You are on page 1of 9

Chemical Engineering Science 60 (2005) 5696 5704 www.elsevier.

com/locate/ces

A general correlation for pressure drop in a Kenics static mixer


Hyun-Seob Song , Sang Phil Han
Corporate R&D, LG Chem, Ltd./Research Park, 104-1 Moonji-dong, Yuseong-gu, Daejeon 305-380, Republic of Korea Received 17 May 2004; received in revised form 25 January 2005; accepted 21 April 2005 Available online 27 June 2005

Abstract A new pressure drop correlation in a Kenics static mixer has been developed. Pressure drop data were generated from computational uid dynamics (CFD) calculations, avoiding the experimental limitations in obtaining comprehensive data enough for developing a reliable pressure drop correlation. Dimensional analysis reveals that the pressure drop characteristic of the Kenics static mixer can be described by three dimensionless groups, i.e., the friction factor, Reynolds number (Re), and aspect ratio of a mixing element (AR). A systematic graphical analysis led to a single master curve governing the pressure drop behavior of the Kenics static mixer, which had never been achieved before. We derived a pressure drop correlation tting well with the obtained master curve in a general form into which the AR effect on the pressure drop is directly incorporated. Unlike the already existing correlations available in the literature, the correlation proposed in this study can cover the whole range of Re from laminar to turbulence. The reliability of the proposed correlation was validated by the comparison with various pressure drop data reported in the literature. 2005 Elsevier Ltd. All rights reserved.
Keywords: Pressure drop correlation; Kenics static mixer; Computational uid dynamics (CFD); Dimensional analysis; Fluid mechanics; Graphical analysis

1. Introduction Kenics static mixers are being widely used in the food and chemical industries for a variety of mixing applications such as single-phase homogenization, liquidliquid dispersion, or gasliquid dispersion, etc. A Kenics static mixer features a series of helical mixing elements within a long cylindrical tube. The helical blades are twisted alternatively in left- and right-hand directions by 180 and the successive ones are aligned at 90 each other. This peculiar conguration of the mixing elements allows the uids to be mixed in an effective way. The mixing principles can be explained by periodic splitting and remixing at the junctions and stretching and folding within the elements (Bakker et al., 1998). Until recently, a number of studies have been reported on the ow characteristics and mixing performance of the Kenics static mixers (Hobbs and Muzzio, 1998a,b; Hobbs et al., 1998;

Corresponding author: Tel.: +82 42 866 2666; fax: +82 42 861 2057.

E-mail address: songh@lgchem.com (H.-S. Song). 0009-2509/$ - see front matter 2005 Elsevier Ltd. All rights reserved. doi:10.1016/j.ces.2005.04.084

Byrde and Sawley, 1999; Rauline et al., 2000; Fourcade et al., 2001; Szalai and Muzzio, 2003; van Wageningen et al., 2004). Despite the various advantages of the Kenics static mixers mentioned above, its use can often be restricted due to the substantially large pressure drop caused by the helical elements. It might be envisaged that the mixing strength of the Kenics static mixer is achieved at the expense of high pressure drop. In this regard, it has been an important issue to make a reliable pressure drop correlation in the Kenics mixer. The benets of a priori estimating pressure drop are straightforward. In a designing stage, it is a prerequisite to calculate the power of the pump. Besides, pressure drop estimation is valuable in determining the possible operating window of the mixer. Since there are a host of parameters affecting pressure loss in the Kenics static mixer, it is convenient to represent pressure drop correlation in terms of fewer dimensionless groups. It is well known that the pressure drop correlation in an empty tube is simply given in terms of the friction factor (Cf ) and Reynolds number (Re). Dimensional analysis

