Acessibilidade / Reportar erro

THE EFFECT OF SURFACTANT AND HIGH MOLECULAR WEIGHT POLYMER ADDITION ON PRESSURE DROP REDUCTION IN PIPE FLOW

Abstract

The main aim of this paper is to present a possibility to enhance the drag reduction effect in straight pipe flow by the simultaneous addition to the transported liquid of a small amount of high molecular weight polymers and surfactants. Qualitative analysis of the polymer-micellar additive influence on the shape and character of flow resistance curves has been performed. Also multicomponent polymer-micellar solution flow resistance curves were compared with appropriate single additive polymer or surfactant solution flow resistance curves. The experimental data shows that, for polymer-micellar solutions, the stable transitional zone between the laminar and the turbulent flow regions is extended toward higher values of the critical Reynolds numbers. Occurrence of the phenomenon can be explained by the flow laminarization caused by polymer-micellar aggregates. Existence of the third extended drag reduction zone in the turbulent range of flow has also been observed for the first time.

Keywords:
Non-Newtonian flow; Pressure drops reduction; Polymer-surfactant additive; Polymer-micellar aggregate

INTRODUCTION

Abnormal flow drag reduction by surfactant or polymer additives is intensively examined and described in the subject literature (Matras, 1984Matras, Z., Przepływ cieczu Tomsa w przewodach kołowych. Politechnika Krakowska, Monografia 29 (1984). (In Polish).; Borostow, 2008Borostow, W., Drag reduction in flow: Review of applications, mechanism and prediction. Journal of Industrial and Engineiring Chemistry, 14, 409-416, (2008).; White and Mungal, 2008White, C. M. and Mungal, M. G., Mechanics and predictions of turbulent drag reduction with polymer additives. Annular Review of Fluid Mechanics, 40, 235-256 (2008).; Wang et al., 2011Wang, Y., Yu, B., Zakin, J. L. and Shi, H., Review on drag reduction and its heat transfer by additives. Advances in Mechanical Engineering, 10, 17 pages (2011).; Shu-Peng, 2012Cai, S.-P., Drag reduction of a cationic surfactant solution and its shear stress relaxation. Journal of Hydrodynamics, 24(2), 202-206 (2012).; Tamano et al., 2010Tamano, S., Ito, M., Kato, K. and Yokota, K., Turbulent drag reduction in nonionic surfactant solutions. Physics of Fluids, 22(5), 055102 (2010).;). This phenomenon allows a significant increase of flow rate without increasing power demand, or vice versa - to reduce power demand while maintaining a constant flow rate. It provides large potential possibilities for the application of this effect in different industry branches, particularly in the oil industry or in heating, firefighting, transport of slurries, sludge and brines (Motier et al., 1996Motier, J. F., Chou, L. C. and Kommareddi, N. S., Commercial drag reduction: Past, present, and future. Proceedings of the ASME Fluids Engineering Division Summer Meeting, San Diego, Calif., USA (1996).; Figueredo and Sabadini, 2003Figueredo, C. R. and Sabadini, E., Firefighting foam stability: The effect of the drag reducer poly(ethylene) oxide. Colloids and Surfaces, A: Physicochemical and Engineering Aspects, 215(1-3), 77-86 (2003). ; Dujmovich and Gallegos, 2005Dujmovich, T. and Gallegos, A., Drag reducers improve throughput, cut costs. Offshore, 65(12), pp. 55-58 (2005). ; Wang et al., 2011Wang, Y., Yu, B., Zakin, J. L. and Shi, H., Review on drag reduction and its heat transfer by additives. Advances in Mechanical Engineering, 10, 17 pages (2011).). Causes of the described drag reduction have been perceived in the existence of a new internal solution structure which is formed in the moment when the special additives are introduced into the solution. Addition of high molecular weight polymer agents into the solvent results in macromolecule solution formation, which has a crucial influence on turbulence structure in the flow (Borostow, 2008Borostow, W., Drag reduction in flow: Review of applications, mechanism and prediction. Journal of Industrial and Engineiring Chemistry, 14, 409-416, (2008).; White and Mungal, 2008White, C. M. and Mungal, M. G., Mechanics and predictions of turbulent drag reduction with polymer additives. Annular Review of Fluid Mechanics, 40, 235-256 (2008).; Tamano et al., 2010Tamano, S., Ito, M., Kato, K. and Yokota, K., Turbulent drag reduction in nonionic surfactant solutions. Physics of Fluids, 22(5), 055102 (2010).). Although, there is still strong debate on whether a single polymer molecule or clusters of polymer molecules are responsible for the drag reduction effect, experimental results clearly prove that even a dozen ppm of polymer concentration in the solvent induces an efficient drag reduction effect in the turbulent range of flow (White and Mungal, 2008White, C. M. and Mungal, M. G., Mechanics and predictions of turbulent drag reduction with polymer additives. Annular Review of Fluid Mechanics, 40, 235-256 (2008).; Wang et al., 2011Wang, Y., Yu, B., Zakin, J. L. and Shi, H., Review on drag reduction and its heat transfer by additives. Advances in Mechanical Engineering, 10, 17 pages (2011).; Zadrazil et al., 2012Zadrazil, I., Bismarck, A., Hewitt, G. F. and Markides, C. N., Shear layers in the turbulent pipe flow of drag reducing polymer solutions. Chemical Engineering Science, 72, 142-154 (2012).).

