Sei sulla pagina 1di 14

Journal of Membrane Science 325 (2008) 809822

Contents lists available at ScienceDirect

Journal of Membrane Science


journal homepage: www.elsevier.com/locate/memsci

Numerical study of two-dimensional multi-layer spacer designs for minimum drag and maximum mass transfer
G.A. Fimbres-Weihs, D.E. Wiley
School of Chemical Engineering and Industrial Chemistry, UNESCO Centre for Membrane Science and Technology, University of New South Wales, Sydney, NSW 2052, Australia

a r t i c l e

i n f o

a b s t r a c t
Based on the insights gained from previous published works, a series of multi-layer spacer designs for use in spiral wound membrane modules are proposed and evaluated via computational uid dynamics simulations. The lament diameter to channel height ratio of traditional cylindrical laments is reduced from 0.6 to 0.4 and 0.3, and one or two layers of elliptical laments with various attack angles are introduced in the middle region of the channel. The mass transport equations are solved in conjunction with the momentum and continuity equations for a solute with Schmidt number of 600, and the hydraulic Reynolds number is varied from 50 up to 800. Spacer performance is evaluated via a basic permeate processing cost analysis. The proposed designs did not lower processing costs when operating at hydraulic Reynolds numbers above 200, but showed potential for reducing costs in the steady laminar ow regime, at hydraulic Reynolds numbers equal to or less than 200. Implications for design improvements of spacer meshes, such as extra layers of spacer laments to direct the bulk ow towards the membrane walls, and changes to the lament proles to reduce form drag are discussed. 2008 Elsevier B.V. All rights reserved.

Article history: Received 2 July 2008 Received in revised form 24 August 2008 Accepted 3 September 2008 Available online 16 September 2008 Keywords: Computational uid dynamics Mass transfer Novel spacer Spacer design Multi-layer spacer

1. Introduction The spacer meshes in spiral wound membrane (SWM) modules that help keep the membrane leaves apart also serve other purposes. Flow disruption and boundary layer destabilization caused by the obstruction of the channel due to the spacer laments result in increased mixing and mass transfer, albeit at the expense of increases in energy losses. These effects have been demonstrated in various published works [17]. The benets of mass transfer enhancement often outweigh the disadvantages caused by increased energy losses, such that it becomes more economically attractive to operate membrane systems that utilize spacer meshes, compared to those that do not [3,8]. Despite the benecial impact of spacer meshes on mass transfer, some energy losses do not translate into enhanced mass transfer [7]. Different spacer congurations will lead to different mass transfer to energy losses ratios [9]. In addition, recent studies [6,7] suggest that certain ow characteristics lead to increased mass transfer (e.g. ow towards the membrane wall and wall shear perpendicular to the bulk ow direction). Therefore, it is expected that spacer lament designs that promote these ow characteristics, while keeping increases in energy losses to a minimum, would

Corresponding author. Tel.: +61 2 9385 5541. E-mail address: d.wiley@unsw.edu.au (D.E. Wiley). 0376-7388/$ see front matter 2008 Elsevier B.V. All rights reserved. doi:10.1016/j.memsci.2008.09.005

result in further economic improvements for the operation of membrane units. Spacer congurations consisting of more than the typical two layers of laments have been proposed [10,11]. The efcacy of multi-layer spacers is due to the ability of smaller near-wall laments to promote the formation of recirculation regions and vortex shedding in the vicinity of the membrane wall, thus disrupting the boundary layer. Additionally, the middle spacer lament layer or layers serve the purpose of re-directing the low-concentration bulk ow towards the membrane walls. Since form drag caused by the recirculation regions and vortex shedding behind the middle layers does not result in boundary layer separation or reattachment, it is therefore possible to reduce energy losses by optimizing the lament prole of the middle layer spacer laments. Kim and Kim [12] developed a numerical algorithm to calculate the minimum drag prole in 2D ow. They found that this prole depends on the Reynolds number. They also found that elliptical proles are very close to the optimal proles, with a total drag never more than 0.1% higher. The total drag of a circular prole was 4.54% higher at a Reynolds number of 1, but this ratio quickly increased to above 30% higher at a Reynolds number of 40. As elliptical laments present desirable drag characteristics, they appear as the obvious alternative for a middle-layer spacer lament. Moreover, the ability to manufacture elliptical lament proles should only depend on the ability to extrude this desired prole shape before constructing

810

G.A. Fimbres-Weihs, D.E. Wiley / Journal of Membrane Science 325 (2008) 809822

the spacer mesh, which is not considered a difcult task for modern extrusion equipment. This paper considers possible multi-layer spacer conguration designs for improving the mass transfer enhancement characteristics of typical zigzag spacers. For this purpose, various congurations incorporating elliptical laments as the middle spacer layers are simulated and analyzed. Due to the higher number of degrees of freedom in 3D spacer geometries, only 2D ow is considered in this study. The reason for this is that the main aim of this work is to understand the effect of these multi-layer spacer designs on mass transfer, serving as a screening tool for further detailed analysis and design of spacers. Williamson [13] reported that, under ideal conditions, the transition from 2D to 3D ow in a cylinder wake occurs at a cylinder Reynolds number of about 194, which translates to hydraulic Reynolds numbers above 500 for all of the geometries considered in this work, and above 745 for the geometries with a df /hch ratio of 0.3. In addition, Zovatto and Pedrizzetti [14] showed that the presence of nearby walls further delays the transition to 3D ow. Furthermore, Iwatsu et al. [15,16] suggest that for hydraulic Reynolds numbers above 800 the differences between 2D and 3D ow are signicant, such that the characteristics of 3D ow cannot be extrapolated from 2D calculations. Therefore, laminar steady and unsteady ow conditions with hydraulic Reynolds numbers up to 800 are investigated in this paper for spacer-lled channels. Moreover, an economic analysis is carried out in order to compare the relative performance of the spacer designs in terms of overall cost-effectiveness of the entire channel.

keeping the ratio of lm /df constant, and by introducing a middle spacer layer consisting of submerged elliptical laments with various angles of attack ( e ). The lm /df ratio was kept constant in order to maintain the same membrane area covered by the spacer mesh. As a consequence of keeping the lm /df ratio constant, changes in lament diameter for a given Reynolds number do not affect the aspect ratio (length to height) of the recirculation regions formed due to the presence of the obstruction. Comparisons of different mass transfer enhancement techniques are often made at the same Reynolds number [5]. However, this may not be the best option when energy losses are signicant. Huang and Tao [26] proposed making comparisons at the same pressure drop and at the same pumping power. From the denition of the Fanning friction factor used in this paper, the necessary pumping power per unit length can be calculated by 2fAT u3 Ws 2AT 3 3 Q pch eff = Reh f = = 2 d4 L L dh h (1)

From Eq. (1) it can be seen that for two different ows of the same uid to have the same pumping power per unit length, the following condition must be satised:
3 Reh1 f1 1 AT1 4 dh1

3 Reh2 f2 2 AT2 4 dh2

(2)

2. Problem description, assumptions and methods The commercial CFD code ANSYS CFX-11.0 is used to solve the continuity, momentum and mass transport equations [17] in a spacer-lled channel. Previous work [18,19] has shown that neither gravity nor density variation will have a signicant effect on the solutions obtained, therefore constant properties are employed and the effect of gravity was excluded. In order to further simplify this system the uid is assumed to be Newtonian, the ow twodimensional, and a binary mixture of water and salt is considered, with no sources of salt in the uid. Mass transfer is incorporated in the form of a dissolving wall boundary condition. Although a permeable wall boundary condition [19,20] in which the wall mass fraction is dependent on the permeation velocity would be more physically accurate, the relative magnitude of the permeation velocity is usually a few orders of magnitude smaller than the average uid velocity, such that the effect of permeation on wall shear and mass transfer enhancement is minimal [21]. In addition, for channel geometries and ow conditions typically found in NF and RO, Geraldes and Afonso [22] have recently demonstrated that Sherwood numbers obtained for cases without permeation can be used to predict the wall conditions in cases where permeation is present. The impermeable-dissolving wall model is therefore capable of providing valuable insights into the ow eld-induced mass transfer phenomena taking place inside narrow channels, and the results obtained should still represent a good approximation of the ow and mass transfer in SWM modules. The geometries analyzed in this paper are variations of the 2D zigzag spacer arrangement [6,2325]. A basic zigzag geometry was chosen because it presents the most similarities to spacers used in real membrane modules. Moreover, using parameters df /hch = 0.6 and lm /hch = 4, it was also found to be better performing than the other geometries studied by Schwinge et al. [24], with regards to its mass transfer and pressure loss characteristics. Modications to this conguration were made by decreasing the df /hch ratio while