H.-S. Song, S.P. Han / Chemical Engineering Science 60 (2005) 5696 5704

5697

of the Kenics static mixer conducted in this study shows that the element aspect ratio (AR), i.e., ratio of diameter to length of a mixing element, is another important dimensionless parameter to be considered in addition to Cf and Re. As will be shown later, Cf (and subsequently pressure drop) signicantly varies depending on AR, especially at high Re regime. Despite this important role of AR, its effect has not been properly considered in most of the existing correlations. Recently, a few groups reported more general forms for pressure drop correlation by explicitly incorporating the AR effect (Lecjaks et al., 1987; Joshi et al., 1995), but it is unfortunate that they cannot be applied to the high Re regime. The absence of the general pressure drop correlation applicable to the whole range of Re for the Kenics mixer motivates this study. A new pressure drop correlation for a Kenics static mixer has been developed in this study by taking the AR effect into account. A single master curve describing the pressure drop characteristics in the Kenics mixer is successfully obtained through a systematic graphical analysis. Unlike the already existing correlations, the application of the proposed one is not restricted to a certain range of Re, but covering the entire range of Re even over 106 . The prediction capability of the proposed correlation is conrmed by comparing the existing literature data.

Table 1 The geometric parameters of the Kenics static mixers Type A B C D (m) 0.00493 0.00810 0.01092 L (m) 0.292 0.432 0.629 AR 1.74 1.67 1.80

2. Grid dependence test The pressure drop data of the Kenics mixer in this study have been obtained from computational uid dynamics (CFD) calculations using a commercial code, FLUENT V6.0. In order to enhance the accuracy of the simulation results, we investigate here the dependence of the calculated pressure drop on the mesh density. A grid dependence test was conducted under the laminar condition (i.e., Re = 0.24) using the six-elements Kenics mixer. The mixer geometries, uid ow rate and uid properties taken for the grid test is as follows: inner diameter of the tube = 0.00493 m, the aspect ratio of a mixing element = 1.74; ow rate = 0.3 m/s; viscosity = 0.5 Pa s, and density = 800 kg/m3 . The geometric parameters employed here are the same as those of the Stratos Tube Mixers (See type A in Table 1). The mesh size can be controlled by adjusting the following two parameters. One is the number of the meshes on the cross-section ncross , and the other is the number of the grids on the axial length per element, nx (Fig. 1(a)). Fig. 1(b) shows how the pressure drop is changed with ncross and nx . The coarse grids (i.e., ncross = 60) give a relatively low value for pressure drop, but, as expected, the converged value is obtained as the grid density increases. Considering the tradeoff between the accuracy and the computational burden, we determined ncross and nx to be 224 and 12, respectively, which corresponds to the grid density of 1.6 1010 /m3 . The resulting grid allocation is constantly maintained throughout this study, not only for the calculation of lami-

Fig. 1. Grid dependence test: (a) denition of the two parameters, ncross and nx , controlling the grid density; (b) pressure drop variation depending on ncross and nx .

nar ows where the grid dependence tests were conducted, but also for the calculation of turbulent ows. The suitability of the determined grids for the turbulent regime is addressed here. For calculating the turbulent ow, we have employed the standard k . model with the standard wall function, which requires that the cells near the wall should meet the y + requirements. Theoretically, the location of the cells near the wall should be adapted for the loglaw to be valid at every Re. However, it has been found that the effect of the location of the rst cells from the wall and the type of the wall function on the pressure drop, which is the only variable we are concerned with here, is negligible in our calculations. From this nding, no

5698

H.-S. Song, S.P. Han / Chemical Engineering Science 60 (2005) 5696 5704

adaptation on the near-wall cells depending on the ow regime was made. The thin-wall assumption is introduced on the helical blade, i.e., the thickness of the helical blade has not been considered in the CFD calculations. This assumption sounds reasonable as the thickness of the blade is relatively small in comparison with the tube diameter.