In case of application of surfactants as drag reducing additives, formation of micelle structures is observed (Tamano et al., 2010Tamano, S., Ito, M., Kato, K. and Yokota, K., Turbulent drag reduction in nonionic surfactant solutions. Physics of Fluids, 22(5), 055102 (2010).; Wang et al., 2011Wang, Y., Yu, B., Zakin, J. L. and Shi, H., Review on drag reduction and its heat transfer by additives. Advances in Mechanical Engineering, 10, 17 pages (2011).). In order to improve the micellarization process effectiveness in surfactant solutions, small amounts of electrolytes are applied (e.g. sodium salicylate or sodium bromide). At the no motion condition, the mentioned structures are chaotic. Only during fluid flow shearing do both macromolecules and micelles start to arrange in characteristic orientations, in accordance with the principle of the minimum resistance. In both surfactant and polymer induced experiments on the drag reduction effect, a hypothetical mechanism of the phenomenon that is widely accepted is the interaction between polymer or surfactant molecules and the flow turbulence structure. On the basis of advanced measurement techniques, such as Particle Image Velocimetry (PIV), it is observed that polymer additives can lead to reduction or elimination of the ejections of low-momentum fluid from the wall region to the outer velocity region (Warholic et al., 2001Warholic, M. D., Heist, D. K., Katcher, M. and Hanratty, T. J., A study with particle- image velocimetry of the influence of drag-reducing polymers on the structure of turbulence. Exp. Fluids, 31, 474-483 (2001).). It is also observed that the presence of polymers leads to a decrease in the frequency and the intensity of large-scale ejections when compared to a Newtonian solvent (Liberatore et al., 2004Liberatore, M. W., Baik, S., Mchugh, A. J. and Hanratty, T. J., Turbulent drag reduction of polyacrylamide solutions: Effect of degradation on molecular weight distribution. J. Non-Newtonian Fluid Mech., 123, 175-183 (2004).) and to the reduction of the magnitude and frequency of the small-scale eddies (Liberatore et al., 2004). Usually, elongational viscosity or elasticity of the polymer chain (White and Mungal, 2008White, C. M. and Mungal, M. G., Mechanics and predictions of turbulent drag reduction with polymer additives. Annular Review of Fluid Mechanics, 40, 235-256 (2008).) are proposed to explain the hypothetical polymer drag reduction mechanics. On the other hand, for a surfactant aqueous solution, which exhibit neither viscoelastic properties nor the presence of elongational viscosity (Wang et al., 2008Wang, Y., Yu, B., Zakin, J. L. and Shi, H., Review on drag reduction and its heat transfer by additives. Advances in Mechanical Engineering, 10, 17 pages (2011).), a local shear-thickening hypothesis is proposed (Wang et al., 2008Wang, Y., Yu, B., Zakin, J. L. and Shi, H., Review on drag reduction and its heat transfer by additives. Advances in Mechanical Engineering, 10, 17 pages (2011).; Hadri et al., 2010Hadri, F., Besq, A., Guillou, S. and Makhloufi, R., Drag reduction with an aqueous solution of CTAC-NaSal: Study of the wall slip with a Couette geometry. Comptes Rendus Mécanique, 338(3), 152-157 (2010).).

A novel and poorly recognized effect is a phenomenon of fluid flow drag reduction by the simultaneous addition into the solvent of both high molecular weight polymer and surfactant with salt. In the few published works related to this subject (Minatti et al., 1996Minatti, E., Zanette, D., Salt effects on the interaction of poly(ethylene oxide) and sodium dodecyl sulfate measured by conductivity. Colloids Surfaces, A: Phisicochem. Eng. Aspects, 113, 237 (1996).; Hou et al., 1999Hou, Z., Li, Z. and Wang, H., Interaction between poly(ethylene oxide) and sodium dodecyl sulfonate as studied by surface tension, conductivity, viscosity, electron spin resonance and nuclear magnetic resonance. Colloid Polym. Sci., 277, 1011-1018 (1999).; Suksamranchit et al., 2006Suksamranchit, S., Sirivat, A. and Jamieson, A. M., Polymer-surfactant complex formation and its effect on turbulent wall shear stress. Journal of Colloid and Interface Science, 294(1), 212-221 (2006).; Jung et al., 2011) the internal structure formation and chemical reaction process in polymer-micellar solutions are mainly highlighted. Initial attempts at experimental examination of the drag reduction effect have been performed (Matras et al., 2008Matras, Z., Malcher, T. and Gzyl-Malcher, B., The influence of polymer-surfactant aggregates on drag reduction. Thin Solids Films, 516, 8848-8851 (2008).; Mohsenipour et al., 2013Mohsenipour, A. A., Pal, R. and Prajapati, K., Effect of cationic surfactant addition on the drag reduction behaviour of anionic polymer solutions. The Canadian Journal of Chemical Engineering, 91(1), 181-189 (2013).). The results confirmed that simultaneous addition into the solvent of these additives combines and intensifies the positive features of their purely polymer and micellar analogues, providing additional extension of the drag reduction zones. Moreover, the researchers indicate that this new effect requires a comprehensive experimental study to gain a deeper knowledge of this phenomenon.