The geometries considered in this work were chosen to have the 4 same /dh ratio, such that comparisons at the same Power number 3 f ) value would have the same pumping power requirements (Reh per unit volume of the respective channel As a result, increases 3 in mass transfer performance at the same Reh f value are directly related to changes in spacer arrangement, and would represent a better performing spacer conguration regardless of the economic impact of pumping costs. A general unit cell for the family of geometries considered in this paper is shown in Fig. 1. The aspect ratio of the elliptical laments (aE /ae ) was chosen such that it was as close as possible as that of the optimal ellipse for minimum drag, as reported by Kim and Kim [12]. This gave an aE /ae ratio of 5 for the cases with a df /hch ratio of 0.4, and an aE /ae ratio of 4 for the cases with a df /hch ratio of 0.3. For the cases where the angle of attack of the elliptical laments was not zero, the elliptical laments were tilted in opposite directions in order for the lift generated by these laments to cancel each other. The distance from the centre of the elliptical laments to the channel walls was chosen to be either half of the channel height or as given by the following equation: he =
1 (h 3 ch

ae ) + 0.3df

(3)

The congurations considered are portrayed in Fig. 2. In the 3-layer (3L) conguration, the elliptical laments are placed equi-distant from both membrane walls. In the 4-layer (4L) congurations, successive elliptical laments are placed at a distance from one of the membrane walls as given by Eq. (3). This was done in order to allow a reduction in elliptical lament size while keeping the spacer conguration representative of a real-world spacer mesh, i.e. without leaving gaps in the channel height. For each 4L spacer there are two possible locations for the rst elliptical lament, depending on which membrane wall it is closer to. If the rst elliptical lament is closer to the membrane wall to which the upstream circular lament is attached, the conguration is designated as Low (-L). Conversely, if the lament is closer to the membrane wall to which the downstream circular lament is attached, then it is designated as High (-H). The designation by which each spacer conguration will be referred to in this paper is indicated in parentheses in Fig. 2.

G.A. Fimbres-Weihs, D.E. Wiley / Journal of Membrane Science 325 (2008) 809822

811

Fig. 1. Generic unit cell geometry of the spacer type modeled in this paper.

It must be noted that geometries 2L04 and 2L03 are not representative of typical spacer geometries because the spacer height is smaller than the channel height, and a gap is present in between the spacer laments, as seen in Fig. 2. In real world membrane modules, the spacer laments usually touch each other, as they are welded together in layers or woven so as to keep the membrane leaves apart. Nonetheless, these geometries could be manufactured if a third lament parallel to the ow is introduced, or via the use of spacer protuberances. This paper does not consider the additional manufacturing costs that such designs might incur. For steady laminar ow at low Reynolds numbers, it generally takes several repetitions of the unit cell before the Sherwood number reaches a limiting value. In order to avoid the need to simulate an excessively long channel at low Reynolds numbers, two separate simulations were carried out. The rst consisted of an array of at least eight repetitions of the unit cell, where the entrance and exit length dimensions were chosen based on the ndings of Schwinge et al. [24]: an entrance length of 10 times the channel height and an exit length twice as long as the entrance length. These long entrance and exit lengths were necessary to fully develop the velocity proles before the beginning of the spacer array, and to prevent the outlet condition interfering with the recirculation regions after the last lament. The second simulation involved three repetitions of the unit cell and the application of the methodology for

periodic wrapping described by Fimbres-Weihs and Wiley [7]. The rst simulation allows quantication of the entrance effects in the channel, while the second simulation permits the calculation of the characteristics far away from the channel inlet. For unsteady ow, the Sherwood number achieves its limiting value much closer to the channel entrance than in the case of steady laminar ow, often after as few as ve repetitions of the periodic unit cell [27]. In addition, the results of Fimbres-Weihs et al. [6] revealed that at least six spacer laments were necessary for vortices to appear in unsteady ow. Therefore, the same channel characteristics were used for the unsteady runs as for the entrance region runs under steady laminar ow. As in our previous works [6,7], a Schmidt number of 600 was chosen, which is characteristic of typical monovalent salts, such as sodium chloride. The channel walls over the entrance and exit lengths as well as the spacer surfaces are treated as non-slip walls with no mass transfer, where both velocity components and the mass fraction gradient normal to the boundary are set to zero (u = v = 0 and Y /nw = 0). The membrane walls are also treated as non-slip walls (u = v = 0), but the mass fraction at the wall is xed at a constant value (Y = Yw ). For the entrance region and unsteady runs, a at velocity prole with u = uavg , v = 0 and Y = 0 is specied at the inlet of the ow domain. At the outlet, an average reference pressure of zero is spec-

Fig. 2. Schematic of the unit cells for the different spacer congurations analyzed in this paper.

812

G.A. Fimbres-Weihs, D.E. Wiley / Journal of Membrane Science 325 (2008) 809822

ied. For the steady laminar runs under fully developed conditions, the inlet and outlet boundaries were treated as fully developed velocity and mass fraction proles, using the approach described by Fimbres-Weihs and Wiley [7]. The results were taken from the middle unit cell, following the approach of Rosaguti et al. [28], in order to avoid the error introduced due to the lack of gradient information at the periodic boundaries. Mesh independent solutions were obtained using an unstructured mesh with element sizes of the order of at most 3% of the circular lament diameter. This element size was determined after a series of runs with increasingly ner meshes. Inated boundaries were used for all the boundaries with a non-slip condition, where the thickness of the rst grid element layer was of the order of at most 0.15% of the circular lament diameter. Angular resolution on the surfaces of the laments was set to 17 , and to 15 for the corners between the laments and the membrane wall. The time steps used for the unsteady runs were determined by decreasing the time step size until the solution did not depend on this parameter. Typically, the RMS Courant number for the unsteady runs was less than 0.2, and the maximum Courant number was less than 3. Following the approach used by Fimbres-Weihs et al. [6], an approximate steady-state solution was chosen as the initial state (t = 0) for the transient simulations. Using this initial condition, the statistical quantities of all the variables stabilized after approximately one residence time. 2.1. Channel interpolation In this work, solutions to the momentum and mass transport equations are obtained for the entrance region and for conditions of fully developed velocity and mass fraction proles. These two results can be used to estimate the mass transfer and pressure drop conditions for the region where entrance effects are not dominant, but the mass transfer conditions are not yet fully developed. Churchill and Usagi [29] developed a simple correlation to approximate intermediate solutions for transfer processes for which asymptotic solutions are known at large and small values of the independent variable. Mass transfer in a narrow channel under laminar ow, such as investigated in this paper, ts this category of problems with the Sherwood number as the dependent variable, and the dimensionless channel length as the independent variable. Following the approach of Churchill and Usagi, the following is a suitable interpolation expression:
B Sh = [(Ax ) + ShN ] fd N 1/N

Fig. 3. Conceptual diagram of interpolation functions, as dened by Eqs. (4) and (5), showing asymptotes and interpolation curves.

