Next Article in Journal
Thermodynamic vs. Kinetic Basis for Polymorph Selection
Next Article in Special Issue
Modeling of the Free Radical Copolymerization Kinetics of n-Butyl Acrylate, Methyl Methacrylate and 2-Ethylhexyl Acrylate Using PREDICI®
Previous Article in Journal
Optimization of Ultrasound-Assisted Extraction on Antioxidative Activity of Malus toringoides Using Response Surface Methodology
Previous Article in Special Issue
Internal Model Control for Rank-Deficient System with Time Delays Based on Damped Pseudo-Inverse
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Mathematical Modeling of the Reverse Osmosis Concentration Process of a Glucose Solution

Biomass Energy Technology Research Center, Biogas Institute of Ministry of Agriculture, Chengdu 610041, China
*
Author to whom correspondence should be addressed.
Processes 2019, 7(5), 271; https://doi.org/10.3390/pr7050271
Submission received: 21 March 2019 / Revised: 23 April 2019 / Accepted: 27 April 2019 / Published: 8 May 2019
(This article belongs to the Special Issue Modeling, Simulation and Control of Chemical Processes)

Abstract

:
A mathematical modeling of glucose–water separation through a reverse osmosis (RO) membrane was developed to research the membrane’s performance during the mass transfer process. The model was developed by coupling the concentration–polarization (CP) model, which uses one-dimensional flow assumption, with the irreversible thermodynamic Spiegler–Kedem model. A nonlinear parameter estimation technique was used to determine the model parameters Lp (hydraulic permeability constant), σ (reflection coefficient), and Bs (solute transport coefficient). Experimental data were obtained from the treatment of a pre-treated glucose solution using a laboratory-scale RO system, and studies on the validation of the model using experimental results are presented. The calculated results are consistent with the experimental data. The proposed model describes the RO membrane concentration process and deduces the expression of k (mass transfer coefficient in the CP layer). The verification shows that the expression of k well-describes the reverse osmosis mass transfer of a glucose solution.

1. Introduction

In the last few decades, in China, with the development of industry and the increase in pollution, water shortage has been getting worse. Wastewater recycling can effectively alleviate this situation and contribute to the sustainability of development. Considering the relatively efficient and simple process, membrane technology is a promising method for water treatment [1,2,3,4]. Additionally, it is reported that reverse osmosis (RO) technology is a relatively appropriate technique for manipulating sugar concentration [5]. In recent years, RO has gained widespread interest and has become a widely adopted modern tool for industrial applications and research laboratories. This has resulted in a great demand for RO membranes, making it necessary to reduce the cost of the RO system. According to the literature reported [6], the development of a proper mathematical model that adequately describes the performance of the RO process is crucial for the optimum design of an RO membrane which can make the RO system more efficient, thereby reducing the overall cost.
A number of mechanistic and mathematical models have been proposed to describe the mass transfer and hydrodynamic permeability in RO. In the RO separation processes, the mass transfer is mainly governed by the inside-membrane mass transfer and the outside-membrane mass transfer. The inside-membrane mass transfer is explained by the solution–diffusion model, proposed by Lonsdale et al. [7,8] and the preferential sorption–capillary flow model, proposed by Sourirajan et al. [8,9]. The main subject of investigation in the outside-membrane mass transfer is the mass transfer process near the membrane, where concentration–polarization (CP) significantly affects the mass transfer process. Hence, the study of outside-membrane mass transfer focuses on the modeling of CP; it is necessary to model the CP phenomenon to predict the RO separation process.
Various attempts were made to understand and describe the CP phenomenon by a theoretical analysis, and the main models were based on film theory, such as the Kimura–Sourirajan model [10,11,12,13,14] and the Spiegler–Kedem model [15,16,17,18,19], which are both related to the mass transfer coefficient. In other literature, a model based on the solution–diffusion mass transport theory, CP, and pressure-dependent dynamic membrane resistance was developed to predict the performance of RO for mixed salt and dye solutions [20,21]. Pastagia et al. [22] developed an unsteady-state mass transfer model for a binary solute system composed of reactive black dye and a reactive red dye. Moreover, Jamal [23] developed a model without CP for the prediction of feed concentration, permeate concentration, rejection, and flux as a function of the operating time. Overall, these models proved to be useful and applicable in the RO separation process.
The objective of this work was to research the mass transfer during the RO concentration process. To fulfill this requirement, a mathematical model based on glucose–water separation through an RO membrane was developed. Considering the actual structure of the RO membrane, the irreversible thermodynamic Spiegler–Kedem model is appropriate for describing the inside-membrane mass transfer, while a CP model based on the one-dimensional flow assumption was selected to explain the outside-membrane mass transfer. The mathematical model proposed in this study couples the CP model with the irreversible thermodynamic Spiegler–Kedem model. Experimental data (glucose concentration experiment) were obtained from the treatment of a pretreated glucose solution using a laboratory-scale RO system, and studies on the validation of the model using experimental results are presented The calculated results are consistent with the experimental data. Also, the expression of k is deduced.