3. The friction factor vs. the Reynolds number It seems exceedingly difcult to directly derive the pressure drop correlation since the number of the relevant process variables is so large in the Kenics mixer. This cumbersome situation can be alleviated through the dimensional analysis. The dimensional analysis is based on the so-called pi theorem, the essence of which lies in the compression of the size of the relevant physical quantities into fewer dimensionless groups (Zlokarnik, 2002). The detailed procedure of applying dimensional analysis is explained in the Appendix and only the principal results are addressed here. The pressure drop ( P ) can be expressed as a function of the following six process variables: f ( P , D, Lp , Le , , , u) = 0, (1)

where D is the inner diameter of the pipe, Lp the length of the pipe, Le the length of a mixing element, the uid viscosity, the uid density, and u the uid ow rate. The application of dimensional analysis transforms Eq. (1) to a relation among four dimensionless groups: g(Eu, Lp /D, AR, Re) = 0, (2)

where Eu = P /( u2 ), AR = Le /D , and Re = D u/ . Further, it is assumed that the Eu and Lp /D can be combined into Cf as often done in an empty tube. The question on the validity of this assumption will be clear later. Consequently, we have three dimensionless parameters only, i.e., Cf , Re, and AR, in terms of which the pressure drop correlation will be represented. Three different types of Kenics static mixers are considered for the pressure drop calculations, which are distinguished from one another in geometry such as tube diameter, tube length, and element aspect ratio as shown in Table 1. The mixers of Table 1 are among the widely used Koplo Stratos Tube Mixers, which have been used in our laboratory. For the time being, we conne our concern to elucidating the basic relation between Cf and Re with AR xed to a certain value, deferring the issue of incorporating AR into the correlation to later. Through investigating the Cf .Re relation, we can check the validity of the earlier assumption that Cf can replace Eu and Lp /D , and more importantly, get an insight into the pressure drop feature of the Kenics static mixer. The change of pressure drop with process parameters for the type A of Table 1 is presented in Fig. 2 where the

Fig. 2. The relation of Cf and Re for the Kenics static mixer of type A in Table 1 when (a) ow rate; (b) uid viscosity in addition to (a); (c) tube length in addition to (b) is varied.

H.-S. Song, S.P. Han / Chemical Engineering Science 60 (2005) 5696 5704

5699

Fig. 3. The Cf .Re curves calculated using laminar and standard k . models.

in the Kenics static mixer weakly depends on the viscosity enhancement by turbulence. It is conceivable that the resistance of the uid ow is mainly by the mixing elements while the contribution of the turbulence to the pressure drop is relatively small. In principle, the turbulence ows obey the NavierStokes (NS) equations. We could obtain the accurate simulation results without applying turbulence modeling by directly solving the unsteady NS equations using the direct numerical simulation (DNS). It is required that the grid in DNS is dense enough to capture the smallest motion of uid. In this regard, it might be expected that if the sufciently ne grids are used, the calculated pressure drop values could be similar between the case where the NS equations are solved with turbulence models and the case where only the NS equations are solved (as in laminar ow simulation), which partly explains the above observations. From all of these considerations, it is concluded that the employment of the standard k . model is a reasonable choice for calculating the pressure drop in the Kenics static mixer. It seems unlikely that the accuracy of the pressure drop data is improved simply by employing more advanced turbulence models. 4. Incorporation of the AR effect So far, we have examined the relation between Cf and Re when the AR is xed to be 1.74 (see Table 1). Now we are in a position to incorporate the AR effect into the correlation. We extend the pressure drop calculations to the different AR. The type B and type C of Table 1 might be good examples for this purpose. The AR effect was, however, not clearly observed in those mixers although the Cf values tend to scatter to some extent at high Re (Fig. 4(a)). In order to see more vividly the AR effect, we articially make the variation of AR wider from 1.0 to 2.5 for the Kenics mixer type A. As Fig. 4(b) shows, the effect of AR on the pressure drop is substantial when Re is large. The necessity of incorporating AR into the pressure drop correlation is obvious from this gure. In order to obtain a general pressure drop correlation including the AR effect, we need to obtain a single master curve describing the pressure drop behavior in the Kenics static mixer. As seen in Fig. 4(b), the transition point from downhill to plateau in the curves also varies with AR. If we dene the Re to be the x-coordinate of the intersection point of the two asymptotic lines as in Fig. 5(a), Re is the function of AR. Taking the logarithm of AR and Re provides a linear relationship between them (Fig. 5(b)), i.e., log(Re ) = 2.15 log(AR) + 2.09. (3)

