.4 °C, respectively. The refrigerant production increased as the solution temperature increased; however, the refrigerant production decreased with operation time because the LiCl concentration increased. On the basis of the experimental results, the H2O/LiCl mixture can be used to replace the conventional H2O/LiBr mixture for absorption cooling systems that integrate membrane devices.

Author Contributions

Conceptualization, J.I.-B., U.D.-C., Y.R.G.-L., I.L.M.-C. and W.R.; methodology, J.I.-B., U.D.-C., Y.R.G.-L., I.L.M.-C. and W.R.; writing—original draft preparation, J.I.-B., U.D.-C., Y.R.G.-L. and W.R.; writing—review and editing, J.I.-B., U.D.-C., Y.R.G.-L., I.L.M.-C. and W.R.; supervision, J.I.-B., U.D.-C., Y.R.G.-L., I.L.M.-C. and W.R.; project administration W.R.; funding acquisition, W.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the PAPIIT-UNAM IT100920 project.

Institutional Review Board Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding.

Acknowledgments

J.I.-B. and U.D.-C. express thanks to the FORDECYT 297486 and IMTA-RD2203.1 projects. The authors appreciate the technical support provided by Olga García Villa.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Arshi Banu, P.S.; Sudharsan, N.M. Review of water based vapour absorption cooling systems using thermodynamic analysis. Renew. Sustain. Energy Rev. 2017, 82, 3750–3761. [Google Scholar] [CrossRef]
  2. Jian, S.; Lin, F.; Zhang, S. A review of working fluids of absorption cycles. Renew. Sustain. Energy Rev. 2012, 16, 1899–1906. [Google Scholar]
  3. Altamirano, A.; Le Pierrès, N.; Stutz, B. Review of small-capacity single-stage continuous absorption systems operating on binary working fluids for cooling: Theoretical, experimental and commercial cycles. Int. J. Refrig. 2019, 106, 350–373. [Google Scholar] [CrossRef]
  4. Gommed, K.; Grossman, G.; Ziegler, F. Experimental investigation of a LiCI-water open absorption system for cooling and dehumidification. J. Sol. Energy Eng. 2004, 126, 710–715. [Google Scholar] [CrossRef]
  5. Kim, K.; Ameel, T.; Wood, B. Performance evaluations of LiCl and LiBr for ab- sorber design applications in the open-cycle absorption refrigeration system. J. Sol. Energy Eng. 1997, 119, 165–173. [Google Scholar] [CrossRef]
  6. Ahmad, T.; Azhar, M.; Sinha, M.K.; Meraj, M.; Mahbubul, I.M.; Ahmad, A. Energy analysis of lithium bromide-water and lithium chloride-water based single effect vapour absorption refrigeration system: A comparison study. Clean. Eng. Technol. 2022, 7, 100432. [Google Scholar] [CrossRef]
  7. Saravanan, R.; Maiya, M.P. Thermodynamic comparison of water-based working fluid combinations for a vapour absorption refrigeration system. Appl. Therm. Eng. 1998, 18, 553–568. [Google Scholar] [CrossRef]
  8. Horuz, I. A comparison between ammonia-water and water-lithium bromide solutions in vapor absorption refrigeration systems. Int. Commun. Heat Mass Transf. 1998, 25, 711–721. [Google Scholar] [CrossRef]
  9. Barragán Reyes, R.M.; Arellano Gómez, V.M.; García-Gutiérrez, A. Performance modelling of single and double absorption heat transformers. Curr. Appl. Phys. 2010, 10, S244–S248. [Google Scholar] [CrossRef]
  10. Patel, J.; Pandya, B.; Mudgal, A. Exergy based analysis of LiCl-H2O absorption cooling system. Energy Procedia 2017, 109, 261–269. [Google Scholar] [CrossRef]