The presence of polymer macromolecules in the surfactant solution enhance the micelle structure formation ability. It leads to the formation of micellar structure at a lower concentration. The newly formed macromolecule is called an aggregate (Matras et al., 2008Matras, Z., Malcher, T. and Gzyl-Malcher, B., The influence of polymer-surfactant aggregates on drag reduction. Thin Solids Films, 516, 8848-8851 (2008).; Mohsenipour et al., 2013Mohsenipour, A. A., Pal, R. and Prajapati, K., Effect of cationic surfactant addition on the drag reduction behaviour of anionic polymer solutions. The Canadian Journal of Chemical Engineering, 91(1), 181-189 (2013).). Adding a slight amount of salt (e.g. NaCl or NaSal) to the high molecular weight polymer and surfactant solution causes micelle size growth. The number of micelles linked with the polymer chains also increases. Furthermore, the addition of the salt can increase the solution viscosity.

The aim of this paper is to perform an analysis of the drag reduction efficiency induced by simultaneous addition to the solvent of both surfactants and high molecular weight polymer, comparing to the drag reduction effect obtained by addition of pure polymer or pure surfactant agents.

CHARACTERISTIC OF POLYMER-MICELLAR SOLUTION INTERNAL STRUCTURE

Simultaneous addition to the solvent of small amounts of polymer and surfactant additives triggers initiation of the micellarization process at much lower concentration, compared to the critical micelle concentration (CMC) (Deo et al., 2007; Jönsson et al., 1998Jönsson, B., Lindman, B., Holmberg, K. and Kronberg, B., Surfactants and Polymers in Aqueous Solution. John Wiley & Sons Chichester, UK (1998).). This concentration at which micelle formation occurs in the presence of polymer macromolecules is called the critical aggregation concentration (CAC). The newly formed polymer-micelle macromolecules are called aggregates (Diamand and Andelman, 1999Diamant, H. and Andelman, D., Onset of self-assembly in polymer-surfactant systems. Europhysics Letters, 48(2), 170-176 (1999).; Matras et al., 2008Matras, Z., Malcher, T. and Gzyl-Malcher, B., The influence of polymer-surfactant aggregates on drag reduction. Thin Solids Films, 516, 8848-8851 (2008).). From the experimental study of polymer-micellar aqueous solutions (Jönsson et al. 1998Jönsson, B., Lindman, B., Holmberg, K. and Kronberg, B., Surfactants and Polymers in Aqueous Solution. John Wiley & Sons Chichester, UK (1998).; Mohsenipour et al., 2013Mohsenipour, A. A., Pal, R. and Prajapati, K., Effect of cationic surfactant addition on the drag reduction behaviour of anionic polymer solutions. The Canadian Journal of Chemical Engineering, 91(1), 181-189 (2013).), the mechanism of aggregates formation process can be described. Initially polymer and surfactants molecules occur in the solution independently. The situation significantly changes when a small amount of salt is introduced into the solution. According to Minatti et al. (1996Minatti, E., Zanette, D., Salt effects on the interaction of poly(ethylene oxide) and sodium dodecyl sulfate measured by conductivity. Colloids Surfaces, A: Phisicochem. Eng. Aspects, 113, 237 (1996).) and Matras et al. (2008)Matras, Z., Malcher, T. and Gzyl-Malcher, B., The influence of polymer-surfactant aggregates on drag reduction. Thin Solids Films, 516, 8848-8851 (2008)., it causes surfactant micelle growth, linking polymer together with micelles and forming aggregates. Consequently, it increases solution viscosity values. It should be pointed out that salt addition causes a significant viscosity decrease. This is justified by the more intense interaction between polymer chains. Further increase of the polymer concentration leads to a polymer saturation point (PSP) - the maximum viscosity value of the mixture is reached.

MATERIALS AND METHODS

Having analysed the level of difficulty of the planned experimental tests and taking into account the type of physical quantities to be measured, the experiment was performed using a modern capillarypipe rheometer, designed and constructed in the Division of Fluid Mechanics laboratory at the Cracow University of Technology (Matras and Walczak, 2003). The device allows the operator to conduct a comprehensive identification of rheological characteristics of the examined liquid in laboratory conditions.

The versatility of the described capillary-pipe rheometer allows the operator to assign not only classical experimental flow curves in the laminar range of flow, but also to examine and interpret properly the flow characteristics of fluids which behave differently when compared with purely viscous non-Newtonian fluids in the turbulent flow region. This applies primarily to the solutions of polymers and surfactants, which are the subject of research in this work, as well as to the fluids which can be considered to be on the border between the physical continuum and multi-phase systems, e.g., rheostable (purely viscous) or viscoelastic suspensions. A schematic diagram of the capillary-pipe rheometer is illustrated in Figure 1.