2. Theory

In the RO separation process, the mass transfer occurs mainly in the feed boundary layer (CP layer) and inside the membrane. The general process of mass transfer is shown schematically in Figure 1. In order to build a mass transfer model, a one-dimensional flow is assumed to be valid for the transport of solvent and solute through the membrane. In addition, the CP layer is assumed to be fully developed. On this basis, the mass transfer equation is expressed as
J w C p = J w C D d c d x
where Jw is the solvent (pure water) flux, Cp is the permeate solution concentration, C is the solute concentration in the CP layer, and D is the diffusion coefficient. The Equation (1) is transformed into the following equation
D d c d x = J w C J w C p
According to Figure 1, x ranges from 0 to δcp, and C ranges from Cf to Cδ,1. Integrating Equation (2) gives the following equation
C δ , 1 C p C f C p = e ( J w δ c p D ) = e ( J w k )
where Cδ,1 is the solute concentration at the membrane surface (feed side, as shown in Figure 1), Cf is the feed solution concentration, δcp is the CP layer thickness, and k is the mass transfer coefficient in the CP layer and is defined as
k = D δ c p
According to Figure 1, the solute flux Js is expressed by the following equation
J s = B s ( C δ , 1 C δ , 2 ) = B s ( C δ , 1 C p ) = J w C p
where Bs is the solute transport coefficient.
For the reverse osmosis process, the total flux of the solvent and solute (Jv), which represents the volumetric flux on the permeate side of the RO membrane and can reflect the concentration capacity of the RO membrane, is expressed by the following equation
J v = J w + J s J w
According to the principle of mass conservation, we get the following equations
Q f C f = Q b C b + Q p C p
Q f = Q b + Q p
where Qf, Qb,, and Qp are the feed solution flow, retentate flow, and permeate flow, respectively. Cb is the retentate solution concentration, and the relationship between Qp and Jw is shown in the following equation
J v = Q p S
where S is the effective RO membrane area.
The recovery (y), which represents the water production capacity, is defined as the fraction of the feed flow which passes through the membrane. The higher the recovery, the stronger the water production capacity of the RO system, and the stronger the concentration capacity of the RO system. The value of y is calculated by the following equation
y = Q p Q f × 100 %
To describe the mass transfer process of the RO membrane, we examined the parameters of the glucose solution and the microstructure of the RO membrane (PA2-4040, HYDECANME, HYDRANAUTICS, Oceanside, CA, USA). The effective thickness (calculated by deducting the thickness of the support layer) of the RO membrane was 0.2 μm. In contrast, the diameter of the glucose molecule (<1 × 10−3 μm) and the diameter of the water molecule (≈4 × 10−4 μm) are much smaller than the RO membrane thickness, and the membrane pore of the RO membrane was generally less than 1 nm (roughly the same as the diameter of the glucose molecule). Therefore, we believe that the solution interacted with the RO membrane during the membrane separation of the glucose solution, and the solution-diffusion model was discarded. In this work, the irreversible thermodynamic Spiegler–Kedem model [24,25,26,27] was assumed to be appropriate for explaining the separation performance of solute through the membrane. Therefore, Jv is expressed by the following equation
J v = L p ( Δ p σ Δ π )
where Lp is the hydraulic permeability constant, △p is the transmembrane pressure, △π is the difference in the osmotic pressure across the membrane, and σ is the reflection coefficient. The reflection coefficient represents the solute separation capability of a membrane, where σ = 0 means no separation, and σ = 1 means complete separation (100% separation). For permeable membranes, 0 < σ < 1. △π is calculated using the following equation
Δ π = R T ( C δ , 1 C δ , 2 ) = R T ( C δ , 1 C p )
where R is the gas law constant, T is the temperature, and Cδ,2 is the solute concentration at the membrane surface (permeate side, as shown in Figure 1).
The observed membrane rejection fraction Ro is given by
R o = C f C p C f
and the real rejection fraction Rr is given by
R r = C δ , 1 C δ , 2 C δ , 1 = C δ , 1 C p C δ , 1
According to the Spiegler–Kedem equation, the real rejection fraction Rr is expressed as
R r = σ ( 1 F ) 1 σ F
F is defined as
F = exp [ J v ( 1 σ ) / B s ]
where Bs is the solute transport coefficient.
Rewriting Equation (15), we get
J v = B s 1 σ ln σ ( 1 R r ) σ R r
Substituting Equations (13) and (14) in Equation (3), we get
ln ( 1 R o R o ) = J v k + ln ( 1 R r R r )
Rewriting Equation (18), we get
k = J v / ln R r ( 1 R o ) R o ( 1 R r )
Substituting Equations (3) and (12) in Equation (11), we get
J v = L p ( Δ p σ Δ π ) = L p [ Δ p σ R T ( C f C p ) e ( J v k ) ]
Substituting Equation (19) in Equation (20), we get
J v = L p [ Δ p σ R T ( C f C p ) R r ( 1 R o ) R o ( 1 R r ) ]