  11. Bhowmick, A.; Kundu, B. Thermo-economic optimization and comparison study of LiBr-H2O and LiCl-H2O working pair in absorption cooling systems based on genetic algorithm. Int. J. Energy Res. 2021, 45, 3938–3954. [Google Scholar]
  12. Bellos, E.; Tzivanidis, C.; Antonopoulos, K.A. Exergetic and energetic comparison of LiCl-H2O and LiBr-H2O working pairs in a solar absorption cooling system. Energy Convers. Manag. 2016, 123, 453–461. [Google Scholar] [CrossRef]
  13. Bellos, E.; Tzivanidis, C.; Pavlovic, S.; Stefanovic, V. Thermodynamic investigation of LiCl-H2O working pair in a double effect absorption chiller driven by parabolic trough collectors. Therm. Sci. Eng. Prog. 2017, 3, 75–87. [Google Scholar] [CrossRef]
  14. Labus, J.M.; Bruno, J.C.; Coronas, A. Review on absorption technology with emphasis on small capacity absorption machines. Therm. Sci. 2013, 17, 739–762. [Google Scholar] [CrossRef]
  15. Altamirano, A.; Stutz, B.; Le Pierrès, N. Review of small-capacity single-stage continuous absorption systems operating on binary working fluids for cooling: Compact exchanger technologies. Int. J. Refrig. 2020, 114, 118–147. [Google Scholar] [CrossRef]
  16. Yu, D.; Chung, J.; Moghaddam, S. Parametric study of water vapor absorption into a constrained thin film of lithium bromide solution. Int. J. Heat Mass Transf. 2012, 55, 5687–5695. [Google Scholar] [CrossRef]
  17. De Vega, M.; Venegas, M.; García-Hernando, N. Modeling and performance analysis of an absorption chiller with a microchannel membrane-based absorber using LiBr-H2O, LiCl-H2O, and LiNO3-NH3. Int. J. Energy Res. 2018, 42, 3544–3558. [Google Scholar] [CrossRef]
  18. Asfand, F.; Stiriba, Y.; Bourouis, M. Performance evaluation of membrane-based absorbers employing H2O/(LiBr + LiI + LiNO3 + LiCl) and H2O/(LiNO3 + KNO3 + NaNO3) as working pairs in absorption cooling systems. Energy 2016, 115, 781–790. [Google Scholar] [CrossRef]
  19. Yang, M.; Low, E.; Lim Law, C.; Chen, J.-C.; Show, P.L.; Huang, S.-M. Heat and mass transfer in a counter flow parallel plate membrane-based absorption heat pump (PMAHP). Int. J. Therm. Sci. 2022, 171, 107227. [Google Scholar] [CrossRef]
  20. Huang, S.-M. Coupled heat and mass transfer in a cross-flow hollow fiber membrane absorption heat pump (HFMAHP). Appl. Therm. Eng. 2017, 111, 1119–1128. [Google Scholar] [CrossRef]
  21. Huang, S.M. Heat and mass transfer in a quasi-counter flow parallel-plate membrane-based absorption heat pump (QPMAHP). J. Membr. Sci. 2015, 496, 39–47. [Google Scholar] [CrossRef]
  22. Woods, J.; Pellegrino, J.; Kozubal, E.; Burch, J. Design and experimental characterization of a membrane-based absorption heat pump. J. Membr. Sci. 2011, 378, 85–94. [Google Scholar] [CrossRef]
  23. Alsaadi, A.S.; Ghaffour, N.; Li, J.-D.; Gray, S.; Francis, L.; Maab, H.; Amy, G.L. Modeling of air gap membrane distillation process: A theoretical and experimental study. J. Membr. Sci. 2013, 445, 53–65. [Google Scholar] [CrossRef] [Green Version]
  24. Shahu, V.T.; Thombre, S.B. Air gap membrane distillation: A review. J. Renew. Sustain. Energy 2019, 11, 045901. [Google Scholar] [CrossRef]
  25. Ibarra-Bahena, J.; Rivera, W.; Nanco-Mejía, S.D.; Romero, R.J.; Venegas-Reyes, E.; Dehesa-Carrasco, U. Experimental performance of a membrane desorber operating under simulated warm weather condensation temperatures. Membranes 2021, 11, 447. [Google Scholar] [CrossRef]