value of AR is xed to be 1.74. Fig. 2(a) shows the relation between Cf and Re where the ow rates are varied from 0.0017 to 13 kg/s while the uid properties are constant, i.e., = 800 kg/m3 and = 0.2 Pa s. Cf linearly decreases with Re when Re is low, but becomes constant as Re becomes larger, which is the typical pressure drop curve as observed in an empty tube. The Cf .Re curves coincide with each other even when the material (e.g., viscosity) and geometrical parameters (e.g., the pipe length) are varied (Figs. 2(b) and (c)). These results indicate that the pressure drop characteristics are represented by the single curve at least when AR is xed, consequently justifying the foregoing assumption on the introduction of Cf . Before proceeding further, we address here in brief the issue on the choice of ow model (i.e., laminar or turbulence) in calculating the pressure drop of the Kenics static mixer. While the uid ow is laminar at Re less than 2300 in an empty tube, it is obscure to dene such critical Re in the Kenics mixer because the ow nature of the latter is substantially different from that of the former due to the mixing elements. Even if we could nd the critical Re in the Kenics static mixer, it is still unclear which ow model should be used in the transition regime between laminar and turbulence. In Fig. 3, Cf values are plotted against Re using the laminar and the turbulence models, respectively, over the whole range of Re where the standard k . model is employed for turbulent ow calculations. Contrary to the authors expectation, both curves coincide well with each other. It is understandable that the difference between the two curves is negligible at low Re regime where the turbulence strength is insignicant. Interestingly, the two curves do not show appreciable bifurcation even up to high Re range considered here as shown in Fig. 3, indicating that the pressure drop

This indicates that the x-coordinates of the transition points coincide with each other if we set the x-axis to be Re/AR2.15 (Fig. 5(c)). Once the x-axis is redened to be Re/AR2.15 , interestingly the distance between any two curves is independent of the x-coordinate. If we take the Cf values along the

5700

H.-S. Song, S.P. Han / Chemical Engineering Science 60 (2005) 5696 5704

Fig. 4. The Cf .Re curves for different AR values: (a) AR = 1.74, 1.67, and 1.80 (as in Table 1); (b) AR = 1.0, 1.5, 2.0, and 2.5.

on AR: (a) denition of Re ; (b) linear relation between log(Re ) and log(AR); (c) denition of C ; (d) linear Fig. 5. Dependence of Re and Cf f ) and log(AR). relation between log(Cf

H.-S. Song, S.P. Han / Chemical Engineering Science 60 (2005) 5696 5704

5701

x-axis into three regimes depending on the Re/AR2.15 value, i.e., (i) 0 < Re/AR2.15 < 100, (ii) 100 < Re/AR2.15 < 1000, and (iii) 1000 < Re/AR2.15 . Each regime is assumed to be tted well by the prescribed function below: Cf AR2.04 = K(Re/AR2.15 )n . (5)

By taking logarithms on both sides of Eq. (5), we obtain the linear equation with slope of n and y-intercept of log K . The optimal values of n and K are summarized in Table 2. 5. Comparison with literature data The performance of the pressure drop correlation derived above was compared with the literature data. Table 3 shows the pressure drop and the corresponding Cf values at Re of 10 for the case of AR of 1.5. The pressure drop prediction by the proposed correlation is in agreement with the literature average within about 25% discrepancy under the specied condition, which is reasonably acceptable. In order to verify the validity of the proposed correlation over a wider range, we compared it with various correlations available in the literature. The pressure drop correlations reported so far are well summarized in Rauline et al. (1998). It should be pointed out that the AR effect on the pressure drop has not been explicitly reected in most of the existing correlations as shown in Table 4, i.e., they only describe the relations between Cf and Re when AR is xed. The correlations by Lecjaks et al. (1987) and Joshi et al. (1995) include the AR effect, but their applications are limited to low Re range only.
Table 3 Cf and

Fig. 6. A single master curve representing the pressure drop characteristics of the Kenics static mixers.

