Elsevier

Chemical Engineering Science

Volume 101, 20 September 2013, Pages 329-344
Chemical Engineering Science

Laminar liquid–liquid dispersion in the SMX static mixer

https://doi.org/10.1016/j.ces.2013.06.047Get rights and content

Highlights

  • Superficial velocity and continuous phase viscosity are major effects on mean droplet diameter.

  • Mean droplet diameter decreased with increasing surfactant concentration.

  • Droplet size distributions were measured for above process parameters.

  • Primary droplet diameter is related to the surface area to void volume ratio of the mixer elements.

  • Model for mean droplet diameter based on drop breakup work and interfacial tension.

Abstract

The liquid–liquid dispersion of two immiscible viscous fluids was studied in an SMX static mixer in the laminar flow regime. The dispersed and continuous liquid phases were silicone oils and aqueous solutions of high fructose corn syrup, respectively. Several different phase viscosities were used and the effect of interfacial tension was examined by adding different concentrations of sodium dodecyl sulphate in the aqueous phase. Flow rates of the two phases were varied, and the effect of the volume fraction (holdup) of dispersed phase was explored for a few cases. The droplets were photographed with backlighting and the images were analysed using the Hough transform method. The drop size distributions for each set of conditions were typically obtained on the basis of about 200 images. Mean drop size was expressed as the mass mean diameter D43.

It was found that the “tail” at the higher end of the droplet size distribution decreased with increasing superficial velocity and continuous phase viscosities and the same effect was seen in the values of D43. It was also found that D43 decreased with lowering of the interfacial tension. The effect of the dispersed phase viscosity was significant only at non-zero surfactant concentrations. The holdup was found to have a minimal effect on D43.

A two-stage breakup model has been developed relating D43 to the capillary number. The model uses the concept that primary drops are formed near the entry to the mixer, with a size determined by the spacing between the cross bars in the element. They are then further broken up by a hydrodynamic mechanism based on an energy balance between the energy input due to the viscous flow, and the surface energy increase due to the breakup of the primary drops. This model provides an approximate basis for correlating the data on D43.

Introduction

Motionless mixers or static mixers have been in use since the 1970s (Thakur et al., 2003). They are found in a wide array of industries including grain processing, pharmaceuticals and cosmetics, petrochemicals, refining, and specialty chemicals. They are used in continuous processes in both distributive and dispersive mixing applications.

In general static mixers are more energy efficient than continuous stirred tanks with rotating impellers and their use is favoured for systems that require relatively short residence times and plug flow conditions, for example in the blending or dispersing of fluids that are sensitive to high shear. They are also a more cost effective option for high pressure services. These benefits are realized because of the static mixer's mechanical simplicity and compactness.

A typical static mixer contains a series of mechanical inserts, chosen to suit a particular application and to optimize the mixing performance. These inserts can be broadly categorized according to twisted-ribbon, plate and structured geometries. The twisted-ribbon type is ideal for mixing of low viscosity fluids (such as gas–liquid dispersions) under turbulent conditions. Plate type mixers are also used for applications in the turbulent flow regime, as these designs create vortices. Structured packings have more complex geometry and are often placed such that each cartridge is oriented 90° relative to its neighbours. The most versatile in this category are the SMX and SMXL types (Sulzer Chemtech USA, Inc., Tulsa, Oklahoma) which have geometries with cross bars angled at 45° and 30° for SMX and SMXL, respectively. The SMX packing, which is studied in this work, is said to be the best design choice for plug flow in tubular reactors and for blending and dispersing high viscosity fluids in the laminar flow regime (Paul et al., 2004, p. 428).

In order to scale up static mixers reliably in liquid–liquid dispersion applications, it is necessary to develop correlations or fundamental models to predict key characteristics such as pressure drop, mean droplet size and size distributions. The latter two characteristics depend on the fluid behaviour of multiple phases, which is affected by the two liquid phase viscosities and interfacial tension. Such correlations and basic insights also help to optimize the packing geometry for better performance. In liquid–liquid dispersions with SMX packing for example, there is a strong industrial motivation to eliminate the occurrence of oversized droplets in order to produce a more uniform droplet size distribution.

Much numerical and experimental work has been done in the area of droplet breakup since the early contribution by Grace (1982). For example, Rauline et al. (1998) have done numerical investigations to characterize elongational flow in various static mixers. Liu (2005) has done numerical and experimental studies to study flow fields inside an SMX mixer. He focused on single drop breakup in the laminar regime (Liu et al., 2005). Direct investigations of droplet size distribution in SMX mixers have been done by Legrand et al. (2001), Rama Rao et al. (2007), Fradette et al. (2007), Theron et al. (2010) and Das et al. (2005). Legrand et al. (2001) studied droplet dispersion in the turbulent regime and have based their correlations on the classical Kolmogorov theory, whereas Fradette et al. (2007) studied dispersion in the laminar regime. In all these studies, a major limiting factor was the measurement of droplet size distribution. In most cases, droplet size measurement was done using the micro encapsulation technique where the droplets were hardened at the surface by means of interfacial polymerization. Rama Rao et al. (2007) and Fradette et al. (2007) used image/video capture and then manually counted and measured the droplets to obtain a distribution. In the latter case the detailed measurements seriously limited the amount of data that could be acquired. These limitations can be reduced by automated data collection and analysis which can help experimenters to conduct more runs under different conditions, and come up with better correlations.

