Vous êtes sur la page 1sur 42

Chapter 6 DYNAMIC BREAKUP OF LIQUID-LIQUID JETS

Many, it may even be said, most of the still unexplained phenomena of Acoustics are connected with the instability of jets of uid. For this instability there are two causes; the rst is operative in the case of jets of heavy liquids, e.g., water, projected into air (whose relative density is negligible), and has been investigated by Plateau in his admirable researches on the gures of a liquid mass, withdrawn from the action of gravity. It consists in the operation of the capillary force, whose eect is to render the innite cylinder an unstable form of equilibrium and to favour its disintegration into detached masses whose aggregate surface is less than that of the cylinder. The other cause of instability, which is operative even when the jet and its environment are of the same material, is of a more dynamical character. Lord Rayleigh, On the instability of jets (1879) 6.1 Background The liquid-liquid jets discussed in chapter 5 eventually break up due to increasing amplitude of disturbance waves on their surface. In this chapter we investigate the length of the resulting jets, which is dependent on many factors. The axisymmetric, dynamic breakup of a Newtonian liquid jet injected vertically into another immiscible Newtonian liquid at various Reynolds numbers is investigated. When a liquid is injected into another liquid at a velocity above a certain critical value, a jet forms, rises to a certain length, and then breaks up into drops (Meister, 1966; Meister and Scheele, 1967; Scheele and Meister, 1968; Meister and Scheele, 1969a, 1969b; Richards, 1978; Richards and Scheele, 1985; Richards et al., 1993; Richards et al., 1994a). These drops result in the creation of large new surface area, which leads to enhanced heat and mass transfer (Skelland and Walker, 1989). 107

108 This has immediate implications for design of, for example, sieve-tray liquid-liquid extractors. The prediction of the jet dynamics is thus important as a means to explore both quantitative aspects of performance, e.g., to calculate the size of the drops formed, and qualitative features, such as a test of stability theories. The technique applied to steady laminar jets as described in chapter 5 has been slightly modied to represent better the physical stability limits of the system, and is used in the present chapter to predict jet length and subsequent breakup into drops. Jet breakup has been extensively studied, both experimentally and theoretically, for jets injected into air with the air modelled as a vacuum or an inviscid uid; a brief review up to 1990 is given by Mansour and Lundgren (1990). Theoretical developments started with the elegant analysis of Rayleigh (1879) for the case of an innite free inviscid jet injected into air. In the liquid-liquid jet problem additional, non-trivial, eects enter due to the presence of an outside continuous phase. The Rayleigh stability analysis was extended by Tomotika (1935) to a viscous cylinder surrounded by a viscous uid. However, he solved the general eigenvalue relation only for the limiting case of a very viscous liquid in another very viscous liquid. The general eigenvalue equation was rst solved numerically by Meister and Scheele (1967) and covered all the previous special cases such as a viscous liquid jet in a gas by Weber (1931) and an inviscid liquid jet in an inviscid liquid by Christiansen and Hixson (1955, 1957). In a series of publications, Meister and Scheele (Meister, 1966; Meister and Scheele, 1967; Scheele and Meister, 1968; Meister and Scheele, 1969a, 1969b) worked to develop an understanding of the jet and drop formation based on experiments obtained with 15 liquid-liquid systems. They described the general behavior of the jets as follows. For low ow rates drops form, grow, and break o from the nozzle at regular intervals. Above a certain critical velocity a jet is formed that rises to a certain length from the nozzle at which point it breaks up into drops. At still higher nozzle velocities, the jet disrupts into small drops.

109 Much of the earlier literature on liquid-liquid jet breakup has been reviewed by Meister and Scheele (1969a) and more recently by Skelland and Walker (1989) (up to 1989). Meister and Scheele (1969a), using the Tomotika (1935) stability equation, developed an expression for the jet length at breakup that was an improvement over that used by Smith and Moss (1917). This equation involved a Galilean translation of the innite Tomotika stationary jet at an average velocity of the interface, calculated by assuming a certain form for the velocity prole inside the jet and by solving for the unknown coecients invoking overall momentum and mass balances. We believe that this theory, although developed in the late 60s, still represents the best one available for engineering design estimates of the jet length. For liquid jets injected into air, Sterling and Sleicher (1975) included in the analysis the aerodynamic interaction between the jet and the air and found that the presence of the air led to an enhanced growth rate of the axisymmetric disturbances. In particular, they examined the case of a viscous liquid injected into a stationary inviscid medium, and they developed an equation for the growth rate which was an improvement over the one used by Weber (1931). Several attempts have been made to include in the stability analysis the relative motion of the continuous phase in liquid-liquid systems, the most recent being those of Kitamura et al. (1982), Bright (1985), and Russo and Steen (1989). Kitamura et al. (1982) experimentally varied the motion of the continuous phase to be either faster, the same as, or slower than the jet down to the case of a stagnant continuous phase. They found that the jet shortened as the absolute value of the continuous phase velocity relative to the jet increased from zero. Moreover, the condition of a zero relative phase velocity was found to give the best agreement between Tomotikas stationary jet solution (translated at the average velocity of the jet) and the experimental measurements. Thus, they attributed the discrepancy between experiments and Tomotikas analysis in the nonzero relative velocity cases to the