and friction factor behavior of an entire spiral wound module at a given Reynolds number. 2.2. Cost estimation In order to effectively compare the relative performance of different spacer geometries in real-world operations, an economic analysis is needed. However, a thorough economic analysis would have to take into account many factors that vary greatly with geographical region and also over time including salaries, equipment cost and interest rates. Such a complete analysis is outside the scope of this paper. However, it is possible to obtain an approximation of total costs by analyzing the direct costs for the production of permeate, without taking into account pre-treatment costs, which are independent of the spacer geometry in the SWM units. In addition, Maskan et al. [30] explain that cleaning costs are largely proportional to the membrane area, and can therefore be considered as being incorporated with the membrane cost. Moreover, due to the nature of the spacer prototypes considered in this chapter, estimation of their manufacturing costs would be outside the scope of this work. Therefore, the cost analysis carried out in this chapter does not directly take into account membrane cleaning costs, and ignores pre-treatment and spacer costs, focusing instead on direct permeate processing costs. 2.2.1. Permeate calculation The local permeate ux can be calculated following the approach of Kedem and Katchalsky [31] and Merten [32], which yields the following expression: Jsln = Lp ( ptm
tm )

(4)

A similar expression is also applicable for the friction factor:


B N f = [(Ax+ ) + ffd ] N 1/N

(5)

Entrance region effects are quantied by parameters A and B in Eq. (4), and parameters A and B in Eq. (5). Parameters N and N quantify the transition from entrance region to fully developed ow, as seen in Fig. 3, i.e. a very high N or N would signify a sharp transition, which would follow the entrance region trend until it reaches with the fully developed value, and then follow the latter. A lower N or N would result in a smoother transition. Such a function is similar to a hyperbolic function that describes the decline in Sherwood number and friction factor over the entrance region, and that decays to the asymptotic value for far downstream regions of the channel as shown in Fig. 3. The parameters in Eqs. (4) and (5) were estimated via a non-linear regression using the results from the channel simulations in the laminar regime. Using this interpolation method, a single equation for each of Sh and f can therefore be obtained using the data from the two different simulations in this paper. The single equation should be suitable for describing the Sherwood number

(6)

The osmotic pressure of the solute can be approximated by a linear expression [33]: = Y (7)

where the value for the osmotic pressure coefcient () used was the value reported by Geraldes et al. [34], and is shown in Table 1. Eqs. (6) and (7) can be combined to obtain an expression for the local permeate ux in terms of the local wall solute mass fraction. The local solute mass fraction at the membrane wall can be directly obtained from CFD simulation results using a permeable wall boundary condition. However, since the simulations reported in this paper use an impermeable wall, i.e. a dissolving wall where there is no uid permeating through the wall, a more convenient approach is to use the mass transfer coefcient data obtained from

G.A. Fimbres-Weihs, D.E. Wiley / Journal of Membrane Science 325 (2008) 809822 Table 1 Case study parameters for cost analysis Module length (L) Channel width (wch ) Salt rejection Feed mass fraction (Yb,in ) Inlet transmembrane pressure ( ptm,in ) Reection coefcient ( ) Osmotic pressure coefcient () Membrane permeability (Lp ) Channel height (hch ) Energy cost (Ce ) Membrane cost (Cm ) Amortization factor (Fa ) Pump efciency (pump ) Operation time (top ) 90 cm 90 cm 99.6% 0.025 60 atm 1 8.051 107 Pa 3.94 106 m/(s atm) 0.001 m $0.10 kW/h $100/m2 0.4/yr 0.6 8000 h/yr

813

where the membrane area (Am ) is the product of the width and length of the channel, and the average permeate ux (Jsln,avg ) can be estimated as the average of the local permeate ux values calculated for the inlet and outlet ow conditions of the membrane module. In this paper, the permeate ux at the inlet ow conditions is calculated using Eq. (10) and the inlet conditions ( pin , kmt,in,per and Yb,in ). Likewise, the permeate ux at the outlet is calculated using Eq. (10) and the outlet conditions ( pout , kmt,out,per and Yb,out ) in an iterative manner. In order to calculate the outlet operating conditions, the following relationships are used: Qin = uavg,in hch wch Qout = Qin Qp (14) (15) (16)

the simulations. The local mass transfer coefcient is dened as kmt = D (Yw Yb ) Y y (8)
w

Yb,out =

Qin Yb,in Qp Yp Qout Qout hch wch ptm,in


2L 3 dh 2 Reh,avg f

uavg,out =

(17) pch (18) (19)

From the mass balance of solute at the membrane surface, it is also known that: Jsln Yw = D Y y + Jsln Yp
w

ptm,out = pch = 2

(9)

Combining these equations yields the following expression for the permeate ux: Jsln 1 = (Lp 2 ptm + kmt ) 1 ( kmt Lp 2
2

ptm )

+ kmt Lp (Yb Yp )

(10)

Eq. (10) can be used to calculate the local permeate ux in terms of the operating parameters and the mass transfer coefcient. However, the mass transfer coefcient obtained from the CFD simulation results in this paper is calculated using impermeable wall conditions. In the real case, the permeation of uid through the membrane wall alters the mass transfer conditions, and hence the mass transfer coefcient, when compared to the dissolving wall case. Geraldes and Afonso [22] have shown that the mass transfer coefcients for permeable wall conditions can be related to the impermeable (dissolving wall) mass transfer coefcients using the following relationship. Shper kmt,per = = = Shimp kmt,imp + (1 + 0.26
1.4 1.7

In Eq. (19), the average Reynolds number is calculated as the average of the inlet and outlet Reynolds numbers. The iterative procedure for the calculation of the permeate ux at the outlet is slightly more complex than that for the permeate ux at the inlet. This is because the outlet permeate ux estimate is also used for calculating the permeate ow (Qp ) through Eq. (13), which in turn is used for calculating the outlet Reynolds number through Eqs. (15) and (17). The impermeable Sherwood numbers and friction factors used for calculating the inlet and outlet permeate uxes are obtained using the interpolation Eqs. (4) and (5), respectively. The parameters in the interpolation equations are, in turn, estimated via a non-linear regression using the results from the entrance region and fully developed channel simulations, as described in Section 2.1. Once converged values for both the inlet and outlet permeate uxes are obtained, the value for the permeate ow can be obtained through Eq. (13). 2.2.2. Cost calculation Knowing the permeate ow, the capital cost for the membrane unit per cubic meter of permeate ow can be calculated using the following expression: Cc = Am Cm Fa Qp top (20)

(11)

where the variable is the ratio of permeation Peclet number to the impermeable Sherwood number, and can be calculated using the following equation: = Jsln kmt,imp (12)

Operating costs are proportional to the pumping energy, which is equal to: Ws = pch Qin (21)

Therefore, Eqs. (10)(12) can be used to calculate the local permeate ux at given operating conditions ptm , kmt and Yb . Now the permeate ux is an input required for calculating the permeable wall mass transfer coefcient. Thus, an iterative procedure is used to obtain an estimate of the permeate ux accurate to six signicant gures. The permeate production of a membrane module is directly proportional to the average permeate ux, and can be calculated using the following expression: Qp = Am Jsln,avg (13)

The operating cost per cubic meter of permeate ow is then given by Cop = Ws Ce Qp pump (22)

Finally, the total processing cost per cubic meter of permeate ow is calculated as the sum of the operating and capital costs: Ctot = Cop + Cc (23)

814

G.A. Fimbres-Weihs, D.E. Wiley / Journal of Membrane Science 325 (2008) 809822

Fig. 4. Friction factor dependence on hydraulic Reynolds number for 2-layer spacerlled channels with varying df /hch ratios.

Fig. 6. Sherwood number distribution along channel length for spacer-lled and empty channels, at a hydraulic Reynolds number of 200.

