Foam free drainage and bubbles size for surfactant concentrations below the CMC

https://doi.org/10.1016/j.colsurfa.2015.09.050Get rights and content

Highlights

  • Role of bubble size on the free drainage of moderately stable foams.

  • Incorporation of data into the modified L–L model yields surface shear viscosity values.

Abstract

Foams produced with surfactant concentration below the Critical Micelle Concentration (CMC) are usually moderately stable due to high drainage rates and intense bubbles coarsening/coalescence. Aim of this work is to examine how such low surfactant concentration affects foam destabilization and elucidate the interplay between free drainage and bubble size variation. Foam destabilization experiments are conducted at varying SDS concentrations (below the CMC) where the evolution of liquid fraction and bubbles size is registered simultaneously. Instantaneous volume measurements of the drained liquid and the remaining foam yield the evolution of the global liquid fraction and drainage rate in the foam. Continuous electrical conductance measurements give the local liquid fraction and drainage rate in the foam. Microphotographs allow estimation of bubble size distribution and bubble population at regular time intervals. The present data show that the lifetime of moderately stable foams depend largely on surfactant concentration below the CMC but this effect does not scale linearly with surfactant concentration. Furthermore, measurements are fitted to semi-empirical expressions and are compared to a modified Leonard and Lemlich (L–L) drainage model that has been expanded to incorporate bubble size evolution. The latter is a rough approximation based on certain assumptions but it is a fair approach given the excessive difficulty of detailed numerical calculations. The comparison reveals the significant role of bubble size on the free drainage of moderately stable foams. Interestingly, incorporation of global liquid fraction data into the modified L–L model yields surface shear viscosity values in agreement with literature.

Introduction

A foam comprises a mixture of liquid and bubbles, stabilized by surfactant, having an anisotropic distribution of liquid fraction and bubble size that both evolve dynamically [3], [36]. Free foam drainage is a complex process since each of the following may take place simultaneously: gravitational liquid flow (liquid drains out of the foam chiefly through Plateau Borders (PB) and nodes), shear stresses imparted by the gas–liquid interfaces, capillarity induced liquid suction (opposing gravity especially at the foam bottom), bubbles size increase and bubble population decrease because of coalescence (merging of adjacent bubbles as result of rupture of the liquid films between them), and coarsening (enlargement of large bubbles by gas diffusion from smaller adjacent bubbles due to pressure differential between bubbles of unequal size-Ostwald ripening) [3].

The presence of surfactants in a foam can, in principle, influence both the gas–liquid interfacial properties (e.g., surface tension, surface dilatational viscosity, surface shear viscosity, surface elasticity) and the bulk liquid viscosity. Yet, for the bulk viscosity to be affected very large concentrations -well above the CMC- are required. Many preceding works (extended reviews on this topic can be found elsewhere e.g. [9], [44] have shown that the type and concentration of surfactant modifies considerably the gas–liquid interfacial properties and, therefore, can change the flow type inside the foam’s liquid network (i.e. PB, nodes and films) and can lead to significant alteration of drainage rates. However, most studies examine the above either in foam systems that maintain their initial bubbles size during drainage (i.e. surfactant concentrations well above the CMC and at high liquid fractions) or/and overlook the evolution of bubbles size, e.g., Cervantes-Martinez et al. [8], Boos et al. [3].

Several works have shown that bubble size evolution due to coarsening and free foam drainage are closely interrelated. Hilgenfeldt et al. [18], have shown that strong bubble coarsening leads to shorter drainage time (accelerated drainage) which is independent from the initial liquid content, ϵ0. Both Magrabi et al. [25] and Vera and Durian [47] measured bubble size and liquid void fraction at various positions along the foam column during free drainage. Magrabi et al. [25] concluded that at their particular system, foam destabilization takes place at three stages: initially (<200s) drainage dominates over coarsening, later (200s–900s) drainage and coarsening occur concurrently and eventually (>900s) coarsening prevails over drainage. Vera and Durian [47] measured increased coarsening rates for drier foams and suggested that coarsening plays an unavoidable role in free drainage. Saint-Jalmes and Langevin [36] studied the coupling between drainage and coarsening of aqueous foams and showed that the type of flow inside the PBs depends on both bubbles size and gas–liquid interfacial properties. The above works do not take into account the effect of the type or/and surfactant concentration on the interaction between drainage and coarsening and, additionally, consider bubbles coalescence negligible.