110 relative motion of the continuous phase, rather than to other factors, such as, for example, a velocity prole relaxation. Bright (1985) attempted to perform a linear viscous stability analysis assuming a constant, but unequal, velocity in each liquid phase. One diculty with this study rests on the inconsistency of having a discontinuity of velocity at the interface for uids that are not assumed to be inviscid. Even assuming that the equations are formulated correctly, we have found that the particular solution provided by Bright (a nonhomogeneous modied Bessel ODE) is incorrect; its correct form is much more complex than the one provided by his equation (18). Details of the derivation can be found in appendix H. Russo and Steen (1989), in a study on liquid bridge stability, performed numerically a linear stability analysis for an innite cylinder of viscous uid with a constant shear applied to the outer free surface. From their predicted growth rates, they found qualitative agreement with the data of Meister and Scheele (1969a), but they noted that they could not expect quantitative agreement since they neglected not only the relaxation of the velocity prole along the jet, but also the possibility of spatially growing instabilities (Russo and Steen, 1989). Linear stability analyses can be further characterized as temporal or spatial. Temporal stability analyses (e.g., Tomotika, 1935; Christiansen and Hixson, 1957; Mansour and Lundgren, 1990) examine how spatially periodic disturbances grow in time. In contrast, spatial stability studies assume that disturbances starting at the nozzle tip are temporally periodic waves that travel spatially with the jet (Keller et al., 1973; Busker et al., 1989; Russo and Steen, 1989). The most severe assumption is that of a uniform (at least in the axial direction) base ow. Furthermore, note that the usefulness of eigenvalue analysis, through linear stability analysis, for a non-self-adjoint system of equations (e.g., the Navier-Stokes equations) has been recently questioned as far as its relevance to the time evolution of nite amplitude disturbances is concerned, by the discovery of solutions to the subcritical linearized problem of Poiseuille and Couette ow that grow greatly even though

111 all eigenmodes decay monotonically (Trefethen et al., 1993). Thus, all the previous liquid-liquid stability theories, as based on rather drastic simplifying assumptions, have limited validity. As was observed by Meister and Scheele (1967) regarding previous stability analyses, the agreement between experimental results and theoretical instability analyses has been good when experiments are performed which closely approximate the assumptions made in the theoretical development. The diculties are associated with the various eects (such as viscous, buoyancy, surface tension, inertial forces, jet contraction, velocity prole relaxation, and relative motion of the continuous phase) that are not all fully accounted for in any of the previous theories. This lack of an adequate theoretical description which is driven by the inherent complexity of the problem led us to the present direct numerical simulation of the jet breakup process. The only assumptions involved with the present approach are those of laminar ow of the two Newtonian uids with constant bulk densities and viscosities, constant interfacial surface tension, and axisymmetric disturbances. Of these the latter assumption has been shown experimentally to fail in the region above a certain critical velocity where the jet lengthens appreciably to a maximum length (Meister and Scheele, 1969a). However, in the region below this critical velocity, experimental observation (Meister and Scheele, 1969a) shows that axisymmetric disturbances dominate, and in consequence, this is the region of validity of the present study (see Figure 7.1 which is Figure 1 of Meister and Scheele, 1969a). The motivation for selecting the VOF method in this numerical study has been elaborated previously in chapters 1 and 3 (Hirt, 1968; Nichols et al., 1980; Hirt and Nichols, 1981; Kothe et al., 1991; Brackbill et al., 1992; Richards et al., 1993). Primarily, the present numerical algorithm can capture the irregular shapes of free surfaces, such as those that are realized during necking and detachment of drops from the jet; this has not been possible with other methods devised for single phase jets without considerable modication of the algorithms. For example, Mansour and Lundgren (1990) followed the breakup of an initially stationary