3. Experimental

3.1. Membrane and Module

A polyamide composite membrane (PA2-4040, HYDECANME, USA) was chosen as the reverse osmosis experiment membrane material in the present work. The membrane characteristics are shown in Table 1.
The experimental device (Figure 2) mainly consisted of a cooling and heating circulation tank to change the temperature of the model solution, a pump for feeding the glucose solution, a valve (check valve) to control the solution flux, the reverse osmosis membrane module (including pressure gauges and a flow meter), and a device for sampling the solution.

3.2. Preparation of Model Solutions and Concentration Experiment

In this study, glucose anhydrous (AR) was purchased from the Kelong Chemical Reagent Factory, Chengdu, China. Model solutions (24.8–166.8 mol·m−3) were prepared using ultrapure water, by adding suitable amounts of the analytes (calculated).
The glucose model solutions were selected to carry out reverse osmosis concentration experiments using the reverse osmosis system (Figure 2). Model solutions were transferred into the membrane module by a feed pump. This was followed by the outflow of concentrate and permeate from the reverse osmosis membrane module and the reflow into the circulation tank. During this procedure, the concentrate was sampled by a sampling device and subsequently analyzed to evaluate the separation effect of the reverse osmosis system. The concentrations of the glucose solutions (amount of sample for each test: 20 µL) were quantified using a high-performance liquid chromatography system (HPLC, Agilent LC1200, Agilent, Santa Clara, CA, USA) equipped with a differential refraction detector (RID) and an Aminex HPX-87H column (Bio-Rad Co., Hercules, CA, USA). Experimental data were measured three times to obtain an average value under each experimental condition.