3. Results and discussion 3.1. Two-layer spacer geometries The rst series of runs conducted was for 2-layer (2L) spacers with circular laments, i.e. without elliptical lament layers. The df /hch ratio was varied from 0.6 down to 0.3, while keeping the ratio of lm /df at a constant value of 6.667. As the lament diameter is decreased, the cross-section of the channel that is obstructed is also decreased. It is therefore expected that for lower lament diameter to channel height ratios the friction factor will be lower, approaching the value for an empty channel. Fig. 4 shows that this is the case. The effect of lament diameter on Sherwood number is more complex than its effect on the friction factor. For empty channels, the Sherwood number for ow with a fully developed concentration prole does not depend on the Reynolds number. However, as shown in Fig. 5, for spacer-lled channels the Sherwood number increases as the Reynolds and Power numbers are increased, as a consequence of increased boundary layer destabilization at higher Reynolds numbers. At low Reynolds numbers, where no boundary layer separation occurs and creeping ow conditions are prevalent, the Sherwood number will be lower for the spacer-lled channels than for the empty channel as a result of the spacer laments covering the membrane surface and reducing the available mass transfer

area, i.e. membrane coverage. This effect is evidenced for Power numbers below 105 in Fig. 5. For empty channels, entrance region effects caused by the developing boundary layer result in a higher Sherwood number near the channel inlet [35]. As the distance from the channel inlet is increased, the Sherwood number eventually reaches its asymptotic value. Since further increases in channel length do not affect the Sherwood number, the concentration prole is considered to be fully developed. This phenomenon is also encountered in spacer lled channels as can be observed in Fig. 6, which was obtained by making use of the interpolation functions described by Eqs. (4) and (5). As a consequence of the entrance region effect, the ratio of energy losses to mass transfer is increased as the channel length is increased. This suggests that short and wide channels will result in lower energy needs for given mass transfer needs. As mentioned in Section 2. the geometries considered in this 4 work were chosen to have the same /dh in order for comparisons 3 at the same Reh f value to have the same pumping power require3 ments per unit volume of the respective channel. The group Reh fL is therefore a measure of accumulated energy losses at a specied channel length (L). In order to make this value dimensionless, it can be multiplied by 4 /dh (which was kept constant for all the geometries modeled in this paper), thus giving the dimensionless group
3 Reh f 4 L/dh . From Fig. 6 it can be seen that for a given energy loss, the Sherwood number is always higher for spacer lled channels than for an empty channel. In addition the empty channel requires less energy to attain a fully developed concentration prole than the spacer-lled geometries tested. This is to be expected given that pressure drop for the empty channel is signicantly lower than for the spacer-lled cases, as was shown in Fig. 4. In addition, Fig. 6 shows that the geometries with a higher Sherwood number under fully developed concentration prole conditions will usually have a higher Sherwood number at any given value for the energy loss throughout the entire length of the channel. This is conrmed in Fig. 6, where it is shown that the lines for the different spacers do not intersect each other, and the asymptotic value of the Sherwood number for the 2L03 geometry is slightly higher than that for the empty channel. 1 1

3.2. Multi-layer spacer geometries In order to assess the effect of the angle at which the elliptical laments intersect the oncoming bulk ow, simulations of congurations with different angles of attack were undertaken. Because of

Fig. 5. Sherwood number dependence on Power number for 2-layer spacer-lled channels with varying df /hch ratios.

G.A. Fimbres-Weihs, D.E. Wiley / Journal of Membrane Science 325 (2008) 809822

815

the large number of geometric variations explored in this paper, and due to the high computational and time demands of transient simulations, it would not have been practical to carry out transient simulations for all of the spacer geometries. Therefore, for the purposes of isolating the effect of the angle of attack, the hydraulic Reynolds number was not varied and kept constant at a value of 200, which is the highest hydraulic Reynolds number for which all of the geometries present steady ow conditions. In addition, this Reynolds number is representative of typical operating conditions in real-world membrane systems which operate in the steady-laminar ow regime. Aerodynamic studies show [36] that for airfoils, form drag and its variation are small at small angles of attack. Moreover, as the angle of attack is increased above a threshold value (usually in the range of 5 to 15 for Reynolds numbers of the order of 100,000) ow separation causes form drag to increase drastically (usually doubling or tripling its value). Therefore, the angle of attack of the elliptical laments in this study was kept within the range of +10 to 10 . Due to the relative positioning of the circular and elliptical spacer laments and the denition of the angle of attack, a negative angle of attack is expected to align the elliptical spacer lament with the bulk ow such that form drag is reduced. Fig. 7 shows that as the attack angle increases, a larger region of stagnant uid forms downstream of the elliptical laments, thus increasing form drag. It can also be seen that this region of stagnant ow, caused by boundary layer separation, is smaller for the negative angles of attack than for the positive values. At an angle of 5 the velocity of the uid owing above and below the elliptical lament is approximately equal and the stagnant region is smaller than at the other angles of attack. Therefore, for the 3L04 spacer, form drag is expected to be lower for an attack angle of 5 than for the other values. The effect of the angle of attack on form drag for all spacer congurations and angles of attack can be seen in Fig. 8. For all these congurations the lowest percentage of energy losses due to form drag is attained at an angle in the range of 5 to 7.5 . In the case of the 3L04 conguration, form drag is reduced as the angle of attack is reduced from 0 to 5 , and increases again as the angle is reduced further to 10 . This suggests that, as expected from Fig. 7, an angle of 10 overshoots the negative angle value necessary to align the elliptical lament with the bulk ow. As a result, this angle causes

Fig. 8. Variation of the percentage of energy losses due to form drag with angle of attack for multi-layer spacer geometries, at a hydraulic Reynolds number of 200.

a larger region of stagnant uid behind the obstacle, thus leading to a larger percentage of form drag than at a 5 angle. Since form drag is the main component of pressure drop, it can provide an indication of the behavior of the friction factor as the angle of attack is changed. As can be seen in Fig. 9, however, the pattern of lower friction factor for negative angles of attack is only followed by the 4L04L and 4L03L congurations. For the 3L04, 4L04H and 4L03H congurations the friction factor is at its lowest value at an angle of attack between 5 and 0 . The increase in friction factor at negative angles of attack for the -L congurations is due to an increase in the viscous drag. Fig. 8 conrms that the percentage of pressure drop due to form drag for these cases is lower than for the other congurations. The reason for the increase in viscous drag is visualized in Fig. 10. It can be seen in this gure that for the 4L03L and 4L04L congurations the relative positioning of the elliptical lament to the circular laments causes most of the uid to ow on one side of the elliptical laments. This channeling of the uid creates a large area of stagnant or slow moving uid downstream of the circular laments, between the elliptical laments and the membrane wall. The relatively slow owing uid does not contribute signicantly to viscous drag on the elliptical lament and the membrane

Fig. 7. Velocity contour plots for the 3L04 spacer congurations at various angles of attack, and a hydraulic Reynolds number of 200.

Fig. 9. Dependence of friction factor on angle of attack for multi-layer spacer geometries, at a hydraulic Reynolds number of 200.

816

G.A. Fimbres-Weihs, D.E. Wiley / Journal of Membrane Science 325 (2008) 809822

Fig. 10. Velocity contour plots for the 4L03H, 4L03L, 4L04H and 4L04L spacer congurations at an attack angle of 10 , and a hydraulic Reynolds number of 200.

Fig. 11. Dependence of Sherwood number on angle of attack for multi-layer spacer geometries, at a hydraulic Reynolds number of 200.

wall surfaces. In contrast, for the 4L03H and 4L04H congurations the ow is more evenly distributed on both sides of the elliptical lament. Therefore, both sides of the elliptical laments and membrane walls experience higher viscous drag in the -H congurations than in the -L congurations. This results in a higher friction factor for the -H congurations at the 10 angle of attack than at the 0 angle, despite form drag being lower for the negative angle. For the effect of the angle of attack on the Sherwood number, it is expected that both positive and negative angles of attack would increase the average Sherwood number over the unit cell by directing ow away from the bulk and towards the membrane walls. As seen in Fig. 11, this is only the case for the 3L04 and (to a lesser extent) 4L04H congurations. For those cases, the nonzero angles of attack cause low concentration uid to be diverted from the bulk ow towards the membrane surface, causing a larger degree of boundary layer renewal and thus a higher Sherwood number. In contrast, the 4L03H and 4L03L congurations show a local minima for the average Sherwood number when the angle of attack is 10 or +10 . In these latter cases, the large angles of attack reduce the velocity of the uid owing between the elliptical lament and the membrane wall as shown previously in Fig. 10. This in turn reduces the wall shear rate and increases the concentration boundary layer thickness. Since less uid is being forced to make contact with the membrane surface, less boundary layer renewal occurs and, as a result, the average Sherwood number is also reduced. This effect is evidenced by comparing the velocity and mass fraction contour plots in Fig. 12, which show that the bulk of the uid channels past the ow obstructions follow-