112 cylinder of inviscid liquid using a nonlinear boundary element method. They were able to follow the solution until necking just before breakup; then, in a second stage of calculation, they cut the forming satellite drop at the throat and continued the development of the satellite shape. Boundary element techniques could be used, but they would involve a restart after every necking-drop detachment and a rather elaborate remeshing to capture the inertial eects. In contrast, the VOF method allows very naturally arbitrary jet and drop shapes to be calculated. The objectives of the present study are to use the VOF based numerical method that we have developed for calculating liquid-liquid jet dynamics, from the startup of their time evolution to their apparent (or pseudo) steady-state involving the region from the nozzle exit to their breakup through necking and detachment of drops. Our aim is to predict the jet length and the shape, as well as the subsequent shape and size of drops that are formed. We compare the results with the available experimental data of Meister and Scheele (1969a) and Bright (1985) and the available linear stability theory of Meister and Scheele (1969a). We also show that quantitative agreement of the present model is obtained with the dynamic linear initial-value problem of Christiansen and Hixson (1955, 1957). In section 6.2 we present the denition and formulation of the governing equations for the solution to the problem of a Newtonian liquid jet injected vertically into another Newtonian quiescent liquid. The analytical solution to the linear inviscid liquid-liquid cylinder breakup problem is illustrated, and the stability analyses of Tomotika (1935) and Meister and Scheele (1969a) for viscous liquid-liquid jets are discussed. In section 6.3 we then discuss the numerical implementation of the governing equations to the liquid-liquid jet problem. In section 6.4 the numerical simulation is compared with the inviscid liquid-liquid cylinder breakup solution. The present numerical simulation of free surface dynamics and jet lengths is compared to the existing experimental data (Meister and Scheele, 1969a; Bright, 1985) and the Meister and Scheele (1969a) stability analysis. Finally, in section 6.5 we reach conclusions based on the results of the present work.

113 6.2 Problem Denition and Formulation The ow conguration investigated here is shown in Figure 6.1. It corresponds to the experimental setup of Meister and Scheele (1966, 1969a) and Bright (1985), details of which can be found in those references. Briey, the problem investigated involves the jet of an alkane (either n-heptane or n-decane) injected vertically from a circular nozzle upwards into a tank of stationary mutually saturated immiscible water. Figure 6.1 shows the stationary tank of uid 1 (water) with density 1 and viscosity 1 with the jet of uid 2 (alkane) owing upward with density 2 and viscosity 2 , from a nozzle of inner radius R and outer radius Ro with average velocity v . The constant interfacial surface tension is , and the gravitational acceleration, g , is directed downward. The distance from the nozzle tip to the top of the tank is L1 and to the bottom is L2 . The distance from the centerline of the nozzle to the outer tank wall is L3 and the nozzle is of total length L4 with the velocity prole fully developed at length L5 from the nozzle tip. Disturbances originating in the nozzle propagate with the jet ow and eventually lead to breakup of the jet, with the length of the jet given by L, and the wavelength of the growing disturbance given by .

6.2.1 Governing Equations Equations (3.1) to (3.13) are solved with appropriate boundary conditions as discussed in section 5.2.1. For the nozzle and the bottom of the tank no-slip conditions are used: u=v=0 For the axis of symmetry at r = 0: v =0 r (6.1)

u = 0,

(6.2)

114

Outflow Boundary Continuous Fluid 1 F=0 1, 1

Tank Wall

Dispersed Fluid 2 F=1 2, 2 Axis of Symmetry

L g z r

L1

Jet Interface r=a(z)

L3

L5

L2 L4

Inflow Boundary v

R Ro

Nozzle

Figure 6.1: Liquid-liquid jet ow conguration (not to scale).

115 For the inow into the nozzle, at distance L5 from the nozzle tip, fully developed Poiseuille ow is assumed: u = 0, v = 2 v 1 r
2

(6.3)

where r r/R is the dimensionless radial distance, and v is the average velocity in the nozzle. Dimensionless variables in this work are denoted with an asterisk using R as the characteristic length and v as the characteristic velocity. The experimental nozzle length ratio L4 /R was assumed to be suciently large to guarantee condition (6.3) (Meister and Scheele, 1969a). In the VOF approach, uid is allowed to ow through the mesh with a minimum of upstream inuence. For the outow boundary at the top of the mesh, z = L1 , it is assumed that there is no change in the axial direction (Nichols et al., 1980; Hirt and Nichols, 1981; Richards et al., 1993): u v = =0 (6.4) z z To allow for the presence of a lateral wall far away from the nozzle (L3 /R 1) without increasing the computational grid size excessively, it is as-

sumed that there is no change along the radial direction at the outer boundary of the computational grid at r = L3 , L3 < L3 : u v = =0 r r (6.5)