4. Results and Discussion

4.1. Influence of Experimental Parameters of the Reverse Osmosis System

The pump used was a vertical multistage centrifugal pump (CDLF1, 50HZ, CNP, Hangzhou, China), and in the experimental design, one of the valves was installed at the outlet of the RO membrane (valve 3, Figure 2). The feed flow and transmembrane pressure in the system were changed by adjusting the valve 3. The relationship between the feed solution and transmembrane pressure is shown in Figure 3. According to Figure 3, the feed flow rate decreased when the transmembrane pressure increased, and the two parameters were roughly linear.
Figure 4 reflects the effect of transmembrane pressure on the RO process. According to Figure 4, within the range of experimental parameters (310 kPa < △p < 617 kPa; 36 mol·m−3 < Cf < 140 mol·m−3), when other experimental conditions are constant, Jv, y, Ro, and Cb increase with an increase in transmembrane pressure. Jv and △p are approximately linear, which is consistent with the irreversible thermodynamic Spiegler–Kedem model selected in this paper (Equation (11)). As the △p increased, y increased slowly, followed by an increase in growth rate, and the Cb maintained approximately the same trend. According to Equation (10), the recovery formula is a ratio. Combined with Figure 3 and Jv trend over △p. In the formula of y, the numerator increases linearly with △p (consistent with the trend of Jv, Equation (9)), and the denominator decreases approximately linearly with △p. Therefore, the overall ratio should increase, and the rate of the value increase should be gradual, which is consistent with the trend of y in Figure 4. According to Figure 4b, with the increase of △p, Ro is slightly improved, although the trend is not obvious. According to Figure 3, Qf decreased with increases of △p. This means that a unit volume of feed solution can acquire more pressure to facilitate the RO process. The macroscopic performance of this process was an increase in Ro and Jv.
Figure 5 shows the effect of Cf on the RO process. It can be seen in Figure 5 that an increase in Cf led to an increase in the Cb and decreases in Jv and y. The trend of Ro was not significant (slightly lower trend). In the RO process, the CP phenomenon occurs near the boundary layer of the membrane. When Qf is constant, the increase in Cf leads to an enhanced CP phenomenon, which can result in a decrease in Jv, followed by a decrease in y. On the other hand, the enhancement of the CP phenomenon also leads to a slightly lower trend for Ro. The trend of Cb increasing with Cf was predictable and imaginable.

4.2. Model Validation

Some mathematical formulas for the mass transfer in the RO process were given in the previous section (Section 2). In this section, model validation was carried out to verify that the calculated values of these mathematical formulas were consistent with the experimental data. In order to verify the models, we first determined some parameter values in the model. The non-linear parameter estimation technique of Levenberg–Marquardt, combined with the Gauss–Newton algorithm [17,20,28,29,30,31], was used in this study to estimate the parameters, and the calculated values of Bs and σ were 0.85 × 10−8 m·s−1 and 0.9981, respectively. We also calculated the Lp and Cδ,1; the results are shown in Table 2. From Table 2 and Equation (14), we further calculated the real rejection fraction Rr. As shown in Table 2, Lp remained basically the same.
In the RO experiment, the feed solution flow Qf and the permeate flow Qp were obtained according to the flow meter. Then, the solvent (pure water) flux Jv was obtained (experimental value) based on Equation (9). Combined with Equation (21) and Table 2, we calculated the Jv (theoretical value). The comparison of the two Jv values is shown in Figure 6. Meanwhile, we calculated the relative error between each group of two Jv values, and the maximum value of the relative error was 12.81%.

4.3. Estimation of the Mass Transfer Coefficient, k

The mass transfer coefficient (k) represents the transfer rate of the solute in the CP layer. The k value can reflect the strength of the concentration polarization, and with a known k value, the salt diffusion coefficient and boundary layer thickness can be determined using a dimensionless analysis. The RO membrane (PA2-4040, HYDECANME, USA) used in this work was a spiral wound membrane. The mass transfer coefficient (k) was calculated according to the following empirical formula [12]
k = α R e β S c γ D d h
where α, β, and γ are constants that vary for different physical situations, which are determined experimentally, and Re is the Reynolds number. Experiments indicate that the flow rate of the fluid, the fluid channel size, and the physical properties of the fluid (density and viscosity) affect the type of flow. These main influencing factors can be combined into a dimensionless parameter (Re), the value of which can reflect the type of flow. Sc is the Schmidt number, and dh is the hydraulic diameter of the channel. Re and Sc are defined as
R e = ρ u d h μ , S c = μ ρ D
where u, ρ, and μ are the velocity, the density, and the viscosity of the fluid, respectively.
For a spiral wound membrane, dh is approximated by the following formula
d h 2 h
where h is the rectangular feed channel thickness (h = 0.7112 mm, HYDECANME).
According to Equation (22) and the experimental work of this article, When Qf changes during the RO process (since the Cf does not change, Sc and D/dh are constant) only Re changes. Therefore, Equation (22) is rewritten by the following equation
k = α 1 R e β
where α1 represents a combination constant.
Combining Table 2 with Equation (19), we can calculate the value of k and Re under different feed solution flow conditions (as shown in Table 3), then we performed a power function fitting according to Table 3 and Equation (25) and β = 0.9243. Equation (22) is rewritten by the following equation
k = α R e 0.9243 S c γ D d h
Also, when Cf changed during the RO process, ρ, μ, and D of the solution were slightly changed as a result of the change in the concentration of the solution, leading to a slight change in Re. Since the change in Re was slight, we ignored it (Re ≈ 95) in this paper. Equation (22) is rewritten by the following equation
k = α 2 S c γ
where α2 represents a variable value.
We calculated the value of k and Sc under different feed solution concentration conditions, as shown in Table 4, and the value of D was estimated from the literature [30]). According to Table 4, using the same method as above, the value of γ was obtained (γ = 0.3495). Equation (22) was rewritten by the following equation
k = α R e 0.9243 S c 0.3495 D d h
Finally, we calculated the value of α according to Table 4 and modified the value of α according to Table 3, α = 0.031, and obtained the final expression of the mass transfer coefficient
k = 0.031 R e 0.9243 S c 0.3495 D d h
In this paper, according to Equation (29), the Reynolds number was the main factor affecting the mass transfer coefficient. Combining Table 3 and Equation (29), we plotted the experimental and theoretical values of k, as shown in Figure 7. We also calculated the relative error between each group of the two k values; the maximum value of the relative error was 12.2%.