ing a zigzag path, and generates large stagnant areas where the solute concentration builds up and mass transfer into the bulk uid is low. For the case of the 4L04L conguration, the Sherwood number tends to increase as the angle of attack is increased. This is because the negative angle of attack reduces the amount of uid that ows near the membrane wall, whereas the positive angle of attack increases the amount of low concentration bulk uid that approaches the membrane wall. It is well known that mass transfer can be enhanced by boundary layer destabilization through the use of ow obstructions [1,3739]. These obstructions, in turn, increase the channel pressure drop, i.e. energy losses. Since different angles of attack result in different combinations of mass transfer enhancement and energy losses, the Sherwood number to friction factor ratio can be used to identify the angle of attack that achieves the best balance between mass transfer enhancement and energy losses. This comparison is shown in Fig. 13. For all but the 4L04L and 4L03L congurations, the Sherwood number to friction factor ratio was higher at a zero angle of attack. This means that for all but the 4L04L and 4L03L congurations, the benet of an increased Sherwood number is outweighed by the increase in energy losses caused by a non-zero angle of attack. Both the 4L04L and 4L03L congurations show a higher Sherwood number to friction factor ratio at the 5 angle of attack. It is interesting to note that, although both the 4L04L and 4L03L spacer congurations showed a lower friction factor due to the negative angle of attack, the 4L03L conguration prevented ow from

Fig. 12. Velocity and solute mass fraction contour plots for the 4L03L spacer conguration for various angles of attack, at a hydraulic Reynolds number of 200.

G.A. Fimbres-Weihs, D.E. Wiley / Journal of Membrane Science 325 (2008) 809822

817

Fig. 15. Dependence of cost components on hydraulic Reynolds number for two sample geometries. A membrane cost of $100/m2 was assumed. Fig. 13. Dependence of the Sherwood number to friction factor ratio on angle of attack for multi-layer spacer geometries, at a hydraulic Reynolds number of 200.

reaching the region near the membrane wall, whereas the 4L04L conguration did not exhibit this behavior. For the 3L04 spacers, the conguration with a 10 angle of attack is aligned with the ow, and has lower friction factor than the conguration with a +10 angle of attack. From Fig. 14 it can be seen that the conguration with a 10 angle of attack also has more evenly distributed local Sherwood number values along the membrane wall than the other angles of attack, which results in one of the highest overall Sherwood numbers of this series of spacers. However, these overall Sherwood number values are still lower than for the conguration without an elliptical lament (2L04). In general, it can be said that increases in the angle of attack lead to increases in both the friction factor and Sherwood number. Although there are specic situations in which due to the geometrical alignment of the spacer laments this relationship might not hold, it has been shown that this trend is generally followed. In addition, positioning the elliptical laments in the -L conguration will result in higher Sherwood numbers. It must be said, however, that this trend was found at a constant lm /df ratio of 6.667, and there is the possibility of a different trend if the lm /df ratio is changed signicantly from this value. 3.3. Economic analysis of spacer performance The economic analysis of the different spacer congurations was carried out following the approach described in Section 2.2. In order to isolate the impact of the spacer conguration used, only processing costs were calculated, i.e. feed pre-treatment costs and operating costs other than pumping through the mem-

brane module (e.g. cleaning costs) were neglected. In addition, membrane-cleaning costs are assumed to be proportional to membrane area (and hence included in the membrane cost). Spacer manufacturing costs are also neglected. Therefore, the costs reported in this section do not represent the total cost of permeate production, and should only be considered as a comparative indicator of spacer performance, as explained in Section 2.2. A set of cost parameters, representative of what is currently found in literature [3,9,40], was utilized for the cost analysis case study. These parameters are summarized in Table 1. In addition, three different membrane costs were considered: a case study base cost of $100/m2 , a higher cost of $500/m2 , and a lower cost of $10/m2 of membrane. The high membrane cost is representative of a membrane system which is very prone to fouling, and thus includes the cost of cleaning and replacement in case of blockage. The low membrane cost is representative of the possible future membrane costs, since current trends indicate that membrane production costs will continue to decrease. As the Reynolds number is increased, the components of the total cost (Cc and Cop ) vary in opposite directions. Given that energy losses increase as the Reynolds number increases, the operating costs also increase. On the other hand, increases in Reynolds number lead to increases in Sherwood number, which in turn leads to decreased membrane area needs due to increased mixing and mass transfer. Therefore, typical cost curves follow a U shaped trend (see Fig. 15). At low Reynolds numbers, well within the laminar ow regime, capital costs dominate due to low mass transfer and low energy losses. As the Reynolds number is increased and capital costs are reduced, operating costs are increased until they become the dominant component. In the region where operating costs dominate, the total cost increases as the Reynolds number is increased further. Therefore, there exists a point in any total cost curve at which the total cost is minimized, and either an increase or decrease of Reynolds number would lead to an increase in total cost. The Reynolds number at which this minimum total cost occurs will be referred to as the optimal Reynolds number for operation. For the data presented in Fig. 15, the optimal Reynolds number for the empty channel is between 1000 and 2000, and around 400 for the 4L03H spacer geometry. 3.3.1. Base membrane cost of $100/m2 Fig. 16 depicts the effect of lament diameter and Reynolds number on permeate processing costs for the geometries analyzed in this paper. Despite energy losses encountered in spacer lled channels being larger than for empty channels, the cost of permeate production using spacers is lower than the cost with-

Fig. 14. Local Sherwood number for the bottom membrane wall of 2L04 and 3L04 spacer geometries with varying angles of attack, at a hydraulic Reynolds number of 200.

818

G.A. Fimbres-Weihs, D.E. Wiley / Journal of Membrane Science 325 (2008) 809822

Fig. 16. Dependence of total processing cost per cubic meter of permeate on hydraulic Reynolds number, for empty, 2-, 3- and 4-layer spacer-lled channels. The graph on the left shows effect of lament diameter on cost. The graph on the right compares costs for selected multi-layer spacers against empty channel and 2L06 spacer. A membrane cost of $100/m2 was assumed.

out using spacers. Modules using spacers have smaller membrane area requirements due to the higher Sherwood numbers attained, thus offsetting the higher energy costs required for their operation. This result agrees with previous studies [3,8]. However, at hydraulic Reynolds numbers below 100, spacer geometries 2L03 and 2L04 do not present signicant cost benets over the empty channel. As explained in Section 3.1. This is due to the amount of membrane area covered by the spacer laments in these congurations. Membrane coverage results in lower Sherwood numbers, and hence higher permeate processing costs than for the channel without spacers. In any case, operating at hydraulic Reynolds numbers below 100 would not be recommended under the conditions of this study, as Fig. 16 shows that lower permeate processing costs can be achieved by operating in the hydraulic Reynolds number range of 300 to 1000. In addition, operating at low Reynolds numbers also increases the potential impact of fouling. The results presented in Fig. 16 show that lower processing costs are achieved by the 2L04 and 2L06 spacer geometries. For both of these geometries, as well as for the 2L03 spacer geometry, the trends indicate lower costs as the Reynolds number is increased. Given that it is not clear whether a minimum in the cost curve has been reached at the highest Reynolds number for which data was collected, it is unclear which lament diameter to channel height ratio will achieve the lowest permeate processing cost. Moreover, ow becomes unsteady for these geometries at Reynolds numbers higher than the ones presented in Fig. 16. Therefore, transient data is needed before the best performing spacer conguration for these cost parameters and ow conditions can be determined. Nonetheless, one conclusion can be drawn from these simulations results: for 2D 2L zigzag spacer meshes operating in the laminar ow regime, signicant permeate cost reductions cannot be achieved by reducing the lament diameter to channel height ratio from 0.6 to 0.4 or 0.3. The multi-layer spacers geometries proposed in this paper show lower permeate processing costs than the 2L06 geometry for hydraulic Reynolds numbers below 200. This is partially due to the decrease in open volume inside the channel (due to the added spacer layers) which leads to an increase in local velocity and wall shear rates. It is also due to the increased mass transfer enhancement caused by the elliptical laments directing low concentration uid towards the membrane wall. However, the added lament

layers also increase energy losses and, thus, operating costs. At hydraulic Reynolds numbers above 300 the ratio of energy costs to membrane costs is higher for the multi-layer congurations than for the 2L06 spacer and, as a result, lower permeate processing costs are achieved with the 2L06 spacer at hydraulic Reynolds numbers above 300. Fig. 17 shows a comparison of the different geometries at a df /hch ratio of 0.4. The results shown in Fig. 17 are similar to those seen in Fig. 16, in that multi-layer spacers present lower permeate processing costs at hydraulic Reynolds numbers below 200. In addition, the 3L04 spacer at an angle of attack of 0 results in lower costs than its 4L04H counterpart. From the trends shown in Fig. 16 it is difcult to predict which geometry would be the better overall performer as the Reynolds number is increased beyond 200, which would lead to unsteady ow. Transients simulations would be necessary in order to determine which conguration would result in the lowest permeate processing cost. The scenario for the geometries at a df /hch ratio of 0.3 is somewhat similar to that for the geometries at a df /hch ratio of 0.4. As opposed to the cases at df /hch ratio of 0.4 and 0.6, transient simulations were carried out for the 4L03H geometries, and these results