Figure 1
Diagram of the multifunction capillary-pipe rheometer: 1 - straight capillaries and pipes with circular cross-section; 2 - differential pressure drop sensors PD1 and PDF, 3 - electromagnetic flowmeters; 4 - flowmeter controllers; 5 - Multistage rotodynamic pumps with triple-phase electrical engine; 6 - data acquisition system; 7 - microprocessor frequency converter L100; 8 - amplifiers WP-01A; 9 - tanks; 10 - temperature sensors PT 100.

The basic elements of the device are straight copper and stainless steel capillaries and pipes with circular cross-sections (1). They are used for the measurement of fluid flow pressure loss at pipe distance L. Static pressure holes were spotted at distances Linl= Lout ?150 d from the pipe inlet and outlet. This ensures stable flow conditions and eliminates the influence of the so-called "entrance effect."

The fluid flow in the capillary-pipe rheometer is forced by hermetic multistage rotodynamic pumps (5), which suck in the fluid from one of the tanks (9) and pump it then into one of the eight horizontal pipes of different diameters (1). After passing through the electromagnetic flowmeters (3) the fluid returns to the storage tank. The measurements of the pressure losses were performed using PD1 or PDF differential pressure drop sensors (2).

For the temperature control (possibility of maintaining a constant temperature in the measuring system was provided) a resistive temperature sensor (10) was placed in the fluid supply pipe.

Pumps can work in either series or parallel arrangements, depending on the required flow rate or pressure loss value. This allows the operator to obtain a wide range of Reynolds numbers, reaching the value of 3×105 for a measurement system pressure value up to 10 bar, without the loss of fluid continuity (no foaming of the solution or air bubbles).

A microprocessor frequency converter (7) was used to control the pump frequency and consequently volumetric flow rate. The main element supporting the action of the capillary-pipe rheometer is a multi-channel data acquisition system SPIDER 8 (Hottinger Baldwin Messtechnik), arranged to measure the electrical signals from the different sensors (tension, force, pressure, displacement, acceleration and temperature sensors).

Additionally, large volume tanks (1 m3) were used to eliminate the effect of foaming of micellar and polymer-micellar solutions and to minimize the influence of the unavoidable degradation of the polymer-micellar aggregates or macromolecular structures on the measurement results.

During the preparation of the polymer-surfactant solutions, the pH of the chosen drag reducing additives should be particularly considered. Incorrect selection of pH may result in an undesired chemical reaction.

Anionic surfactants cannot be used when aqueous solutions of certain polymers can have an acidic reaction. Cationic and anionic surfactants can be combined with non-ionic polymer solutions.

After preliminary study, the following drag reducers were used for experimental analysis:

  • Poly(ethylene oxide) [CH2 CH2 O]n (PEO) - non-ionic polymer with viscosity-based high molecular weight given by the manufacturer equal to 8·106 g/mol-1, purchased from Sigma-Aldrich, Inc.

  • Cetyltrimetyl ammonium bromide [CH3(CH2)i5N(CH3)3]+Br~ (CTAB) - cationic surfactant purchased from Sigma-Aldrich, Inc.

In order to lower the CAC value, the salt C7H5NaO3 (NaSal) sodium salicylate has been used. Different compositions of polymer, surfactant and salt mass fraction in solvent were used in order to analyse the chemical additive concentration effect. Distilled water was used as solvent. A Polna, Inc. Electrical Distillatory Type DE20 was used to purify tap water. The conductivity of the solvent was of the order of 1 µS/cm.

After addition to the solvent of the appropriate drag reducers, solutions were mixed gently so as not to cause mechanical degradation of polymer chains. The first mixing was performed in cylindrical vessels, by the use of a roller mixer of our own design with very a low rotational speed equal to 1-5 rpm. Then the solution was diluted in the main tanks. Before measurements, mixtures were left to rest for 24 hours.

In order to conduct an analysis of drag reduction efficiency by simultaneous addition to the solvent of both surfactants and high molecular polymer, compared to the drag reduction effect obtained by the addition of pure polymer or pure surfactant agents, 15 solutions with different additive compositions and concentrations were investigated. Designation and the composition of the analysed mixtures are presented in Table 1.

Table 1
Summary of the analysed solution and their designated rheological constants.

Adiabatic steady flow of homogenous solutions was examined in 8 different straight pipes with diameters between 1.8 mm and 21 mm, all with a temperature of 27 °C.

RHEOLOGICAL CHARACTERISTICS AND FLOW RESISTANCE MEASUREMENT RESULTS

In order to identify the rheological characteristics of the analysed solutions, each of the experimental/ pipe flow curves was represented in the form of the functional relationship described by the Equation (1):

(1)

where:

is the shear stress on the pipe wall,

is the pipe shear rate (value of the shear rate on pipe wall).