5. Conclusions

A mathematical modeling of glucose–water separation through an RO membrane was developed in this article. The model coupled the CP model, which used a one-dimensional flow assumption, with the irreversible thermodynamic Spiegler–Kedem model. The non-linear parameter estimation technique was used to determine the model parameters, such as Lp, σ, and Bs. Research on the validation of the model with experimental results was presented, and experimental data were obtained from the treatment of a pre-treated glucose solution using a laboratory-scale RO system. The comparison of the results showed that the calculated values were consistent with the experimental data. According to the calculated results, the maximum relative error between the two values was 12.81%. The expression of k was also deduced. The model well-describes the membrane performance during the mass transfer process, and the verification showed that the expression of k well-describes the reverse osmosis mass transfer of glucose solution.

Author Contributions

The concentrations of the glucose solutions were quantified by co-author H.Q. using HPLC, and the rest was completed independently by C.C.

Funding

This research was funded by Central Public-interest Scientific Institution Basal Research Fund [No. 12018206030202210_04]. And the APC was funded by Central Public-interest Scientific Institution Basal Research Fund [12016206030201701].

Acknowledgments

The authors are grateful to the Central Public-interest Scientific Institution Basal Research Fund [No. 12018206030202210_04] and Central Public-interest Scientific Institution Basal Research Fund [12016206030201701].