Table 2 The optimal parameter values obtained from curve-tting of Eq. (5) 0 < Re/AR2.15 < 100 K n 320 0.86 100 < Re/AR2.15 < 1000 32.0 0.36 1000 < Re/AR2.15 2.66 0

constant x-line (e.g., Re/AR2.15 = 104 ) and take logarithm and AR, this time we obtain the linear relationship of Cf ) and log(AR) (Fig. 5(d)): between log(Cf
) = 2.04 log(AR) + 0.42. log(Cf

P over the six mixing elements at Re = 10 (AR = 1.5) Cf 44.9 48.2 44.7 38.9 26.0 75.0 32.0 44.2 32.8 P 20.2 21.7 20.1 17.5 11.7 33.8 14.4 19.9 14.8

(4)

References Grace (1971) Wilkinson and Cliff (1977) Pahl and Muschelknautz (1982) Lecjaks et al. (1987) Shah and Kale (1991) Joshi et al. (1995) Bakker et al. (1998) Literature average This study

AR2.04 to be constant, and Eq. (4) implies the quantity Cf thus, if we reset the y-axis to be Cf AR2.04 , then we can nally make the four curves into a single one (Fig. 6). To the best of our knowledge, we are the rst to nd a single master curve for the pressure drop correlation in the Kenics static mixer under the variations of the full parameters. Based on the preceding results, we propose a pressure drop correlation in the Kenics static mixer. We divide the

Table 4 Pressure drop correlations in the literature Authors Cybulski and Werner (1986) Grace (1971) Wilkinson and Cliff (1977) Sir and Lecjaks (1982) Lecjaks et al. (1987) Joshi et al. (1995) Correlations P/ P/ P/ P/ Cf = (213.5 + 224.0/AR)/Re + 4.775/AR 0.549 (15 < Re < 1200) P / P0 = 7.41 + 1.04Re0.8 /AR1.04 (Re < 700) P0 = 5.4 + 0.68Re0.5 P0 = 4.86 + 0.68Re0.5 P0 = 7.19 + Re/32 P0 = 5.34 + 0.0211Re AR 1.36 1.5 1.5 2.0 2.05.0 1.52.5

5702

H.-S. Song, S.P. Han / Chemical Engineering Science 60 (2005) 5696 5704

One of the possible reasons for this might lie in the fact that the Kenics static mixers have customarily been employed under the laminar condition, which is, however, not the case in modern applications. The difculty in obtaining experimental data for various operating conditions might be another important reason explaining the limited application range of the previous correlations. This problem is much more serious when the number of parameters to be varied are large as in the Kenics mixers, which was overcome by employing CFD calculations here. As shown in Fig. 7, the Cf predictions by the proposed correlation are quite close to others, but only ours covers up to the high Re range. It is found that the prediction by Joshi et al. (1995) is relatively high as pointed out in their own paper.

6. Conclusions A new correlation for pressure drop has been derived for a Kenics static mixer. Thanks to CFD calculations, comprehensive pressure drop data could be obtained over a wide range of process parameters. From the application of dimensional analysis, it is found that the pressure drop correlation can be represented in terms of three dimensionless groups, i.e., Cf , Re, and AR. Through systematic graphical analysis, we successfully obtain a single master curve describing the pressure drop characteristics in the Kenics static mixer. The proposed pressure drop correlation is distinguished from the existing correlations by the fact that it applies to the whole range of Re from laminar to turbulence, as well as that the AR effect is directly incorporated into the correlation. The comparison with other correlations shows the proposed correlation to be quite satisfactory.

Appendix. Derivation of PI groups The size of the physical quantities can be compressed by dimensional analysis. According to pi theorem, every physical relationship among n physical quantities can be transformed to a relationship among m (less than n) mutually independent dimensionless groups which are called pi groups in the dimensional analysis. The difference between n and m is often equal to the fundamental dimensions associated with the parameters in question. The detailed theoretical aspects of dimensional analysis and the abundant application examples in chemical engineering are available in Zlokarnik (2002). Dimensional analysis begins with dening the target quantity and listing all the parameters affecting it, which comprises the so-called relevance list. We assume the pressure drop ( P ), which is the target quantity here, is affected by the diameter and length of the pipe (D, Lp ), element length (Le ), material properties ( , ), and uid ow

Fig. 7. Comparison of the proposed pressure drop correlation with existing literature correlations: (a) AR = 1.36; (b) AR = 1.5; (c) AR = 2.

