Gas/liquid dispersions with a SMX static mixer in the laminar regime

https://doi.org/10.1016/j.ces.2005.12.022Get rights and content

Abstract

A Sulzer SMX mixer was used to disperse gas into viscous, Newtonian and non-Newtonian fluids. The investigation covered the effect of the dispersed phase volume fraction, the viscosity of the continuous phase, the mixer length and the power draw. The flow regime was kept laminar in all the experiments. The dispersion of gas was carried out with gas concentrations between 1% and 7% in volume. Using the “process viscosity” concept, it was possible to collapse all the measured sizes on a single master curve by using the energy consumption in the mixer as the common variable between the experiments. Comparison was made with a Kenics mixer. The SMX mixer was found to be better adapted to the dispersion task due to its internal structure.

Introduction

The process of dispersing a fluid (liquid, gas) or a solid into a liquid continuous phase is probably one of the most widely spread mixing process in the industry. Despite this popularity, the equipment used for dispersion in continuous laminar processes are not that well known and the knowledge is still, in large parts, owned by the equipment builders. There exist a certain number of studies reporting capacity augmentation, higher process efficiencies, or reaction yield when using static mixing devices with two or more fluid phases in laminar regime (e.g., Schneider et al., 1988; Stringaro and Luder, 1991; Yeon and Choi, 1996). This work is aimed at quantifying the dispersion capacity of the SMX mixer (Sulzer) in the laminar regime and compare it to the performance of other mixers.

Most of the work dealing with laminar dispersion whether is based on Grace (1982) or makes direct references to it. The center of his work is the variation of the critical capillary number for dispersion of drops as a function of the viscosity ratio of the phases. The capillary number is defined as the ratio of the Weber number to the Reynolds number namely:Ca=ηuσ=τ(σ/L).From the dispersed phase point of view, this number can be seen as the ratio of the “destructive” viscous forces responsible for the rupture of the secondary phase over the “protective” surface tension forces opposing the deformation. The inertial forces do not dominate due to the laminar flow regime. In this expression, η is the continuous phase viscosity, u a characteristic velocity and σ is the interfacial tension. The results of Grace were generated for a simple shear flow (Couette apparatus) and for a 2D extensional shear field (four rolls). All these results were obtained for one drop in steady state deformations, which necessitate extremely long times to reach equilibrium. While being very instructive about the break-up mechanisms fundamentals, these experiments do not represent industrial static mixer conditions. Static mixers intrinsically generate transient conditions while the capillary number study by Grace made use of steady state conditions only. The fact that only one drop is studied is also a strong deviation from the industrial situation where billions of drops can be present and interact with each other.

At lower viscosity ratios p<0.1(p=ηdispersed/ηcontinuous), in a simple shear field, very high stresses must be generated in order to break a bubble or a drop. With viscosity ratios above 3.0, steady simple shear flows cannot generate stresses high enough to lead to break-up. With extensional shear flows however, the capillary numbers remain within reasonable values for a very wide range of viscosity ratios. According to these results, flow fields with strong extensional components are preferred for dispersion applications (Rallison, 1984).

The flow type can be defined as the ratio of the deformation (D) to the rotational component (Ω) of the tensor:α=|D||D|+|Ω|whereD+Ω=12(v+vT)+(v-vT),α is helpful in representing the type of two-dimensional flow used: 0 = pure rotational; 0.5 = simple shear; 1 = pure extensional. Depending on the complexity of the geometry used, any intermediate value of α is possible. While this parameter is not objective, that is it cannot be used to compare totally different geometries such as an agitated tank and a static mixer (Astarita, 1979), it is still useful to compare similar geometries like two different static mixer designs or two mixing tanks against each other (e.g., De La Villéon et al., 1998).

The viscosity of the dispersed phase is a major parameter in the dispersion phenomenon. In the turbulent regime, it has been shown that the width of the bubbles/drops distribution increases with the viscosity ratio (Berkman and Calabrese, 1988; Middleman, 1974). Recently, there is some work reporting dispersion results in the turbulent regime (Legrand et al., 2001) using static mixers. The Reynolds number (Re=ρv¯D/μ) is part of every size correlation in this regime, generally of the form: D=D(Rea). The exponent a ranges between 0.1 and 0.2 with variations due to the mixer type.

The laminar regime was theoretically treated by using the energy repartition within a swirl (Middleman, 1974). At this length scale, the drop environment is limited to the movement inside a single swirl and is therefore laminar. The relation is of the formD32Dt=C2σμv3Dvρμ.According to this relation, the average diameter D32 is inversely proportional to the viscosity of the dispersed phase and the average velocity of the flow (1/μ4 and 1/v2). The interfacial tension also plays a major role (σ3). There are no available experimental data to validate this theory.