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Atab, M.S.; Smallbone, A.J.; Roskilly, A.P. A hybrid reverse osmosis/adsorption desalination plant for irrigation and drinking water. Desalination 2018, 444, 44–52. [Google Scholar] [CrossRef]
  2. Chung, T.S.; Zhang, S.; Wang, K.Y.; Su, J.; Ling, M.M. Forward osmosis processes: Yesterday, today and tomorrow. Desalination 2012, 287, 78–81. [Google Scholar] [CrossRef]
  3. Ray, S.S.; Chen, S.S.; Sangeetha, D.; Chang, H.M.; Thanh, C.N.D.; Le, Q.H.; Ku, H.M. Developments in forward osmosis and membrane distillation for desalination of waters. Environ. Chem. Lett. 2018, 16, 1247–1265. [Google Scholar] [CrossRef]
  4. Ray, S.S.; Chen, S.S.; Nguyen, N.C.; Nguyen, H.T.; Dan, N.P.; Thanh, B.X. Exploration of polyelectrolyte incorporated with Triton-X 114 surfactant based osmotic agent for forward osmosis desalination. J. Environ. Manag. 2018, 209, 346–353. [Google Scholar] [CrossRef] [PubMed]
  5. Zhou, F.; Wang, C.; Wei, J. Separation of acetic acid from monosaccharides by NF and RO membranes: Performance comparison. J. Membr. Sci. 2013, 429, 243–251. [Google Scholar] [CrossRef]
  6. Sundaramoorthy, S.; Srinivasan, G.; Murthy, D.V.R. An analytical model for spiral wound reverse osmosis membrane modules: Part I—Model development and parameter estimation. Desalination 2011, 280, 403–411. [Google Scholar] [CrossRef]
  7. Lonsdale, H.K.; Merten, U.; Riley, R.L. Transport properties of cellulose acetate osmotic membranes. J. Appl. Polym. Sci. 1965, 9, 1341–1362. [Google Scholar] [CrossRef]
  8. Merdaw, A.A.; Sharif, A.O.; Derwish, G.A.W. Water permeability in polymeric membranes, Part II. Desalination 2010, 257, 184–194. [Google Scholar] [CrossRef]
  9. Soltanieh, M.; GILL’, W.N. Review of reverse osmosis membranes and transport models. Chem. Eng. Commun. 1981, 12, 279–363. [Google Scholar] [CrossRef]
  10. Goosen, M.F.A.; Sablani, S.; Cin, M.D.; Wilf, M. Effect of cyclic changes in temperature and pressure on permeation properties of composite polyamide seawater reverse osmosis membranes. Sep. Sci. Technol. 2010, 46, 14–26. [Google Scholar] [CrossRef]
  11. Ghernaout, D. Reverse osmosis process membranes modeling—A historical overview. J. Civ. Constr. Environ. Eng. Civ. 2017, 2, 112–122. [Google Scholar]
  12. Qiu, T.Y.; Davies, P.A. Concentration polarization model of spiral-wound membrane modules with application to batch-mode RO desalination of brackish water. Desalination 2015, 368, 36–47. [Google Scholar] [CrossRef]
  13. Kim, D.Y.; Lee, M.H.; Gu, B.; Kim, J.H.; Lee, S.; Yang, D.R. Modeling of solute transport in multi-component solution for reverse osmosis membranes. Desalin. Water Treat. 2010, 15, 20–28. [Google Scholar] [CrossRef]
  14. Khanarmuei, M.; Ahmadisedigh, H.; Ebrahimi, I.; Gosselin, L.; Mokhtari, H. Comparative design of plug and recirculation RO systems; thermoeconomic: Case study. Energy 2017, 121, 205–219. [Google Scholar] [CrossRef]
  15. Spiegler, K.S.; Kedem, O. Thermodynamics of hyperfiltration (reverse osmosis): Criteria for efficient membranes. Desalination 1966, 1, 311–326. [Google Scholar] [CrossRef]
  16. Jain, S.; Gupta, S.K. Analysis of modified surface force pore flow model with concentration polarization and comparison with Spiegler–Kedem model in reverse osmosis systems. J. Membr. Sci. 2004, 232, 45–62. [Google Scholar] [CrossRef]
  17. Murthy, Z.V.P.; Chaudhari, L.B. Separation of binary heavy metals from aqueous solutions by nanofiltration and characterization of the membrane using Spiegler–Kedem model. Chem. Eng. J. 2009, 150, 181–187. [Google Scholar] [CrossRef]
  18. Attarde, D.; Jain, M.; Gupta, S.K. Modeling of a forward osmosis and a pressure-retarded osmosis spiral wound module using the Spiegler-Kedem model and experimental validation. Sep. Purif. Technol. 2016, 164, 182–197. [Google Scholar] [CrossRef]