Fig. 17. Dependence of total processing cost per cubic meter of permeate on hydraulic Reynolds number, for spacer-lled channels with a df /hch ratio of 0.4. A membrane cost of $100/m2 was assumed.

G.A. Fimbres-Weihs, D.E. Wiley / Journal of Membrane Science 325 (2008) 809822

819

Fig. 18. Dependence of total processing cost per cubic meter of permeate on hydraulic Reynolds number, for spacer-lled channels with a df /hch ratio of 0.3. A membrane cost of $100/m2 was assumed.

are presented for the hydraulic Reynolds number of 800 in Fig. 18. It can be seen that the use of 4-layer spacer geometries resulted in lower permeate processing costs for all of the Reynolds numbers tested in the laminar ow regime, but the use of the 2L03 spacer resulted in the lowest processing costs in the unsteady ow regime, at a hydraulic Reynolds number of 800. From the data collected from the transient simulations, it can be concluded that under the assumptions of this study, operating at such a high Reynolds number is not recommended for these 4-layer geometries as the increased energy costs associated with the increased energy losses outweigh the benets of increased permeate production due to increased mass transfer. However, operating at a high Reynolds number may be justied for other reasons not considered here, such as reduced cleaning costs due to reduced fouling. Nevertheless, the low cost achieved by using the 2L03 spacer at a hydraulic Reynolds number of 800 cannot be reduced by using the 4-layer spacers at a df /hch ratio of 0.3. The economic impact of conguration variations was also analyzed. The effect of lament positioning and angle of attack on permeate processing cost for multi-layer spacer meshes are shown in Fig. 19. From these results it can be seen that changes in positioning of the elliptical laments (i.e. -H or -L) have a larger effect on permeate processing costs than changes to the angle of attack. Regardless of the angle of attack, the -L congurations for the 4layer spacers always resulted in lower permeate processing costs than the -H congurations. This result agrees with the predic-

tions using the Sherwood number to friction number ratio (see Fig. 13), and can be attributed to the uid channeling effects seen in Figs. 10 and 12. The positioning of the elliptical spacers in the -L congurations generates channels of faster owing low concentration uid, which cause larger velocity and concentration gradients at the membrane wall, thus resulting in higher mass transfer and lower permeate processing costs. Changes in the angle of attack for the 4L04H spacer do not appear to have a signicant effect on permeate production costs (less than 1%). In contrast, increases to the angle of attack for the 4L04L conguration decrease the total permeate processing cost, albeit only by a small amount (4.1%). In the case of the 3L04 spacer, both positive and negative angles of attack result in a signicant reduction of the total permeate processing cost (16% and 13%, respectively). However, from the data presented in this paper, it is not possible to determine either the value or the direction of the angle of attack that would produce the lowest processing cost. For multi-layer spacer meshes with a df /hch ratio of 0.3, the angle of attack has a signicant effect on permeate processing costs, with the difference between the lowest and highest processing costs being over 30% in the angle range studied. Fig. 19 suggests that the optimal angle of attack is between +10 and 10 . This is because, as seen in Fig. 11, mass transfer is lower for the geometries at attack angles of +10 and 10 than for 0 and 5 . Despite energy losses being lower for the 4L03L case with an angle of attack of 10 than for 0 and 5 (see Fig. 9), the lower Sherwood number results in a lower permeate ux, which in turn increases the operating cost per cubic meter of permeate. Angles above +10 or below 10 are unlikely to reduce processing costs, as aerodynamic studies [36] suggest that ow separation would cause signicant increases to form drag, and hence to operating costs. Given that the low Sherwood numbers observed for the attack angles of +10 and 10 imply a low degree of mixing and boundary layer renewal, it is unlikely that these congurations will achieve lower processing costs for higher Schmidt number solutes. For the df /hch ratios of 0.4 and 0.3, the -L spacer congurations resulted in lower processing costs than the -H congurations. Therefore, it can be concluded that for 2D 4-layer zigzag spacers with elliptical laments as middle layers, the -L conguration will result in lower processing costs than the -H conguration, for the cost parameter conditions used in this paper. Since the attack angles which produced the lowest permeate processing cost for each spacer conguration also presented the highest Sherwood numbers, it can be said that mass transfer had the greatest impact in determining the permeate processing costs in steady ow conditions. The data collected suggests that positioning the middle elliptical lament layers such that they cause a change in the direction of the bulk ow (via positive angles of attack) will result in lower processing costs for the geometries analyzed, despite lower energy losses occurring when the laments are aligned with the ow (at negative angles of attack). However, more data points are needed before optimal operating angles of attack can be determined for these geometries. 3.3.2. Effect of changes in membrane cost When higher membrane costs per unit area are assumed, the difference between spacer-lled channels and empty channels becomes more evident. As seen in Fig. 20 for a membrane cost of $500/m2 , the permeate production cost when utilizing an empty channel is higher than for any of the spacer-lled channel geometries tested within the Reynolds number range of this study. This is because energy losses in the empty channel are up to an order of magnitude lower than for the spacer-lled channels, and therefore membrane costs account for most of the empty channel processing costs. In addition, as membrane costs are increased, the Reynolds

Fig. 19. Dependence of total processing cost per cubic meter of permeate on angle of attack of elliptical laments, for multi-layer spacer-lled channels at a hydraulic Reynolds number of 200. A membrane cost of $100/m2 was assumed.

820

G.A. Fimbres-Weihs, D.E. Wiley / Journal of Membrane Science 325 (2008) 809822

Reynolds number for the 2L06 spacer. This suggests that operating costs dominate in multi-layer spacers, mainly due to the increased pressure losses and form drag caused by the extra lament layers in the multi-layer congurations. In other words, a geometry with a larger amount of form drag will have a lower optimal operating Reynolds number, while a geometry with less form drag will have a higher optimal operating Reynolds number. However, the value of the minimum permeate processing cost is not solely dependent on the value of optimal Reynolds number. Therefore, form drag alone cannot determine which geometry will attain the lowest permeate processing costs. For this scenario, the best performing (lowest processing cost) spacer geometry is the 2L06 spacer, followed by the 3L04 at an angle of attack of 0 . The geometry with the highest processing cost at the optimal Reynolds number is, as in the previous scenarios, the empty channel.

Fig. 20. Dependence of total processing cost per cubic meter of permeate on hydraulic Reynolds number, for empty, 2-, 3- and 4-layer spacer-lled channels. A membrane cost of $500/m2 was assumed.