It is quite clear that the dispersion process is rather complex and it is also evident that the laminar regime is the least popular candidate for this type of process. To our knowledge, there exists only one work on laminar dispersion in a static mixer (Grace, 1982) in which only the Kenics mixer was considered.

The behavior of the size reduction in the mixer can be explained by a kinetic phenomenon. With the dispersed phase passing through the obstacles formed by the element, the drops break in a repetitive process until the kinetic equilibrium is reached between breakup and coalescence. The point where this equilibrium is reached is a function of the dissipated energy and the liquid or gas system involved. The mixing length to establish the equilibrium can be directly compared to the mixing time in tanks (Meyer et al., 1988; Villermaux and Falk, 1994). During the design process of a new application, it is hence of prime importance to evaluate the required number of elements properly.

Grace, in addition to the investigation of the break-up of single drops, used static mixers in his experiments with unsteady flows. His objective was to build a parallel between abrupt changes in flow conditions and the repetitive transient conditions met in static mixers. He showed that variations in the shear rate could lead to the breakup of drops and bubbles without even reaching the critical capillary number. This finding is very instructive about the capacity of a static mixer to generate an efficient size reduction.

The effect of the number of mixing elements can be modeled with the following (Middleman, 1974):D32D32(Nmax)=C1+C2e(-N*C3).Here, the value of D32(Nmax) is simply the minimum size achieved with the maximum number of elements used in the experiments. The values of constants C1C3 have been calculated here based on the results generated in a Kenics (Middleman, 1974) and an in-liner mixer from Lightnin (Al-Taweel and Walker, 1983). They are presented in Table 1. The in-liner mixer is very close in design to the Kenics mixer with their difference laying in the lamellae that are composed of small and partial twisted lamellas instead of an entire 90 or 180 twist. The values of 0.19 vs. 0.40 shown in the same table are also of significant importance for their dispersion capacity characterization and will be discussed later in this paper along with the SMX results.

Based on the principle that the Kenics mixers leads to an interface generation proportional to 2N (N the number of elements) within the laminar flow regime, an experimental investigation shows that this relation is true only after a minimum number of elements (Bigio and MacLaren, 1991). This minimum number of elements would increase with the volume fraction of dispersed phase and the viscosity ratio. Neither details nor data are provided in the paper.

The dispersion of a fluid into a second one requires some sort of an injection point where the first contact occurs between the phases. It is clear that the simplest way to inject a phase into a second one is to use a T connection. This configuration should be avoided (Grace, 1982). When no other alternative is available, a higher number of elements could be used to compensate for the unfavorable injection conditions on a T-connection.

Using simulations (Rauline et al., 1998), it was demonstrated that the most favorable location for the injection is the center of the pipe. Zalc came to the same conclusion numerically with a paper specifically for the SMX mixer (Zalc et al., 2003). Certain mixers appear to be much more sensitive to the injection type and location and the equipment builders can usually provide their injectors design for a given mixing application. According to Rauline et al., the SMX and the Ross ISG mixers should be the least affected by the location of the injection. The explanation laying on the strong extensional field they induce compared to other types of mixers.

In the laminar regime, the flow rate can be linked to the shear rate in the mixer and is called the effective or characteristic shear rate (Grace, 1982):γ˙eff=Cacritσ(Dd/2)1ηc.The characteristic shear rate should be seen as a property of the mixer rather than a design parameter. It is indeed back-calculated from the experimental drop/bubble size distributions. Once it is known for a given mixer, the characteristic shear rate can be of precious help in mixers selection.

In a Kenics mixer, Grace (Grace, 1982) was able to relate the Sauter mean diameter to the nominal wall shear rate expressed by means of the average velocity in the empty tube v, its diameter Dt, namely:D32=C1γ˙wc.In this expression, exponent c=1 when the viscosity ratio is below 1 and c=0.5 when the viscosity ratio is above 1. The constant C1 appears to be a function of the number of elements, the size of the tube, and the dispersed phase volume fraction. The work of Grace did not explicitly mentioned the fact that C1 must also be a strong function of the mixer geometry.

Only a single work pertains to the dispersion in SMX mixers and it comes from Sulzer itself (Mutsakis et al., 1986). They present critical capillary numbers evaluated in the Newtonian case with viscosity ratios below 1.0. The critical capillary numbers all fall between the curves of pure elongation and shear from Grace (1982) and Bentley and Leal (1986). Their experimental curve would qualify the SMX as predominantly extensional with a small shear contribution (Fig. 1). It has been shown numerically (Rauline et al., 1998) that the SMX creates a velocity field with an α of 0.6. This value is closer to shear than to elongation (α=1.0). However, this is an average value that does not reflect the local phenomenon governing the dispersion results (Villermaux, 1988). The distribution curves obtained for the critical capillary number as a function of the viscosity ratio are all obtained for stable conditions, two dimensional flows, and steady state. In the work from Sulzer these conditions were obviously never met. The flow in any static mixing device is periodic, tri-dimensional, non-symmetrical, and the shear stress is never constant. Therefore, the SMX critical capillary number values falling between the elongation and simple shear is hard to justify.