Therefore this work has two research objectives. The first is the capture of the droplet size distribution based on the application of computer vision algorithms for automated measurement. The second objective is the collection of substantial droplet size distribution data with the SMX elements in the laminar regime under varying viscosity ratios, holdups and most importantly under varying interfacial tension. This data will be used to develop a model for correlating the droplet size distribution with the variables involved.

Section snippets

Literature review

In a static mixer, a drop can break up either because of hydrodynamic forces around the drop or as a result of collision with the internals. In the former case, breakup occurs when viscous shear stress from the continuous phase exceeds the drop's internal viscous and surface stresses. Grace (1982) has done fundamental experimental work in this regard with single drops in steady simple shear and 2D extensional flow. He described the breakup in terms of two dimensionless terms namely the

Fluids used

Silicone oil was used as the dispersed phase and an aqueous solution of high fructose corn syrup (HFCS) was used as the continuous phase. The silicone oils were Xiameter-PMX200 silicone fluids from Dow Corning, and the HFCS was supplied by UNIVAR Canada. The properties of the fluids are given in Table 1. For each HFCS solution prepared, the viscosities were measured with a Brookfield Viscometer (using spindle #1) and verified with Zahn Cups. Individual values differed slightly from those given

Results and discussion

Experimental data have been tabulated by Das (2011). They were mainly collected using the 41.18 mm diameter mixer, augmented with limited results using a smaller 15.75 mm diameter mixer due to time constraints.

The effect of holdup (dispersed phase fraction) was studied only for the cases without added surfactant, and only with continuous phase viscosities of 25 and 70 mPa S, and at certain flow rates. Since the holdup effect was minor, these points appear as clusters in the figures that follow. The

Proposed predictive model

An observation-based model has been proposed on the basis of a two-stage mechanism of drop breakup.

In the first stage, very soon after the dispersed phase enters the first packing element, primary drops (with diameter D0), are formed by a slicing interaction with the packing. The diameter D0 is of the same order of magnitude as the gap between the adjacent plates in the packing, consistent with the observed values of Dmax. In the second stage, drops with diameter D<D0 are formed by a mechanism

Conclusions

  • 1.

    Superficial velocity and continuous phase viscosity have the strongest effect on D43. The effect of dispersed phase viscosity was significant only at low and high surfactant concentrations.

  • 2.

    There was a measurable decrease in D43 at low surfactant concentration, and a significant decrease at high surfactant concentration. This confirmed that there is an effect of interfacial tension.

  • 3.

    Droplet size distributions typically were found to have a long and wide tail, which became narrower and shorter

Nomenclature

    Ca

    capillary number (=µcUcd/σ)

    Cacrit

    critical capillary number

    D

    mixer diameter (m)

    D0

    primary drop diameter (m)

    D32

    Sauter mean diameter (m)

    D43

    mass mean diameter (m)

    Df

    final droplet diameter after breakup of a primary drop (m)

    Di

    diameter of an individual drop in a distribution (m)

    Dmax

    maximum diameter (m)

    e

    power dissipation per unit mass (m2/s3)

    E

    overall energy (J)

    kd

    dimensionless constant

    N

    number of droplets formed after breakage

    n

    number of mixing elements

    N0

    number of droplets with an initial diameter D0

    p

Acknowledgements

We are grateful to Procter and Gamble Inc. for funding this research. We also thank Paul Gatt and Dan Wright for their help, respectively in equipment fabrication and instrumentation support. We also acknowledge the partial financial support in the form of a teaching assistantship provided to M.D. by the Department of Chemical Engineering at McMaster University.

References (28)

  • L.M.R. Brás et al.

    Drop distribution determination in a liquid–liquid dispersion by image processing

    Int. J. Chem. Eng.

    (2009)
  • Catalfamo, V., Blum, G.L., Jaffer, S.A. 2003. Apparatus for In-line Mixing and Process of Making Such Apparatus, US...
  • V. Christini et al.

    Drop breakup and fragment size distribution in shear flow

    J. Rheol.

    (2003)
  • Das, M.D., 2011. Laminar Liquid–Liquid Dispersion in SMX Mixer. M.A.Sc. Thesis. McMaster University, Hamilton, Ontario,...
  • Cited by (20)

    • Current advances in liquid–liquid mixing in static mixers: A review

      2022, Chemical Engineering Research and Design
      Citation Excerpt :

      Diversity of construction materials (making static mixers suitable for an extensive spectrum of applications) (Theron and Le Sauze, 2011). Short vs. long residence time distributions to achieve a similar degree of mixing (Thakur et al., 2003; Das et al., 2013). Rate of dilution easy to control and scale (Thakur et al., 2003).

    • Stiffening control of cement-based materials using accelerators in inline mixing processes: Possibilities and challenges

      2021, Cement and Concrete Composites
      Citation Excerpt :

      Unfortunately, striation thickness is hard to measure because of uneven distribution. On the other hand, point to point concentrations (i.e. pixel to pixel concentrations) are relatively easy to measure such as the coefficient of variation (COV) and the relative standard deviations (RSD) of the pixel distribution in images of the hardened samples [109,122,123]. Detecting methods can also adopted from the chemical industries.

    View all citing articles on Scopus
    View full text