  19. Hidalgo, A.M.; León, G.; Gómez, M.; Murcia, M.D.; Gómez, E.; Gómez, J.L. Application of the Spiegler–Kedem–Kachalsky model to the removal of 4-chlorophenol by different nanofiltration membranes. Desalination 2013, 315, 70–75. [Google Scholar] [CrossRef]
  20. Ahmad, A.L.; Chong, M.F.; Bhatia, S. Mathematical modeling of multiple solutes system for reverse osmosis process in palm oil mill effluent (POME) treatment. Chem. Eng. J. 2007, 132, 183–193. [Google Scholar] [CrossRef]
  21. Al-Bastaki, N. Removal of methyl orange dye and Na2SO4 salt from synthetic waste water using reverse osmosis. Chem. Eng. Process. Process Intensif. 2004, 43, 1561–1567. [Google Scholar] [CrossRef]
  22. Pastagia, K.M.; Chakraborty, S.; DasGupta, S.; Basu, J.K.; De, S. Prediction of permeate flux and concentration of two-component dye mixture in batch nanofiltration. J. Membr. Sci. 2003, 218, 195–210. [Google Scholar] [CrossRef]
  23. Jamal, K.; Khan, M.A.; Kamil, M. Mathematical modeling of reverse osmosis systems. Desalination 2004, 160, 29–42. [Google Scholar] [CrossRef]
  24. Ślȩzak, A. Irreversible thermodynamic model equations of the transport across a horizontally mounted membrane. Biophys. Chem. 1989, 34, 91–102. [Google Scholar] [CrossRef]
  25. Van Gauwbergen, D.; Baeyens, J. Modelling reverse osmosis by irreversible thermodynamics. Sep. Purif. Technol. 1998, 13, 117–128. [Google Scholar] [CrossRef]
  26. Sablani, S.S.; Goosen, M.F.A.; Al-Belushi, R.; Wilf, M. Concentration polarization in ultrafiltration and reverse osmosis: A critical review. Desalination 2001, 141, 269–289. [Google Scholar] [CrossRef]
  27. Gupta, V.K.; Hwang, S.T.; Krantz, W.B.; Greenberg, A.R. Characterization of nanofiltration and reverse osmosis membrane performance for aqueous salt solutions using irreversible thermodynamics. Desalination 2007, 208, 1–18. [Google Scholar] [CrossRef]
  28. Marquardt, D.W. An algorithm for least-squares estimation of nonlinear parameters. J. Soc. Ind. Appl. Math. 1963, 11, 431–441. [Google Scholar] [CrossRef]
  29. Mendes, P.; Kell, D. Non-linear optimization of biochemical pathways: Applications to metabolic engineering and parameter estimation. Bioinformatics 1998, 14, 869–883. [Google Scholar] [CrossRef]
  30. Dong, J.; Lu, K.; Xue, J.; Dai, S.; Zhai, R.; Pan, W. Accelerated nonrigid image registration using improved Levenberg–Marquardt method. Inf. Sci. 2018, 423, 66–79. [Google Scholar] [CrossRef]
  31. Zhao, C.; Ma, P.; Zhu, C.; Chen, M. Measurement and correlation on diffusion coefficients of aqueous glucose solutions. J. Chem. Ind. Eng. 2005, 56, 1–5. [Google Scholar]
Figure 1. Schematic diagram of mass transfer in the reverse osmosis (RO) membrane separation process. CP: concentration–polarization.
Figure 1. Schematic diagram of mass transfer in the reverse osmosis (RO) membrane separation process. CP: concentration–polarization.
Processes 07 00271 g001
Figure 2. Schematic illustration of the reverse osmosis system.
Figure 2. Schematic illustration of the reverse osmosis system.
Processes 07 00271 g002
Figure 3. Relation between the feed solution flow and the transmembrane pressure.
Figure 3. Relation between the feed solution flow and the transmembrane pressure.
Processes 07 00271 g003
Figure 4. Influence of transmembrane pressure on RO process. (a) Solvent flux and recovery vary with transmembrane pressure, feed solution concentrate (Cf) = 70.17 mol·m−3, pH = 6, T = 306 K; (b) Solute rejection and retentate solution concentration vary with transmembrane pressure, feed solution concentrate (Cf) = 70.17 mol·m−3, pH = 6, T = 306 K.
Figure 4. Influence of transmembrane pressure on RO process. (a) Solvent flux and recovery vary with transmembrane pressure, feed solution concentrate (Cf) = 70.17 mol·m−3, pH = 6, T = 306 K; (b) Solute rejection and retentate solution concentration vary with transmembrane pressure, feed solution concentrate (Cf) = 70.17 mol·m−3, pH = 6, T = 306 K.