  26. She, X.; Yin, Y.; Xu, M.; Zhang, X. A novel low-grade heat-driven absorption refrigeration system with LiCl–H2O and LiBr–H2O working pairs. Int. J. Refrig. 2015, 58, 219–234. [Google Scholar] [CrossRef]
  27. Ibarra-Bahena, J.; Dehesa-Carrasco, U.; Romero, R.J.; Rivas-Herrera, B.; Rivera, W. Experimental assessment of a hydrophobic membrane-based desorber/condenser with H2O/LiBr mixture for absorption systems. Exp. Therm. Fluid Sci. 2017, 88, 145–159. [Google Scholar] [CrossRef]
  28. Izquierdo-Gil, M.A.; Garcia-Payo, M.C.; Fernandez-Pineda, C. Air gap membrane distillation of sucrose aqueous solutions. J. Membr. Sci. 1999, 155, 291–307. [Google Scholar] [CrossRef]
  29. Dehesa-Carrasco, U.; Pérez-Rábago, C.A.; Arancibia-Bulnes, C.A. Experimental evaluation and modeling of internal temperatures in an air gap membrane distillation unit. Desalination 2013, 326, 47–54. [Google Scholar] [CrossRef]
  30. Shah, R.K.; London, A.L. Laminar Flow Forced Convection in Ducts: A Source Book for Compact Heat Exchanger Analytical Data. Advances in Heat Transfer; Academic Press: New York, NY, USA, 1978; Volume 1. [Google Scholar]
  31. Khayet, M. Membranes and theoretical modeling of membrane distillation: A review. Adv. Colloid Interface Sci. 2011, 164, 56–88. [Google Scholar] [CrossRef] [PubMed]
  32. Bird, R.B.; Stewart, W.E.; Lightfoot, E.N. Transport Phenomena; Wiley: New York, NY, USA, 1976. [Google Scholar]
  33. Woods, J.; Pellegrino, J.; Kozubal, E.; Slayzak, S.; Burch, J. Modeling of a membrane-based absorption heat pump. J. Membr. Sci. 2009, 337, 113–124. [Google Scholar] [CrossRef]
  34. Chaudhari, S.K.; Patel, K.R. Thermodynamic Properties of Aqueous Solutions of Lithium Chloride. Phys. Chem. Liquids 2002, 40, 317–325. [Google Scholar] [CrossRef]
  35. Wimby, J.M.; Berntsson, T.S. Viscosity and density of aqueous solutions of LiBr, LiCl, ZnBr2, CaCl2, and LiNO3. 1. Single salt solutions. J. Chem. Eng. Data 1994, 39, 68–72. [Google Scholar] [CrossRef]
  36. Zhai, C.; Wu, W. Experimental evaluation on heat/mass transfer and pressure drop of a microchannel membrane-based desorber for compact and efficient H2O/LiBr absorption refrigeration. Int. J. Heat Mass Tranf. 2022, 195, 123198. [Google Scholar] [CrossRef]
  37. de Vega, M.; Venegas, M.; García-Hernando, N. Viability on the desorption and air condensation of water in a compact membrane-based microchannel desorber-condenser for cooling applications. Energy Convers. Manag. 2022, 267, 115919. [Google Scholar] [CrossRef]
  38. He, J.; Hirata, R.; Hihara, E.; Dang, C. Desorption characteristic of LiBr–H2O solution in hydrophobic hollow fiber membrane for absorption chiller. Appl. Therm. Eng. 2021, 195, 117164. [Google Scholar] [CrossRef]
  39. Venegas, M.; García-Hernando, N.; De Vega, M. Experimental evaluation of a membrane-based microchannel desorber operating at low desorption temperatures. Appl. Therm. Eng. 2020, 167, 114781. [Google Scholar] [CrossRef]
  40. Ibarra-Bahena, J.; Venegas-Reyes, E.; Galindo-Luna, Y.R.; Rivera, W.; Romero, R.J.; Rodríguez-Martínez, A.; Dehesa-Carrasco, U. Feasibility Analysis of a Membrane Desorber Powered by Thermal Solar Energy for Absorption Cooling Systems. Appl. Sci. 2020, 10, 1110. [Google Scholar] [CrossRef] [Green Version]