It is known that for low surfactant concentrations, especially below the CMC, bubble coalescence may contribute significantly to bubble size variation during foam destabilization [10]. To the best of our knowledge, only Carrier and Colin [7], isolated bubble coalescence from coarsening by using a mixture of gases that has very low mass diffusivity in water. These authors examined the influence of various surfactants (i.e. SDBS and TTAB) at different concentrations on the interplay between coalescence and free drainage. They found that coalescence is dramatically enhanced below a critical liquid fraction which depends on the surfactant concentration and nature of the surfactant but does not depend on bubble size.

The only systematic work that we were able to identify on foam free drainage using surfactant below the CMC is that by Harvey et al. [17] These authors studied the effect of different types (i.e. Dowfroth-250, Dowfroth-400 and SDS) and different concentrations of surfactants (always below the CMC) on foam lifetime and on initial bubbles size. They showed that foam stability is controlled not only by the liquid drainage process, which is a function of surface shear viscosity, μs, but also by coalescence which is a function of other interfacial properties (e.g. surface dilatational viscosity of adsorbed layers). However, this study, being chiefly of technological orientation, did not investigate the effect of surfactant concentration on the evolution of bubble size and liquid fraction during foam drainage.

Scope of this work is to study the influence of surfactant concentration below the CMC on the interplay between bubbles size evolution (due to coarsening and coalescence) and free drainage. Surfactant concentrations below the CMC are used, because at these concentrations foams are only moderately stable. Novel data are presented concerning bubble size evolution and liquid fraction profiles both locally (electrical resistance measurements) and globally (volumetric measurements) for various surfactant concentrations. Additionally, it is examined if the present findings can be described by the well known drainage model of Leonard and Lemlich (1965, L–L model) [53] after incorporating the experimentally observed bubble size evolution to cope with coarsening and coalescence. In the following sections, the foam preparation procedure and a few essential physicochemical properties of the foam solution are presented first, being followed by an outline of the experimental setup and the employed measuring techniques. A section comes next with experimental results and discussion on the underlying phenomena.

Section snippets

Foam preparation

Surfactant solutions are produced using deionized water and sodium dodecyl sulfate (SDS  98% purity; Fluka). Foams are produced employing four concentrations of SDS (300, 600, 1000 and 2000 ppm) which are below the reported CMC value at 30 °C (∼2500 ppm). SDS is known for producing foams unstable with respect to coalescence [12]. A small amount of NaCl 4 × 10−3 M; (≥99% purity Merk) is added to deionized water to yield the ionic strength met in applications where foams are produced from fresh

Evolution of drained liquid volume and foam volume

Table 1 presents initial foam volume produced by whipping 200 ml of the different surfactant solutions. As surfactant concentration increases foamability increases, too. This means that less initial liquid volume, VL,0 is used to produce the same initial amount of foam, VF,0and as a result the foam becomes dryer and more stable (i.e. initial global liquid fraction, ϵ0, decreases). Fig. 2a presents the evolution of foam volume, VF, and drained liquid volume, VL, for different surfactant

Comparison with model

Relation of measurable global drainage quantities to characteristics of a foam is not straightforward since free drainage is difficult to analyze. The situation is different from that of forced drainage where the front velocity is an easily measured variable directly related to foam characteristics [43]. In case of free drainage, there is a continuous spatio-temporal evolution of liquid fraction from the initial to the equilibrium state. This means that a complex spatially distributed foam

Conclusions

In this study we examine the effect of varying surfactant concentrations below the CMC on the free drainage of moderately stable foams. In such systems high drainage rates are combined with intense bubbles coarsening and coalescence.