Interpretation of the experimental results presented in the form of function (1) indicates that the solutions can be successfully approximated with the Ostwald de Waele power-law fluid model. Rheological parameters K (fluid consistency constant) and n (flow behavior index) for each of the analysed solutions are summarized in Table 1. Representative rheological characteristics in the form of experimental/pipe flow curves are illustrated in Figure 2.

Figure 2
Representative experimental flow curves for polymer, surfactant and polymer-surfactant solutions.

Additionally, representative shear viscosity curves vs. shear rate for the analysed solutions of Figure 2 are presented in Figure 3.

Figure 3
Representative shear viscosity curves as a function of shear rate for polymer, surfactant and polymer-surfactant solutions.

The mass fractions of individual additives in the rheologically complex polymer-micellar solution affect the value of the n flow index. This parameter characterizes non-Newtonian properties of a fluid. There is no way to predict a priori, its value for a solution prepared arbitrarily and having different mass fractions of particular drag reducers. It was only observed that the increase of both CTAB and NaSal concentration in the examined solution, at a constant polymer concentration, leads to the intensification of non-Newtonian properties of the fluid, i.e., to the increase of the value of the fluid consistency constant K and the decrease of the n flow index.

The interpretation of experimental data and the assessment of the respective solution additives’ influence on the increase or reduction of the flow resistance and the shape and location of resistance curves depend significantly on the coordinate system adopted for data presentation. Firstly, experimental results of flow resistance are presented in the classical system of dimensionless numbers [Res, cf] described by formulas (2) and (3):

(2)

(3)

and additionally in the form of drag reduction coefficient DR defined as a function of the Reynolds number (2), and described in percentage terms:

(4)

Figure 4 presents the flow resistance curves of polymer, surfactant and polymer-surfactant water solutions, defined in the system of dimensionless numbers (2) and (3). Analyses of flow resistance curves reveal that, in any of the analysed flow ranges, the measurement points do not correspond to the theoretical functions which describe Newtonian fluid flow. The simultaneous addition of even small amounts of high molecular weight polymers and surfactants causes an increase of flow resistance in the laminar range of flow.

Figure 4
The flow resistance curves of polymer, surfactant and polymer-micellar water solutions, defined in the system of cardinal numbers (2) and (3).

The drag reduction coefficient curves (4) are illustrated in Figure 5. It is observed that, in the case of the turbulent flow, simultaneous application of the chemical additives produces the drag reduction effect.

Figure 5
The drag reduction coefficient curves DR = f(Re).

Due to the difficulty in unambiguous determination of the critical value of the Reynolds number Res, the DR values presented in Figure 5 are calculated in such a way that it was assumed that cfs=16/Res in the range of the Reynolds number Res<2100, whereas the formula cfs=0.079/ Res -0.25 was used in the range of Reynolds number Res≥2100. Therefore, a sharp increase of the DR value observed in the transition zone, particularly in the polymer solution (see in Figure 5), does not reflect the actual degree of drag reduction in this range of flow. In the case of the polymer solution flow, in the initial stage of the turbulent flow, no noticeable reduction of flow resistance is observed. Only after exceeding a certain characteristic Reynolds number Res ≈ 1.5×104, the onset of the drag reduction effect occurs and the phenomenon increases with the increase in the value of the Reynolds number. A similar increase in the reduction of the shear resistance for the turbulent flow range, induced with PEO additive, was observed in the drag reduction effect obtained by the use of a rotating-disk apparatus (Kim et al., 2001Kim, C. A., Jo, D. S., Choi, H. J., Kim, C. B., Jhon, M. S., A high-precision rotating disk apparatus for drag reduction characterization. Polymer Testing, 20, 43-48 (2001).).

Furthermore, the value of the critical Reynolds number for which the transition from the laminar flow to the turbulent flow is observed takes various values that depend on pipes diameter, type and concentration of chemical additives introduced into the solvent.

A better interpretation of the simultaneous addition of the polymer and surfactant with salt effect on drag reduction, in comparison with adequate addition of pure polymer or pure surfactant with salt, can be achieved by presentation of the same measurement date in a modified system of “pseudorheostable” numbers [ReM, cfM]. The modified system of dimensionless numbers used in the analysis is described by formulas (5) and (6):

(5)

(6)

As one knows (Matras, 1986), in such a defined dimensionless number system, flow resistance curves of rheostable (purely viscous) non-Newtonian fluids are transformed to a single curve - in the whole range of modified Reynolds number (5) - identical to the classical Newtonian curve described in the laminar range by the Fanning equation and in the turbulent flow by the Blasius formula. Selection of such a coordinate system was dictated additionally by the fact that it facilitates identification and description of the characteristic drag reduction flow zones. In this modified system of pseudorheostable dimensionless numbers [ReM, cfM] each deviation of the experimental flow resistance curve, which indicates abnormal drag reduction from the pseudorheostable Blasius curve, allows for the identification of the influence of specific additives (polymers or/and surfactants with salt) on the range of the drag reduction effect. Figure 6 presents a comparison of flow resistance curves in the modified number system (5) and (6) for 3 types of solutions with different internal structures.