  41. Hong, S.J.; Hihara, E.; Dang, C. Analysis of adiabatic heat and mass transfer of microporous hydrophobic hollow fiber membrane-based generator in vapor absorption refrigeration system. J. Membr. Sci. 2018, 564, 415–427. [Google Scholar] [CrossRef]
  42. Isfahani, R.N.; Fazeli, A.; Bigham, S.; Moghaddam, S. Physics of lithium bromide (LiBr) solution dewatering through vapor venting membranes. Int. J. Multiph. Flow 2014, 58, 27–38. [Google Scholar] [CrossRef]
  43. Bigham, S.; Isfahani, R.N.; Moghaddam, S. Direct molecular diffusion and micro-mixing for rapid dewatering of LiBr solution. Appl. Therm. Eng. 2014, 64, 371–375. [Google Scholar] [CrossRef]
  44. Wang, Z.; Gu, Z.; Feng, S.; Li, Y. Application of vacuum membrane distillation to lithium bromide absorption refrigeration system. Int. J. Refrig. 2009, 32, 1587–1596. [Google Scholar] [CrossRef]
  45. Sudoh, M.; Takuwa, K.; Iizuka, H.; Nagamatsuy, K. Effects of thermal and concentration boundary layers on vapour permeation in membrane distillation of aqueous lithium bromide solution. J. Membr. Sci. 1997, 131, 1–7. [Google Scholar] [CrossRef]
  46. Ibarra-Bahena, J.; Rivera, W.; Romero, R.J.; Montiel-González, M.; Dehesa-Carrasco, U. Novel intermittent absorption cooling system based on membrane separation process. Appl. Therm. Eng. 2018, 136, 718–729. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the AGMD separation process [25].
Figure 1. Schematic diagram of the AGMD separation process [25].
Membranes 12 01184 g001
Figure 2. Schematic diagram of the experimental setup.
Figure 2. Schematic diagram of the experimental setup.
Membranes 12 01184 g002
Figure 3. Experimental membrane desorber/condenser unit [25].
Figure 3. Experimental membrane desorber/condenser unit [25].
Membranes 12 01184 g003
Figure 4. Experimental setup [25].
Figure 4. Experimental setup [25].
Membranes 12 01184 g004
Figure 5. Conceptual diagram of the AGMD process.
Figure 5. Conceptual diagram of the AGMD process.
Membranes 12 01184 g005
Figure 6. Desorption rate as a function of the solution and condensation temperatures.
Figure 6. Desorption rate as a function of the solution and condensation temperatures.
Membranes 12 01184 g006
Figure 7. Refrigerant production as a function of operation time.
Figure 7. Refrigerant production as a function of operation time.
Membranes 12 01184 g007
Figure 8. LiCl concentration as a function of operation time.
Figure 8. LiCl concentration as a function of operation time.
Membranes 12 01184 g008
Figure 9. Comparison between theoretical and experimental desorption rates.
Figure 9. Comparison between theoretical and experimental desorption rates.
Membranes 12 01184 g009
Figure 10. Comparison between theoretical and experimental outlet solution temperatures.
Figure 10. Comparison between theoretical and experimental outlet solution temperatures.
Membranes 12 01184 g010
Figure 11. Comparison between theoretical and experimental outlet cooling temperatures.
Figure 11. Comparison between theoretical and experimental outlet cooling temperatures.
Membranes 12 01184 g011
Table 1. Membrane characteristics.
Table 1. Membrane characteristics.
MaterialPTFE (Polytetrafluoroethylene)
Mean pore diameter (dp)0.22 μm
Porosity (ε)70%
Effective area (Amem)144 cm2
Thickness (δmem)175 μm
Table 2. Uncertainty of the measured variables.
Table 2. Uncertainty of the measured variables.