Processes 07 00271 g004
Figure 5. Influence of the feed solution concentration on RO process. (a) Solvent flux and recovery vary with feed solution concentration, feed solution flow (Qf) = 1.56 × 10−4 m3·s−1, pH = 6, T = 306 K; (b) Solute rejection and retentate solution concentration vary with feed solution concentration, feed solution flow (Qf) = 1.56 × 10−4 m3·s−1, pH = 6, T = 306 K.
Figure 5. Influence of the feed solution concentration on RO process. (a) Solvent flux and recovery vary with feed solution concentration, feed solution flow (Qf) = 1.56 × 10−4 m3·s−1, pH = 6, T = 306 K; (b) Solute rejection and retentate solution concentration vary with feed solution concentration, feed solution flow (Qf) = 1.56 × 10−4 m3·s−1, pH = 6, T = 306 K.
Processes 07 00271 g005
Figure 6. Comparison of the experimental and theoretical solvent flux values.
Figure 6. Comparison of the experimental and theoretical solvent flux values.
Processes 07 00271 g006
Figure 7. Comparison of the k values.
Figure 7. Comparison of the k values.
Processes 07 00271 g007
Table 1. Main characteristics of the used membrane.
Table 1. Main characteristics of the used membrane.
TypePA2-4040
Membrane propertiesCompositionpolyamide
Permeable capacity (average, m3·d−1)7.2
Effective membrane area (m2)7.9
Recovery rate (single, %)15
Usage conditionsMaximum pressure (Mpa)4.14
Temperature (°C)5–45
Maximum flow (m3·h−1)3.6
Table 2. Solute concentration at the membrane surface (Cf = 70.17 mol·m−3, pH = 6, T = 306 K, R = 8.314 J·mol−1·K−1).
Table 2. Solute concentration at the membrane surface (Cf = 70.17 mol·m−3, pH = 6, T = 306 K, R = 8.314 J·mol−1·K−1).
QfpCpCδ,1Lp
(m3·s−1, ×10−6)(Pa, ×103)(mol·m−3)(mol·m−3)(m·s−1·Pa−1, ×10−12)
248.88893113.914270.571041.3923
218.6111346.53.897670.723041.3611
175.5556421.53.397271.067231.3423
152.2222488.53.274871.491081.3238
123.3333505.53.247171.975431.3048
75.55565303.085873.441941.3889
59.72225532.963574.339261.3709
48.33335782.963575.489011.3538
36.1111598.52.807877.800781.353
29.7222616.52.741179.98171.3529
Table 3. Value of the mass transfer coefficient (k) and Reynolds number (Re) under different Qf (Cf = 70.17 mol·m−3, pH = 6, T = 306 K, R = 8.314 J·mol−1·K−1).
Table 3. Value of the mass transfer coefficient (k) and Reynolds number (Re) under different Qf (Cf = 70.17 mol·m−3, pH = 6, T = 306 K, R = 8.314 J·mol−1·K−1).
Qf (m3·s−1, ×10−6)k (m·s−1, ×10−6)dh (m, ×10−4)ReSc
248.888937.61.4224151.2482
218.611133.611.4224132.8482
175.555627.151.4224106.6482
152.222221.331.422492.5482
123.333316.291.422474.9482
75.55569.491.422445.9482
59.72227.861.422436.3482
48.33336.491.422429.4482
36.11114.771.422421.9482
29.72223.881.422418.1482
Table 4. Value of k and Schmidt number (Sc) under different Cf (Qf = 155.56 × 10−6 m3·s−1, pH = 6, T = 306 K, R = 8.314 J·mol−1·K−1).
Table 4. Value of k and Schmidt number (Sc) under different Cf (Qf = 155.56 × 10−6 m3·s−1, pH = 6, T = 306 K, R = 8.314 J·mol−1·K−1).
Cf (mol·m−3)k (m·s−1, ×10−6)ScD/dh (m·s−1, ×10−6)
36.156722.24841.1710
62.444421.954821.1703
87.553321.774801.1696
111.50621.64791.1689
137.14921.44781.1682

Share and Cite

MDPI and ACS Style

Chen, C.; Qin, H. A Mathematical Modeling of the Reverse Osmosis Concentration Process of a Glucose Solution. Processes 2019, 7, 271. https://doi.org/10.3390/pr7050271

AMA Style

Chen C, Qin H. A Mathematical Modeling of the Reverse Osmosis Concentration Process of a Glucose Solution. Processes. 2019; 7(5):271. https://doi.org/10.3390/pr7050271

Chicago/Turabian Style

Chen, Chenghan, and Han Qin. 2019. "A Mathematical Modeling of the Reverse Osmosis Concentration Process of a Glucose Solution" Processes 7, no. 5: 271. https://doi.org/10.3390/pr7050271

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