4. Conclusions A design study for novel multi-layer spacer congurations including elliptical lament spacer layers was conducted. Single and double layers of elliptical spacer laments were incorporated into 2D zigzag congurations of circular laments. The expected effect of the elliptical laments was to direct low concentration uid from the bulk to the near-wall region, and to increase the disruption of boundary layer formation caused by the circular laments. In order to assess the benets of this type of spacer conguration, an economic analysis was also carried out using cost parameters representative of those quoted in literature. The results presented in this work show that, although a comparison of different spacer geometries via power consumption (Re3 f) and mass transfer (Sh) is useful, an economic analysis incorporating permeate production provides a more realistic relative performance assessment, as it accounts for the trade-offs in other performance characteristics. Therefore, in order to determine which spacer geometries will yield lower permeate processing costs, an economic analysis is recommended over a comparison of mass transfer at similar energy losses. According to the economic analysis, the optimal operating Reynolds number (at which total processing costs were minimized for each spacer geometry) was found to be closely related to the amount of form drag generated by the spacer laments. For geometries with a higher amount of form drag the optimal Reynolds number was lower than for those with a lower amount of form drag. It was also found that for higher membrane costs, the optimal operating Reynolds number was increased. For the highest membrane cost assumed in this paper ($500/m2 ), the optimal operating Reynolds number fell in the unsteady ow regime. For low membrane costs of $10/m2 , operating in the laminar ow regime resulted in the lowest processing costs. One of the main components of processing cost which was not incorporated into the economic analysis carried out in this work was that of fouling and cleaning. Since this cost is usually proportional to the membrane area, it was therefore assumed to be included in the membrane cost. If fouling and cleaning costs are signicant, the analysis presented in this work suggests that increasing the operating Reynolds number will reduce total processing costs. The direct effect of multiple spacer layers on fouling was not studied in this paper. However, based on the study of Schwinge et al. [10] and on the results presented here, it is expected that the higher local velocities and shear rates at the membrane wall will reduce the extent of fouling. Another cost component not taken into account in the analysis carried out in this work was the cost of production of the novel spacer geometries. Although new techniques are available for the

number for minimum total cost is also increased. This effect is clearly seen for the 4L03H spacer geometries, for which the processing costs in the unsteady ow regime (i.e. at a hydraulic Reynolds number of 800), are lower than in the steady laminar regime. In contrast, at lower membrane costs, processing costs were higher at a hydraulic Reynolds number of 800 than at the lower hydraulic Reynolds number of 200 (see Fig. 16). Therefore, as membrane costs are increased, it becomes desirable from the economic viewpoint to operate in the unsteady ow regime. This is because the increased energy losses and associated costs of operating in the unsteady ow regime are easily offset by the higher mass transfer rates and permeate uxes attained at higher Reynolds numbers. Conversely, at the lower membrane cost of $10/m2 , the optimal operating Reynolds number for every spacer conguration is lower than the optimal Reynolds number at higher membrane costs. Fig. 21 shows that the advantage of spacer-lled channels against the empty channel is reduced at this lower membrane cost, and that the optimal Reynolds number for the empty channel is higher than the optimal Reynolds number for spacer-lled channels. The reason for this change is that at low membrane costs, energy costs become the main contributor to the total processing cost, and the total cost curve approaches the operating cost curve. Moreover, it can also be seen that the optimal Reynolds number for the multi-layer spacer congurations is lower than the optimal

Fig. 21. Dependence of total processing cost per cubic meter of permeate on hydraulic Reynolds number, for empty, 2-, 3- and 4-layer spacer-lled channels. A membrane cost of $10/m2 was assumed.

G.A. Fimbres-Weihs, D.E. Wiley / Journal of Membrane Science 325 (2008) 809822

821

production of new spacer designs [11,41], common mass produced spacers are currently available at much lower costs, which will have an impact on the economic results presented here. Incorporation of the capital costs of spacers would give an economic advantage to traditional 2-layer spacers over new multi-layer designs, which are more expensive to produce. Therefore, new designs must increase permeate production substantially in order to provide an economic incentive for their production. The multi-layer congurations with elliptical laments showed potential for increased productivity due to mass transfer enhancement. However, lower permeate processing costs were obtained by using a simple 2D zigzag spacer (2L06). The middle layers of elliptical laments in multi-layer spacer congurations promoted mass transfer enhancement when compared to 2-layer spacer lled channels with the same df /hch ratios. However, stagnant uid regions near the membrane walls resulted in lower mass transfer for the multi-layer spacer congurations than for the 2L06 and 2L04 spacers. Variation of the angle of attack of the elliptical laments showed potential to increase mass transfer enhancement at the same or lower energy losses by optimizing the prole of the middle spacer layer, such that little or no recirculation regions are formed on its downstream side. In addition, multi-layer spacer showed potential for reducing processing costs in high-membrane cost scenarios. Although negative attack angles usually reduced energy losses, they failed to direct ow towards the membrane wall, and thus resulted in lower mass transfer and hence higher operating costs per unit volume of permeate. On the other hand, positive angles of attack generally resulted in lower total permeate processing costs, due to increased mass transfer. Thus, mass transfer performance was the main factor in determining the total processing costs for the conditions of this study. As such, improvements in mass transfer enhancement have the potential to reduce both capital and operating processing costs per unit volume of permeate produced. This is possible even if energy losses are increased, since larger permeate uxes will reduce the operating cost per unit volume of permeate. With regards to the relative positioning of circular and elliptical laments in 4-layer spacers, it was found that the low (-L) conguration resulted in higher mass transfer enhancement, and hence lower permeate processing costs. 4L spacers in the -L conguration, coupled with positive angles of attack for the elliptical laments were found to be the most effective in disrupting ow and redirecting low concentration uid towards the membrane walls. However, further studies are required in order to optimize the hydrodynamic prole and angle of attack of the submerged layers, and thus reduce the energy losses incurred by the addition of further ow obstacles. Variation of the lament length to channel height ratio (lm /hch ) should also be explored. In addition, threedimensional simulations of real-world multi-layer spacer designs are needed in order to investigate the 3D ow and mass transfer effects that have not been taken into account in this paper. Current trends indicate that energy costs are increasing, and that membrane production costs are decreasing. This accentuates the need for improved spacer meshes which reduce energy losses in general, and detrimental form drag in particular, in order to reduce processing costs for membrane operations. Multi-layer spacer meshes represent a potential alternative to traditional 2-layer spacer meshes, but further optimization of geometric parameters and lament proles are required. Although none of the proposed multi-layer designs resulted in lower processing costs than traditional 2-layer congurations, especially at lower membrane costs, further study into the 3D ow and mass transfer effects of the middle layer spacer laments would provide invaluable information to either support or rule out multi-layer spacers as a real alternative to traditional spacer meshes.

Acknowledgements The authors would like to acknowledge the Australian Research Council for funding this project through a Discovery grant. One of us (G. F.-W) would like to thank the University of New South Wales and the Faculty of Engineering for scholarship funding.

Nomenclature ellipse minor axis (m) ae aE ellipse major axis (m) Am membrane area (m2 ) AT channel cross-section area (m2 ) A, A, B, B Empirical constants capital cost ($/m3 ) Cc Ce energy cost ($/kWh) Cm membrane cost ($/m2 ) Cop operating cost ($/m3 ) total permeate processing cost ($/m3 ) Ctot de equivalent diameter (m) df lament diameter (m) dh hydraulic diameter (m) D diffusion coefcient (m2 /s) dh p f = Fanning friction factor 2 L
2 u

amortization factor (1/yr) height to ellipse (m) channel height (m) ux of solution through the membrane (kg/m2 s) mass transfer coefcient (m/s) mesh length (m) channel length (m) specic permeate ux of membrane, permeability coefcient (kg/Pa m2 s) n distance in the direction normal to a surface (m) N, N Empirical constants p pressure (Pa) pch channel pressure drop (Pa) 3 Pn = Reh f Power number Q volumetric ow (m3 /s) Reh = ueff dh / hydraulic Reynolds number Sc = / D Schmidt number Sh = kmt dh /D Sherwood number Shloc = (dh /Yw Yavg )(Y /y)w local Sherwood number t time (s) top operating time (h/yr) u velocity in x direction (m/s) ueff = uavg / effective velocity (m/s) v velocity in y direction (m/s) Vtot volume of uid in the channel (m3 ) channel width (m) wch Ws pumping energy (W) x distance in bulk ow direction (m) x+ = L/dh Re dimensionless friction distance x* = L/dh ReSc dimensionless mass transfer distance y distance in direction perpendicular to bulk ow (m) Y salt mass fraction Greek letters void fraction, porosity pump pump efciency e angle of attack of elliptical lament ( ) dynamic viscosity of the uid (kg/m s)