VariableSensor/InstrumentOperation RangeUncertainty
TemperatureRTD PT100−30 to 350 °C±0.1 °C
Volumetric flowVolumetric flowmeter0 to 7 L/min±5.0% f.s. *
Mass flowCoriolis mass flowmeter0 to 4.0 × 10−2 kg/s±0.1%
Distillate water weightElectronic balance0 to 600 g±0.01 g
* f.s., full scale.
Table 3. Experimental operating conditions in continuous operation mode.
Table 3. Experimental operating conditions in continuous operation mode.
ParameterValue
LiCl concentration (% kg salt/kg solution)41.05 ± 0.03
Cooling-water volumetric flow (L/min)2.0 ± 0.35
H2O/LiCl solution mass flow (kg/s)3.50 × 10−2 ± 1.83 × 10−5
H2O/LiCl solution temperature (°C)90.3 ± 0.1
85.3 ± 0.1
80.4 ± 0.1
75.4 ± 0.1
Cooling-water temperature (°C)40.1 ± 0.1
35.1 ± 0.1
30.1 ± 0.1
25.1 ± 0.1
Table 4. Experimental operating conditions in intermittent operation mode.
Table 4. Experimental operating conditions in intermittent operation mode.
ParameterValue
LiCl concentration (% kg salt/kg solution)40.98 ± 0.03
Cooling-water volumetric flow (L/min)2.0 ± 0.35
H2O/LiCl solution mass flow (kg/s)3.00 × 10−2 ± 4.90 × 10−5
H2O/LiCl solution temperature (°C)90.3 ± 0.1
85.3 ± 0.1
80.4 ± 0.1
75.4 ± 0.1
Cooling-water temperature (°C)30.1 ± 0.1
Table 5. Comparison of the experimental desorption rates for membrane desorbers reported in the literature.
Table 5. Comparison of the experimental desorption rates for membrane desorbers reported in the literature.
ReferenceConfigurationdp
(μm)
Working MixtureX
(% w/w) *
Tsol
(°C)
TCon
(°C)
sol
(kg/h)
Jw
(kg/h·m2)
[36]Flat sheet1.0H2O/LiBr50 to 6065 to 854015 to 5014 to 72
[37]Flat sheet0.45H2O/LiBr4370 to 9037 to 472.4 to 3.67.2 to 14.4
[38]Hollow fiber0.2 to 0.4H2O/LiBr58904010 to 300.41 to 2.2
[25]Flat sheet0.22H2O/LiBr49.680 to 9530 to 4590 to 1441.1 to 6.1
[39]Flat sheet0.45H2O/LiBr45.858 to 6025.70.5 to 1.75.8 to 15.1
[40]Flat sheet0.22H2O/LiBr49.875.2 to 95.314.4 to 25.490.01.5 to 5.7
[41]Hollow fiber0.16H2O/LiBr51 to 5865 to 83NA173 to 2690.4 to 3.4
[27]Flat sheet0.45H2O/LiBr45.7 to 58.774.4 to 95.915.6 to 20.058.7 to 90.00.3 to 9.7
[42]Flat sheet0.45H2O/LiBr48 to 5150 to 125NA0.75 to 3.250.0 to 37.8
[43]Flat sheet1.00H2O/LiBr4850 to 125NA2.50.0 to 34.2
[44]Hollow fiber0.16H2O/LiBr5065 to 88NA40 to 1200.3 to 2.0
[45]Flat sheet0.20H2O/LiBr35 to 5535 to 10015NA1.8 to 18
This workFlat sheet0.22H2O/LiCl4175 to 9025 to 401260.26 to 3.49
* kg salt/kg solution.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ibarra-Bahena, J.; Dehesa-Carrasco, U.; Galindo-Luna, Y.R.; Medina-Caballero, I.L.; Rivera, W. Experimental Performance of a Membrane Desorber with a H2O/LiCl Mixture for Absorption Chiller Applications. Membranes 2022, 12, 1184. https://doi.org/10.3390/membranes12121184

AMA Style

Ibarra-Bahena J, Dehesa-Carrasco U, Galindo-Luna YR, Medina-Caballero IL, Rivera W. Experimental Performance of a Membrane Desorber with a H2O/LiCl Mixture for Absorption Chiller Applications. Membranes. 2022; 12(12):1184. https://doi.org/10.3390/membranes12121184

Chicago/Turabian Style

Ibarra-Bahena, Jonathan, Ulises Dehesa-Carrasco, Yuridiana Rocio Galindo-Luna, Iván Leonardo Medina-Caballero, and Wilfrido Rivera. 2022. "Experimental Performance of a Membrane Desorber with a H2O/LiCl Mixture for Absorption Chiller Applications" Membranes 12, no. 12: 1184. https://doi.org/10.3390/membranes12121184

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