Larger surfactant concentrations result to dryer foams, with a more packed structure. Concerning the drainage process, the successful fitting of a sigmoid equation on the drainage data reveals a maximum value on the drainage rate profiles. This maximum value is the

Acknowledgments

The present activity was under the umbrella of COST Actions MP1106 and CM1101 and the European Space Agency funded programs: FASES (Fundamental and Applied Studies of Emulsion Stability) and PASTA (PArticle STAbilised Emulsions and Foams).

References (56)

  • A.V. Nguyen

    Liquid drainage in single Plateau borders of foam

    J. Colloid Interface Sci.

    (2002)
  • M. Papara et al.

    Container effects on the free drainage of wet foams

    Chem. Eng. Sci.

    (2009)
  • A. Patist et al.

    Effect of long chain alcohols on micellar relaxation time and foaming properties of sodium dodecyl sulfate solutions

    J. Colloid Interface Sci.

    (1998)
  • O. Pitois et al.

    Liquid drainage through aqueous foam: study of the flow on the bubble scale

    J. Colloid Interface Sci.

    (2005)
  • P. Stevenson

    Remarks on the shear viscosity of surfaces stabilised with soluble surfactants

    J. Colloid Interface Sci.

    (2005)
  • P. Stevenson

    Dimensional analysis of foam drainage

    Chem. Eng. Sci.

    (2006)
  • T.D. Karapantsios et al.

    On the design of electrical conductance probes for foam drainage applications. Assessment of ring electrodes performance and bubble size effects on measurements

    Colloids Surf. A Physicochemical Eng. Aspects

    (2008)
  • X. Zabulis et al.

    Detection of densely dispersed spherical bubbles in digital images based on a template matching technique Application to wet foams

    Colloids Surf. A Physicochemical Eng. Aspects

    (2007)
  • Y. Yan et al.

    A study on the kinetics of liquid drainage from colloidal gas aphrons (CGAs)

    Colloids Surfaces A Physicochemical Eng. Aspects

    (2005)
  • L. Indrawati et al.

    Characterization of protein stabilized foam formed in a continuous shear mixing apparatus

    J. Food Eng.

    (2008)
  • S. Qian et al.

    Experimental investigation of liquid foams by polarised light scattering technique via the Mueller matrix

    Chem. Eng. Sci.

    (2013)
  • P. Stevenson et al.

    Liquid overflow from a column of rising aqueous froth

    Minerals Eng.

    (2003)
  • A. Bhakta et al.

    Decay of standing foams: drainage, coalescence and collapse

    Adv. Colloid Interface Sci.

    (1997)
  • J. Boos et al.

    Protocol for studying aqueous foams stabilized by surfactant mixtures

    J. Surfactants Deterg.

    (2012)
  • I. Cantat et al.

    Foams Struct. Dyn.

    (2013)
  • V. Carrier et al.

    Coalescence in draining foams

    Langmuir

    (2003)
  • N.D. Denkov et al.

    The role of surfactant type and bubble surface mobility in foam rheology

    Soft Matter

    (2009)
  • M. Durand et al.

    Physicochemical approach to the theory of foam drainage

    Eur. Phys. J. E

    (2002)
  • Cited by (44)

    • Expansion of the concept of critical micelle concentration for the application of a saturated monoacylglyceride emulsifier in aerosol whipping cream

      2022, Food Research International
      Citation Excerpt :

      To understand how the concept of CMC has to be expanded to fit the application of me–sMAG in model AWC, it is necessary to take a closer look at said concept. The concept of CMC discussed in literature is limited to either the stabilisation of emulsions (Das & Kinsella, 1990; McClements & Jafari, 2018; Thanasukarn, Pongsawatmanit, & McClements, 2004) or foams (Lioumbas, Georgiou, Kostoglou, & Karapantsios, 2015; Majeed, Sølling, & Kamal, 2020), systems with only one type of interfacial area (oil–water/water–oil or gas–water/gas-oil). However, the stabilisation of the interface oil–water (unfoamed) and of the interface gas–water (foamed) has to be considered for foamed AWC.

    • Foam stability: The key to inhibiting slug generation in gas–liquid flow

      2022, Journal of Petroleum Science and Engineering
    View all citing articles on Scopus
    View full text