Figure 6
The flow resistance curves of polymer, surfactant and polymer-surfactant water solutions, defined in the system of cardinal numbers (5) and (6).

The results of experimental data analysis indicate that polymer-surfactant additives cause significant drag reduction in a wider range of flow in comparison with pure polymer and pure surfactant solutions. Surfactant and salt additives (micellar solution) induce the appearance of the stable transitional zone, in which a spectacular reduction of flow resistance is observed - usually greater when compared to the same effect achieved with polymer additives. This zone was not observed for the polymer-surfactant solution without the salt additives. Figure 7 demonstrates the influence of NaSal additive on the drag reduction effect. It is clearly evidenced by the flow resistance curves that, with the addition of a small amount of electrolyte (e.g., salts or alcohols), a reformation of spherical micelles into threadlike micelles must proceed. It leads to significant extension of the transitional zone. An efficient drag reduction effect is observed within this zone of flow.

Figure 7
The effect of NaSal additive on the polymer-surfactant flow resistance curve.

In the stable transitional zone, loss of stability of the laminar flow increases gradually when the Reynolds number values grow. In this range the relative drag reduction is the greatest. Beyond a certain second critical value of the Reynolds number the occurrence of an unstable transitional zone is observed. In this range of flow a rapid loss of the drag reduction effect occurs.

Surfactant and salt concentration effects were observed within this zone as illustrated in Figure 8.

Figure 8
Surfactant and salt concentration effects on the flow resistance curves of polymer-micellar aqueous solutions.

An increase of surfactant and salt concentration leads to a further extension of the transitional zone towards greater values of the Reynolds number.

Beyond a certain third critical value of the Reynolds number the fluid starts to behave like a classical rheostable non-Newtonian fluid. An additional abnormal drag reduction zone in the turbulent range is observed. The viscoelastic properties of the solution are a dominant factor. This effect is well illustrated in Figure 9.

Figure 9
Polymer concentration effect on the flow resistance curves of polymer-surfactant aqueous solutions.

In comparison with a purely micellar solution, the collapse of drag reduction is normally observed in this region. It should also be noted that addition to the micellar solution of even small amounts of high molecular weight polymer (about 10ppm) causes a reduction of the non-Newtonian properties of the solution.

The results of drag reduction measurements indicate the effect of pipe diameter on drag reduction efficiency. Figure 10 illustrates the pipe diameter effect on the flow resistance curves of polymer-surfactant water solutions. Increasing the pipe diameter d results in a clear extension of the stable transitional zone towards higher values of the Reynolds number. Moreover, decreasing the pipe diameter value d results in an increase of the drag reduction effect in the third additional turbulent range of flow.

Figure 10
Pipe diameter effect on the flow resistance curves of polymer-surfactant water solutions.

The experimental results reveal that polymer-micellar solutions can be characterized by a lower susceptibility to mechanical degradation during flow or that the degradation can be almost invisible in the analysed range of flow. Figure 11 illustrates a shift of the collapse of the drag reduction effect toward greater values of Reynolds number caused by polymer-surfactant additives.

Figure 11
Shift of the collapse of drag reduction effect toward greater values of Reynolds number caused by polymer-surfactant additives.

Pure PEO solution degrades very fast under high shearing conditions. Collapse of drag reduction is gradually observed in such a case. Experimental results show a considerable increase of the Reynolds number value at which the mechanical degradation of polymer and the collapse of the DR effect are observed in polymer-micellar solution.

HYPOTHETICAL MECHANISM OF DRAG REDUCTION CAUSED BY POLYMER-MICELLAR SURFACTANT ADDITIVES

Most of experimental studies have shown that simultaneous addition of small amounts of polymer and surfactant agents to the solvent causes initiation of the micellarization process at lower concentration, compared to the CMC (Hou et al., 1999Hou, Z., Li, Z. and Wang, H., Interaction between poly(ethylene oxide) and sodium dodecyl sulfonate as studied by surface tension, conductivity, viscosity, electron spin resonance and nuclear magnetic resonance. Colloid Polym. Sci., 277, 1011-1018 (1999).; Matras et al., 2008Matras, Z., Malcher, T. and Gzyl-Malcher, B., The influence of polymer-surfactant aggregates on drag reduction. Thin Solids Films, 516, 8848-8851 (2008).). This concentration is called the critical aggregation concentration (CAC).

In the case of an ionic surfactant mixed with a non-ionic or oppositely-charged polyelectrolyte, only a small part of the polyelectrolyte is absorbed by the surfaces of the micelles. Furthermore, the CAC has a lower magnitude than the original CMC due to the following facts:

  • electrostatic interaction occurs between the polyelectrolyte and the surface of the micelles,

  • hydrophobic interaction occurs between the non-ionic polymer and ionic micelles,

  • surfactant counter-ions on the micelles surface are replaced by the electrolyte or polyelectrolyte,

  • highly charged electrolytes can provide a certain amount of concentrated counter-ions that combine with the micelles.