As seen above, the flow character (α) is constantly changing along the length of the element in a SMX. It was clearly demonstrated that those variations in the flow character are very important and can be very different from one mixer type to the next (Rauline et al., 1998). From the simulations, it is shown that the SMX mixer is creating many more and larger fluctuations along its flow axis while, in a Kenics mixer, the variations are mostly created at the inlet and outlet of each element. This type of result is very instructive about the flow conditions generated inside the mixer and helps understanding why some areas of the internal structure of the mixer are better at reducing the size of the dispersed phase. The effects of the structure of the SMX mixer on mixing were studied in depth by Zalc, both numerically and experimentally with impressive agreement between the two techniques (Zalc et al., 2002). Similar results were obtained numerically in Kenics mixers by following the deformation of massless and dimensionless particles launched from the entrance of the mixer (Avalosse and Crochet, 1997a, Avalosse and Crochet, 1997b). A very recent work also puts in evidence that the entrance and exit effects are of prime importance with the SMX and a new type of static mixer known as the KMX mixer (Heniche et al., 2003). Moreover, the shape of the blades used to compose the mixer have a strong influence on the flow character inside the mixer and hence, its capacity to disperse effectively (Heniche et al., 2004).

The transitional effects were investigated from a much more fundamental point of view in the journal bearing flow (Tjahjadi and Ottino, 1991). When certain flow conditions are met, the flow inside the apparatus can be considered as chaotic and hence, induces the ideal conditions for mixing (Ottino, 1989). The journal bearing flow field can be described analytically while it is totally unfeasible for the SMX mixer due to its intricate geometry. However, experiments clearly demonstrated that the flow field inside the SMX is chaotic (Li et al., 1996). The size distributions generated in the journal jearing apparatus showed narrower distributions with viscosity ratio (p) below 1.0 while, with viscosity ratios above 1.0, the average diameters were smaller. With p<1.0 as it is always the case in gas dispersion, the rupturing process consisting of breaking a single large drop into many smaller ones can be repeated many times until the interfacial tension stops the size reduction process, leading to narrower bubble size distributions.

The very complete review by Thakur et al. (2003) about static mixers in processes also puts in evidence the lack of knowledge about dispersion in laminar regime. This is even truer with the SMX mixer while the Kenics mixer is more the subject of such work.

After this review pertaining to the dispersion in general, and to the SMX in particular, it becomes clear that there is a serious lack of information about the SMX mixer and its performance in laminar dispersion. The objective of this work is hence to characterize experimentally the SMX performance at dispersing gas within a liquid in the laminar flow regime. Two cases will be considered: Newtonian continuous phase and non-Newtonian or shear thinning phase.

Section snippets

Experimental

The experimental bench is schematized in Fig. 2. It comprises three basic components: the circulation loop, the injection feed system, and the size measurement system. A doted parallel circuit indicates the various data acquisition points. Table 2 lists the elements composing the size measurement system along with some specifics of each component.

The main loop can be operated as an open or closed circuit. It comprises the feed tank, the collection tank and the continuous phase pump. The

Gas dispersion results

The average bubble diameters measured after the experiments with glycerol as the continuous phase are presented on Fig. 8 as a function of the average shear. The three curves presented are respectively for 6, 12 and 18 mixing elements in the mixer. Here, we consider the same expression for average shear as Grace (1982). In this expression, the average velocity is based on the total volumetric flow rate of the phases obtained from the inlet conditions divided by the tube diameter and leads toγ˙

Conclusion

The aim of this paper was to present the work completed on the gas dispersion capacity evaluation of the SMX mixer in laminar regime. It was clearly demonstrated that:

  • Elongation is by far the major contributor to the size reduction and simple shear cannot have a large effect. This result is arrived at by comparison with the Kenics mixer in which the shear is known to hold a major influence on the size reduction process.

  • Size reduction is achieved three times faster (in a three times shorter

Notation

Aequation constant
C1,2and3equation constants
Cacapillary number, dimensionless
Ddiameter, m
D32sauter mean diameter, m
Dinjinjection jet/bubble diameter, m
ETtotal energy, J
KsMetzner and Otto constant for apparent shear
Lcharateristic length,m
LMlength of a single static mixer element, m
Maverage of the log-normal bubble size distribution,
dimensionless
Nnumber of static mixing elements, dimensionless
pviscosity ratio (ηdispersed/ηcontinuous), dimensionless
PTtotal power, W
Qvolumetric flow rate, m3/s
Re

Acknowledgements

The financial support of NSERC is gratefully acknowledged.