H.-S. Song, S.P. Han / Chemical Engineering Science 60 (2005) 5696 5704 Table A.1 The dimensions and units of the parameters of a Kenics static mixer Quantity Pressure drop Pipe diameter Pipe length Element length Fluid viscosity Fluid density Fluid velocity Symbol P D Lp Le Dimensions M/LT 2 L L L M/LT M/L3 L/T Units in SI kg/m s2 (=Pa) m m m kg/m s (=Pa s) kg/m3 m/s Table A.2 The core variables and the corresponding pi numbers Core vars. P Lp Le a 1 0 0 0 c 0 1 0 0 d 0 0 1 0 e 0 0 0 1 numbers

5703

= P /( u2 ) = Eu = Lp /D = Le /D = AR = /(D u) = 1/Re, i.e., 4 = Re


1 2 3 4

rate (u): f ( P , D, Lp , Le , , , u) = 0. (A.1)

The thickness of the mixing element is not included in the relevance list, assuming that its effect is insignicant. The number of the mixing elements is also excluded from the list because it can be calculated when Lp and Le are given. As seen in Eq. (A.1), we employed the pipe diameter (D) as the characteristic length of a Kenics static mixer, instead of the hydraulic mean diameter (Dh ). The dimensions and units of the relevance list are summarized in Table A.1. Now, we seek the dimensionless numbers describing the pressure drop in a Kenics static mixer. The dimensionless term is called a (pi) group, which has the general form as follows:
d = ( P ) a D b Lc p Le e f g

two rules: (1) all the dimensions of the system must be represented in the core variables; (2) the core variables must not form a dimensionless group (Duncan and Reimer, 1998). It is convenient to designate the target quantity (i.e., P ) as a core variable. For the other core variables, three parameters commonly varying in the experiments are chosen: Lp , Le , and . Clearly, these four variables satisfy the rst rule. The second rule implies we must choose a core variable for each pi group, e.g., if a pi number includes P , then Lp , Le , or should not be there. We can satisfy this requirement by properly setting the exponent of the core variables, i.e., c, d, and e should be zero in the example. We are free to choose any number for the exponent on P . It is usually convenient to choose 1. Substituting a = 1 and c = d = e = 0 into Eqs. (A.4)(A.6) and solving them, we obtain b = 0, f =1, and g = 2. The pi number resulting from this solution set is the Euler number (Eu):
1

u .

(A.2)

= P /( u2 ) = Eu.

(A.8)

can be expressed in terms of basic dimensions as listed in Table A.1: M a b c d M e M f L g L L L LT L3 T LT 2 a +e+f a +b+c+d e3f +g 2a eg [=]M L T . [=]

(A.3)

to be dimensionless, the exponent on each term of For Eq. (A.3) should be zero, which leads to Mass, M : Length, L: Time, T : a + e + f = 0, a + b + c + d e 3f + g = 0, 2a e g = 0. (A.4) (A.5) (A.6)

The other dimensionless groups are derived in a similar way, resulting in pipe aspect ratio, element aspect ratio (AR), and Re as summarized in Table A.2. The pi numbers for the pressure drop correlations in the Kenics static mixer and in the empty tube coincide with each other except the AR. It is well known that the pressure drop correlation in the empty tube is represented in terms of the friction factor (Cf ) and Re, the former being a combination of Eu and the pipe AR as given below: Cf 2Eu D 2 P D = . Lp u2 Lp (A.9)

According to the pi theorem, the seven physical quantities of Table A.1 are compressed to the four dimensionless groups by subtracting the number of independent equations from the number of unknowns (Zlokarnik, 2002). This signies the pressure drop correlation can be expressed in terms of the four pi numbers as shown in Eq. (A.7), which is much simpler than the original form as given in Eq. (A.1). g(
1, 2, 3, 4 ) = 0.