Fa he hch Jsln kmt lm L Lp

eff

822

G.A. Fimbres-Weihs, D.E. Wiley / Journal of Membrane Science 325 (2008) 809822 [14] L. Zovatto, G. Pedrizzetti, Flow about a circular cylinder between parallel walls, J. Fluid Mech. 440 (2001) 125. [15] R. Iwatsu, K. Ishii, K. Kawamura, K. Kuwahara, J.M. Hyun, Numerical simulation of three-dimensional ow structure in a driven-cavity, Fluid Dynam. Res. 5 (1989) 173189. [16] R. Iwatsu, J.M. Hyun, K. Kuwahara, Analyses of three-dimensional ow calculations in a driven cavity, Fluid Dynam. Res. 6 (1990) 91102. [17] R.B. Bird, W.E. Stewart, E.N. Lightfoot, Transport Phenomena, Wiley, New York, 1960. [18] D. Wiley, D.F. Fletcher, Techniques for computational uid dynamics modelling of ow in membrane channels, J. Membr. Sci. 211 (2003) 127137. [19] D.F. Fletcher, D. Wiley, A computational uids dynamics study of buoyancy effects in reverse osmosis, J. Membr. Sci. 245 (2004) 175181. [20] A. Subramani, S. Kim, E.M.V. Hoek, Pressure, ow, and concentration proles in open and spacer-lled membrane channels, J. Membr. Sci. 277 (2006) 717. [21] N.V. Ndinisa, D.E. Wiley, D.F. Fletcher, Computational uid dynamics simulations of Taylor bubbles in tubular membranesmodel validation and application to laminar ow systems, Chem. Eng. Res. Des. 83 (2005) 4049. [22] V. Geraldes, M.D. Afonso, Generalized mass-transfer correction factor for nanoltration and reverse osmosis, AIChE J. 52 (2006) 33533362. [23] J. Schwinge, D. Wiley, D.F. Fletcher, Simulation of the ow around spacer laments between channel walls. 1. Hydrodynamics, Ind. Eng. Chem. Res. 41 (2002) 29772987. [24] J. Schwinge, D. Wiley, D.F. Fletcher, Simulation of the ow around spacer laments between channel walls. 2. Mass-transfer enhancement, Ind. Eng. Chem. Res. 41 (2002) 48794888. [25] J. Schwinge, D. Wiley, D.F. Fletcher, Simulation of unsteady ow and vortex shedding for narrow spacer-lled channels, Ind. Eng. Chem. Res. 42 (2003) 49624977. [26] H.-Z. Huang, W.-Q. Tao, An experimental study on heat/mass transfer and pressure drop characteristics for arrays of nonuniform plate length positioned obliquely to the ow direction, J. Heat Trans. T. ASME 115 (1993) 568575. [27] Z.-X. Yuan, W.-Q. Tao, Q.-W. Wang, Numerical prediction for laminar forced convection heat transfer in parallel-plate channels with streamwise-periodic rod disturbances, Int. J. Numer. Methods Fluids 28 (1998) 13711387. [28] N.R. Rosaguti, D.F. Fletcher, B.S. Haynes, Laminar ow and heat transfer in a periodic serpentine channel, Chem. Eng. Technol. 28 (2005) 353361. [29] S.W. Churchill, R. Usagi, A general expression for the correlation of rates of transfer and other phenomena, AIChE J. 18 (1972) 11211128. [30] F. Maskan, D.E. Wiley, L.P. Johnston, D.J. Clements, Optimal design of reverse osmosis module networks, AIChE J. 46 (2000) 946954. [31] O. Kedem, A. Katchalsky, Thermodynamic analysis of the permeability of biological membranes to non-electrolytes, Biochim. Biophys. Acta 27 (1958) 229. [32] U. Merten, Flow relationships in reverse osmosis, Ind. Eng. Chem. Fundam. 2 (1963) 229. [33] S. Sourirajan, Reverse Osmosis, Academic Press, New York, NY, 1970. [34] V. Geraldes, V. Semiao, M.N. de Pinho, Flow and mass transfer modelling of nanoltration, J. Membr. Sci. 191 (2001) 109128. [35] A.H.P. Skelland, Diffusional Mass Transfer, John Wiley & Sons, New York, 1974. [36] J.D. Anderson, A History of Aerodynamics, Cambridge University Press, Cambridge, 1997. [37] P. Feron, G.S. Solt, The inuence of separators on hydrodynamics and mass transfer in narrow cells: ow visualisation, Desalination 84 (1991) 137152. [38] G. Belfort, G.A. Guter, An experimental study of electrodialysis hydrodynamics, Desalination 10 (1972) 221262. [39] W.W. Focke, On the mechanism of mass transfer enhancement by eddy promoters, Electrochim. Acta 28 (1983) 11371146. [40] D.E. Wiley, C.J.D. Fell, A.G. Fane, Optimisation of membrane module design for brackish water desalination, Desalination 52 (1985) 249265. [41] J. Balster, I. Pnt, D.F. Stamatialis, M. Wessling, Multi-layer spacer geometries with improved mass transport, J. Membr. Sci. 282 (2006) 351361.

ratio of permeable to impermeable Sherwood number osmotic pressure (Pa) density of the uid (kg/m3 ) reection coefcient osmotic pressure coefcient (Pa) ratio of permeation Peclet number to impermeable Sherwood number

Subscripts avg average value integrated over channel cross-section b mass ow average or bulk ow value fd value at fully developed boundary layer conditions in value at the domain inlet loc value at the local ow conditions out value at the domain outlet p value for the permeate tm trans-membrane value w value at the channel wall

References
[1] D.G. Thomas, Forced convection mass transfer. Part II. Effect of wires located near the edge of the laminar boundary layer on the rate of forced convection from a at plate, AIChE J. 11 (1965) 848852. [2] D.G. Thomas, Forced convection mass transfer. Part III. Increased mass transfer from a at plate caused by the wake from cylinders located near the edge of the boundary layer, AIChE J. 12 (1966) 124130. [3] A.R. Da Costa, A.G. Fane, C.J.D. Fell, A.C.M. Franken, Optimal channel spacer design for ultraltration, J. Membr. Sci. 62 (1991) 275291. [4] R.E. Hicks, W.G.B. Mandersloot, The effect of viscous forces on heat and mass transfer in systems with turbulence promoters and in packed beds, Chem. Eng. Sci. 23 (1968) 12011210. [5] G. Schock, A. Miquel, Mass transfer and pressure loss in spiral wound modules, Desalination 64 (1987) 339352. [6] G.A. Fimbres-Weihs, D.E. Wiley, D.F. Fletcher, Unsteady ows with mass transfer in narrow zigzag spacer-lled channels: a numerical study, Ind. Eng. Chem. Res. 45 (2006) 65946603. [7] G.A. Fimbres-Weihs, D.E. Wiley, Numerical study of mass transfer in threedimensional spacer-lled narrow channels with steady ow, J. Membr. Sci. 306 (2007) 228243. [8] D.G. Thomas, W.L. Grifth, R.M. Keller, The role of turbulence promoters in hyperltration plant optimization, Desalination 9 (1971) 3350. [9] A.R. Da Costa, A.G. Fane, Net-type spacers: effect of conguration on uid ow path and ultraltration ux, Ind. Eng. Chem. Res. 33 (1994) 18451851. [10] J. Schwinge, D.E. Wiley, A.G. Fane, Novel spacer design improves observed ux, J. Membr. Sci. 229 (2004) 5361. [11] F. Li, W. Meindersma, A.B. De Haan, T. Reith, Novel spacers for mass transfer enhancement in membrane separations, J. Membr. Sci. 253 (2005) 112. [12] D.W. Kim, M.-U. Kim, Minimum drag shape in two-dimensional viscous ow, Int. J. Numer. Methods Fluids 21 (1995) 93111. [13] C.H.K. Williamson, Vortex dynamics in the cylinder wake, Ann. Rev. Fluid Mech. 28 (1996) 477539.

Potrebbero piacerti anche