As a result of electrostatic or hydrophobic interaction, the micelles combine with the polymer chains by coiling around them. The final state of the mixture has single threadlike micelles with a part of the polymer macromolecule chain coiled around rigid micelles. According to Hou et al. (1999)Hou, Z., Li, Z. and Wang, H., Interaction between poly(ethylene oxide) and sodium dodecyl sulfonate as studied by surface tension, conductivity, viscosity, electron spin resonance and nuclear magnetic resonance. Colloid Polym. Sci., 277, 1011-1018 (1999)., Matras et al. (2008)Matras, Z., Malcher, T. and Gzyl-Malcher, B., The influence of polymer-surfactant aggregates on drag reduction. Thin Solids Films, 516, 8848-8851 (2008). and Jung et al. (2011), these form the so-called aggregates.

Such newly created polymer-micellar solutions can be characterized by a lower susceptibility to mechanical degradation during flow or the degradation can be almost invisible in the analysed range of flow.

Aggregates, schematically illustrated in Figure 12(a), subjected to the shear stress, take an orientation consistent with the aforementioned principle of minimum resistance - Figure 12(b). With the increasing value of the Reynolds number, internal friction forces stretch and extend the aggregates, leading to the laminarization of the initial phase of the turbulent flow.

Figure 12
Polymer-micellar aggregate: a) in the shape of spherical structures, b) elongated during shear stress action.

Therefore, it can be hypothesized that the rigid rodlike micelles, which create the core of the aggregates, are responsible for reducing the flow resistance in the extended transitional zone between the laminar and the turbulent flow.

The aggregates and micelles are responsible for transmission of internal friction within the liquid. The value of the critical Reynolds number for which the transition to the turbulent zone is observed is greater for polymer-micellar solutions. This means that the stable transition zone is extended. The reason for such behaviour can be the partial disintegration of aggregates to their original forms, i.e., micelles (formed from the surfactant) and macromolecules (formed from the polymer) due to a significant increase of the shear rate. From this moment, both micelles and macromolecules interact separately in the transported solution, causing a further drag reduction effect. Passing further into the turbulent range of flow, micelles lose their orientation and no longer have a major impact on the drag reduction. A key role is played in this zone by the polymer. Not having undergone an earlier degradation, the polymer macromolecules still cause the flow reduction.

In drag reduction caused by the use of polymer-surfactant solution, one cannot talk about the so-called collapse of the drag reduction. It occurs over a wide range of Reynolds numbers. In the turbulent zone, polymer macromolecules undergo a certain mechanical degradation. Decreasing the shear rate leads to the reconstruction of the internal structure of the solution. As a result of electrostatic or hydrophobic interaction, the recreated micelles are combined with the polymer chains by coiling around them. These chains are much shorter and such newly created aggregates do not have the same rheological properties as the original ones. This results in a slight increase of the flow resistance in comparison with a freshly prepared solution.

CONCLUSIONS

A comparative analysis of the drag reduction efficiency produced by pure high molecular polymers, pure surfactants and surfactants with high molecular weight polymer additives indicates that the simultaneous addition of surfactants together with high molecular polymers causes an increase of flow resistance in the laminar range of flow compared to the analogous flow of pure solvent. In the case of the turbulent flow, simultaneous application of the examined chemical additives produces a drag reduction effect. It produces, however, a significant extension of the stable transitional zone between the laminar flow and the turbulent flow. The surfactant with salt additives has the major influence on efficiency of drag reduction in this zone.

Experimental results prove that the simultaneous addition of surfactants and high molecular weight polymers leads to the occurrence of a third significantly extended drag reduction zone in the turbulent range of flow. The dominant factors in that zone are the viscoelastic characteristics of the solution caused by the presence of polymer macromolecules, wherein an increase of the mass fraction of polymer additive increases the efficiency of the drag reduction effect only in the turbulent range of flow.

Comparative studies demonstrated that the analysed polymer-micellar solutions combine and intensify positive features of their purely polymer or purely micellar analogues providing a more efficient drag reduction effect over a wider range of flow.

A hypothetic mechanism for the drag reduction effect caused by polymer-micellar solutions is proposed. Experimental results for the drag reduction effect are consistent in qualitative terms with the proposed hypothesis and confirm the described mechanism of the phenomenon in an indirect way.