Beneting from this result in the empty tube, we reduce the number of pi groups from four to three by replacing 1 and 2 with Cf . The purpose of this study is, therefore, to nd the functional relationship among Cf , Re, and AR. References
Bakker, A., LaRoche, R.D., Marshall, E.M., 1998. Laminar ow in static mixers with helical elements. Published in The Online CFM Book at http://www.bakker.org/cfm Byrde, O., Sawley, M.L., 1999. Optimization of a Kenics static mixer for non-creeping ow conditions. Chemical Engineering Journal 72, 163169. Cybulski, A., Werner, K., 1986. Static mixerscriteria for applications and selection. International Chemical Engineering 26, 171180.

(A.7)

Now, what remains is how to determine the four pi numbers. One of the possible ways to nd the pi groups is to introduce the core variables. The core variables should satisfy

5704

H.-S. Song, S.P. Han / Chemical Engineering Science 60 (2005) 5696 5704 Rauline, D., Tanguy, P.A., Le Blevec, J.-M., Bousquet, J., 1998. Numerical investigation of the performance of several static mixers. Canadian Journal of Chemical Engineering 76, 527535. Rauline, D., Le Blevec, J.-M., Bousquet, J., Tanguy, P.A., 2000. A comparative assessment of the performance of the Kenics and SMX static mixers. Chemical Engineering Research and Design 78, 389396. Shah, N.F., Kale, D.D., 1991. Pressure drop for laminar ow of nonNewtonian uids in static mixer. Chemical Engineering Science 46, 21592161. Sir, J., Lecjaks, Z., 1982. Pressure drop and homogenization efciency of a motionless mixer. Chemical Engineering Communication 16, 325334. Szalai, E.S., Muzzio, F.J., 2003. Fundamental approach to the design and optimization of static mixers. A.I.Ch.E. Journal 49, 26872699. van Wageningen, W.F.C., Kandhai, D., Mudde, R.F., van den Akker, H.E.A., 2004. Dynamic ow in a Kenics static mixer: an assessment of various CFD methods. A.I.Ch.E. Journal 50, 16841696. Wilkinson, W.L., Cliff, M.J., 1977. An investigation into the performance of a static in-line mixer. Second European Conference on Mixing, A2-15A2-29. Zlokarnik, M., 2002. Scale-up in Chemical Engineering. Wiley-VCH, New York.

Duncan, T.M., Reimer, J.A., 1998. Chemical Engineering Design and Analysis: An Introduction. Cambridge University Press, Cambridge. Fourcade, E., Wadley, R., Hoefsloot, H.C., Green, A., Iedema, P.D., 2001. CFD calculation of laminar striation thinning in static mixer reactors. Chemical Engineering Science 56, 67296741. Grace, C.D., 1971. Static mixing and heat transfer. Chemical Processing and Engineering 52 (7), 5759. Hobbs, D.M., Muzzio, F.J., 1998a. Optimization of a static mixer using dynamical systems techniques. Chemical Engineering Science 53, 31993213. Hobbs, D.M., Muzzio, F.J., 1998b. Reynolds number effects on laminar mixing in the Kenics static mixer. Chemical Engineering Journal 70, 93104. Hobbs, D.M., Swanson, P.D., Muzzio, F.J., 1998. Numerical characterization of low Reynolds number ow in the Kenics static mixer. Chemical Engineering Science 53, 15651584. Joshi, P., Nigam, K.D.P., Nauman, E.B., 1995. The Kenics static mixer: new data and proposed correlations. Chemical Engineering Journal 59, 265271. Lecjaks, Z., Machac, I., Sir, J., 1987. Pressure loss in uids owing in pipes equipped with helical screws. International Chemical Engineering 27 (2), 205209. Pahl, M.H., Muschelknautz, E., 1982. Static mixers and their applications. International Chemical Engineering 22, 197205.

You might also like