References (42)

  • T. Avalosse et al.

    Finite element simulation of mixing: 1. Two-dimensional flow in periodic geometry

    A.I.Ch.E. Journal

    (1997)
  • T. Avalosse et al.

    Finite element simulation of mixing: 2. Three-dimensional flow through a Kenics mixer

    A.I.Ch.E. Journal

    (1997)
  • Bade, S., Hoffman, U., 1997. Modeling of the simultaneous comminution and chemical reaction in a non-catalytic...
  • B.J. Bentley et al.

    An experimental investigation of drops deformation and break-up in steady two-dimensional linear flows

    Journal of Fluid Mechanics

    (1986)
  • P.D. Berkman et al.

    Dispersion of viscous liquids by turbulent flow in a static mixer

    A.I.Ch.E. Journal

    (1988)
  • Bigio, D., MacLaren, B., 1991. The effect of viscosity ratio on laminar mixing in the twisted tape motionless mixer....
  • J. De La Villéon et al.

    Numerical investigation of mixing efficiency of helical ribbons

    A.I.Ch.E. Journal

    (1998)
  • Fradette, L., Li, H.-Z., Choplin, L., Tanguy, P., 1997. Efficacité de dispersion gaz–liquide dans un mélangeur statique...
  • Grace, H.P., 1982. Dispersion phenomena in high viscosity immiscible fluid systems and application of static mixers as...
  • M. Heniche et al.

    CFD analysis of a laminar flow through KMX static mixer

    NLGI Spokesman

    (2003)
  • Heniche, M., Reeder, M.F., Tanguy, P.A., Fasano, J., 2004. Numerical comparison of blade shape in static mixing....
  • Cited by (19)

    • CFD modelling of turbulent liquid–liquid dispersion in a static mixer

      2020, Chemical Engineering and Processing - Process Intensification
      Citation Excerpt :

      Static mixers are commonly used in different branches of chemical and process industry [1]: laminar mixing [2,3], enhancement of chemical reactions [4], gas/liquid dispersion [5], mass transfer intensification [6], turbulent dispersion of two immiscible liquids [7] are only few applications to be mentioned.

    • Velocity field characterization of Newtonian and non-Newtonian fluids in SMX mixers using PEPT

      2016, Chemical Engineering Research and Design
      Citation Excerpt :

      SMX mixers, originally developed by Sulzer, are used in a variety of applications, including mixing miscible components (e.g. blending), combining immiscible streams (e.g. emulsification, encapsulation, etc.) and for processes that include a reaction step between the inlet streams (e.g. purification) (Das, 2011; Fradette et al., 2007). Furthermore, SMX mixers can be applied to multiphase processes, dispersing gases or solids through a liquid phase (Fradette et al., 2006; Laporte et al., 2015). SMX mixers are designed for laminar flow applications and rely on breaking and recombining the bulk of the inlet streams into smaller streams, using a series of channels (Paul et al., 2004).

    • The science of foaming

      2015, Advances in Colloid and Interface Science
      Citation Excerpt :

      In this case, bubble generation profits also from the presence of strong viscous and inertial forces. One class of devices is provided by static mixers (Fig. 30a–c) [129–131] which are purpose-designed such that a rapid flow of the gas and the liquid through the interconnected network of pores leads to an optimised shearing action on the bubbles, and hence to bubble break-up (Section 2.3). These mixers can be driven in the laminar regime, leading typically to bubble sizes of the order of 100 μm (Fig. 30d) with a wide range of gas fractions.

    • Flow process conditions to control the void fraction of food foams in static mixers

      2014, Journal of Food Engineering
      Citation Excerpt :

      There are different shapes of static mixers but the cross designed ones, of SMX type, outperform the others, especially in the laminar regime (Meijer et al., 2012. Fradette et al., 2006). Static mixers are usually used for liquid–liquid extraction, liquid–gas absorption, heterogeneous chemical reactions, suspensions, emulsification or solid–liquid mixing (Paglianti, 2008).

    • Effect of impellers configuration on the gas dispersion in high-viscosity fluid using narrow annular gap unit. Part 1: Experimental approach

      2012, Chemical Engineering Science
      Citation Excerpt :

      Over a critical value corresponding to Cac, the bubble breaks into smaller bubbles. Two types of breakup have been observed (Grace, 1982; Fradette et al., 2006): a total breakup (or burst) and a tip breakup. The tip breakup (or tip streaming) occurs in the presence of surfactant and for a low value of Ca whereas the total breakup is observed for much higher values (Fig. 11).

    View all citing articles on Scopus
    View full text