REFERENCES

  • Borostow, W., Drag reduction in flow: Review of applications, mechanism and prediction. Journal of Industrial and Engineiring Chemistry, 14, 409-416, (2008).
  • Cai, S.-P., Drag reduction of a cationic surfactant solution and its shear stress relaxation. Journal of Hydrodynamics, 24(2), 202-206 (2012).
  • Diamant, H. and Andelman, D., Onset of self-assembly in polymer-surfactant systems. Europhysics Letters, 48(2), 170-176 (1999).
  • Dujmovich, T. and Gallegos, A., Drag reducers improve throughput, cut costs. Offshore, 65(12), pp. 55-58 (2005).
  • Figueredo, C. R. and Sabadini, E., Firefighting foam stability: The effect of the drag reducer poly(ethylene) oxide. Colloids and Surfaces, A: Physicochemical and Engineering Aspects, 215(1-3), 77-86 (2003).
  • Goddard, E. D. and Ananthapadmanabhan, K. P., Interactions of Surfactants with Polymers and Proteins. CRC Press, Boca Raton, FL (1993).
  • Hou, Z., Li, Z. and Wang, H., Interaction between poly(ethylene oxide) and sodium dodecyl sulfonate as studied by surface tension, conductivity, viscosity, electron spin resonance and nuclear magnetic resonance. Colloid Polym. Sci., 277, 1011-1018 (1999).
  • Hadri, F., Besq, A., Guillou, S. and Makhloufi, R., Drag reduction with an aqueous solution of CTAC-NaSal: Study of the wall slip with a Couette geometry. Comptes Rendus Mécanique, 338(3), 152-157 (2010).
  • Kim, C. A., Jo, D. S., Choi, H. J., Kim, C. B., Jhon, M. S., A high-precision rotating disk apparatus for drag reduction characterization. Polymer Testing, 20, 43-48 (2001).
  • Kim, J. T., Kim, C. A., Zhang, K., Jang, C. H. and Choi, H. J., Effect of polymer-surfactant interaction on its turbulent drag reduction. Colloids and Surfaces A: Physicochemical and Engineering Aspects, 391(1-3), 125-129 (2011).
  • Jönsson, B., Lindman, B., Holmberg, K. and Kronberg, B., Surfactants and Polymers in Aqueous Solution. John Wiley & Sons Chichester, UK (1998).
  • Liberatore, M. W., Baik, S., Mchugh, A. J. and Hanratty, T. J., Turbulent drag reduction of polyacrylamide solutions: Effect of degradation on molecular weight distribution. J. Non-Newtonian Fluid Mech., 123, 175-183 (2004).
  • Matras, Z., Przepływ cieczu Tomsa w przewodach kołowych. Politechnika Krakowska, Monografia 29 (1984). (In Polish).
  • Matras, Z. and Walczak, S., The capillary-pipe rheometer for the identification of complex properties of multiphase non-Newtonian fluids. Inżynieria i Aparatura Chemiczna, 6, 150-151 (2006).
  • Matras, Z., Malcher, T. and Gzyl-Malcher, B., The influence of polymer-surfactant aggregates on drag reduction. Thin Solids Films, 516, 8848-8851 (2008).
  • Minatti, E., Zanette, D., Salt effects on the interaction of poly(ethylene oxide) and sodium dodecyl sulfate measured by conductivity. Colloids Surfaces, A: Phisicochem. Eng. Aspects, 113, 237 (1996).
  • Mohsenipour, A. A., Pal, R. and Prajapati, K., Effect of cationic surfactant addition on the drag reduction behaviour of anionic polymer solutions. The Canadian Journal of Chemical Engineering, 91(1), 181-189 (2013).
  • Motier, J. F., Chou, L. C. and Kommareddi, N. S., Commercial drag reduction: Past, present, and future. Proceedings of the ASME Fluids Engineering Division Summer Meeting, San Diego, Calif., USA (1996).
  • Suksamranchit, S., Sirivat, A. and Jamieson, A. M., Polymer-surfactant complex formation and its effect on turbulent wall shear stress. Journal of Colloid and Interface Science, 294(1), 212-221 (2006).
  • Tamano, S., Ito, M., Kato, K. and Yokota, K., Turbulent drag reduction in nonionic surfactant solutions. Physics of Fluids, 22(5), 055102 (2010).
  • Virk, P. S., Drag reduction fundamentals. AIChE Journal, 21 (4), 625-656 (1975).
  • Wang, Y., Yu, B., Zakin, J. L. and Shi, H., Review on drag reduction and its heat transfer by additives. Advances in Mechanical Engineering, 10, 17 pages (2011).
  • Warholic, M. D., Heist, D. K., Katcher, M. and Hanratty, T. J., A study with particle- image velocimetry of the influence of drag-reducing polymers on the structure of turbulence. Exp. Fluids, 31, 474-483 (2001).
  • White, C. M. and Mungal, M. G., Mechanics and predictions of turbulent drag reduction with polymer additives. Annular Review of Fluid Mechanics, 40, 235-256 (2008).
  • Zadrazil, I., Bismarck, A., Hewitt, G. F. and Markides, C. N., Shear layers in the turbulent pipe flow of drag reducing polymer solutions. Chemical Engineering Science, 72, 142-154 (2012).

Publication Dates

  • Publication in this collection
    Oct-Dec 2016

History

  • Received
    10 July 2015
  • Reviewed
    23 Aug 2015
  • Accepted
    26 Aug 2015
Brazilian Society of Chemical Engineering Rua Líbero Badaró, 152 , 11. and., 01008-903 São Paulo SP Brazil, Tel.: +55 11 3107-8747, Fax.: +55 11 3104-4649, Fax: +55 11 3104-4649 - São Paulo - SP - Brazil
E-mail: rgiudici@usp.br