The numerical implications of using the boundary condition (6.5) to represent the outer lateral wall (as opposed to a no-slip condition used in the previous work, Richards et al., 1993 ) are discussed in section 6.3. The jet interface is assumed to be pinned at the nozzle lip, a fact that is observed experimentally, so no contact angle needs to be specied in this problem. 6.2.2 Inviscid Free Liquid-Liquid Cylinder Breakup To test the accuracy and performance of our method on the dynamics of jet breakup, it was rst applied to the problem of the capillary breakup of an

116 initially stationary innitely long inviscid liquid cylinder inside another inviscid liquid within an outer free-slip wall. This problem was chosen because it has a closed form analytical solution which was rst derived for arbitrary values of the initial conditions by Christiansen and Hixson (1955, 1957). The detailed derivation is given in appendix F.1. We use the general solution of Christiansen and Hixson (1957) to solve the initial value problem with initial conditions a = R + a0 cos(kz ) and da/dt = 0 at t = 0, where a0 is the initial amplitude, s is the growth rate, and k = 2/ is the axial wavenumber of the disturbance. The particular solution expressed in terms of the velocity potential i , where vi = i , i = 1, 2 in both phases becomes: a = R + a0 cosh (st) cos (kz ) 1 = a0 s k 2 = where: s2 = R 3 2 K0 (kr) + I0 (kr) sinh (st) cos (kz ) K1 () I1 () a0 s k I0 (kr) I1 () sinh (st) cos (kz ) (6.6) (6.7) (6.8)

1 2
I0 () I1 ()

+ 1

K0 () + I0 () K1 () I1 ()

(6.9)

K1 ( )/I1 ( ), kR, and kL3 . Note that this equation reduces to Rayleighs (1879) particular result for an inviscid liquid in a gas if the density of the outer uid is taken to be zero, 1 = 0. Also note that as the outer wall radius is increased, as 0. 6.2.3 Viscous Free Liquid-Liquid Cylinder Breakup The linear stability analysis to calculate the growth rate, s, of a disturbance on an initially stationary, innitely long, viscous liquid cylinder in another viscous liquid was rst derived by Tomotika (1935). The detailed derivation is given in appendix G. His solution is derived by the introduction of the streamfunction u = (1/r) (/z ), v = (1/r) (/r), and by the substitution of normal mode solutions of the form a = R + a (r) exp (ikz + st) into the linearized equations,

117 where a (r) is purely a function of r. The determinant of the coecients of the equations for the growth rate is: I1 () I0 () 2k2 2 I1 () 1 f1 where: f1 = 2k2 2 1 s2 dI1 () k + I0 () + 2 1 I1 () 2 d 1 s1 R 2 dI1 ( ) k + 2 1 I1 ( ) 1 d s1 R2 dK1 () s1 K0 () d 1 (6.11) I1 ( ) K1 () I0 ( ) K0 () 2 2 k2 +m2 2 1 I1 ( ) 2k K1 () f2 f3 K1 ( ) K0 ( ) =0 k2 + m2 1 K1 ( ) f4 (6.10)

f2 = 2km2

(6.12) (6.13) (6.14) (6.15) (6.16)

f3 = 2k2

f4 = 2km1

dK1 ( ) d s1 2 m2 1 =k + 1 s2 2 m2 2 =k + 2

with Rm1 , and Rm2 . This quadratic equation (6.10) is solved for the growth rate s and the dimensionless wavenumber is varied until the maximum value of s is obtained. However, an iterative procedure is necessary since m1 and m2 depend on s (Meister, 1966). This maximum value of s corresponds to the most dangerous wavenumber . Note that the determinant equation (6.10) reduces to the eigenvalue relation (6.9) of the previous section as 1 0 and 2 0 (Meister, 1966). 6.2.4 Meister & Scheeles Linear Stability Analysis for Jet Length Meister and Scheeles (1969a) simplied model (MS) for the jet length, L (see Figure 6.1), is now briey reviewed, both as a convenience to the reader, and

118 to describe the exact algorithm used in this chapter to estimate L based on this theory. Further details can be found in Meister and Scheele (1969a). This theory improves a previous estimate by Smith and Moss (1917) based on the amplication of the critical disturbance obtained from the linear stability theory of the free jet problem: L= R v ln s a0 (6.17)

where v is the average nozzle velocity and s is the maximum growth rate obtained from the Tomotika analysis, equation (6.10). By assuming that the disturbance travels not with the nozzle velocity v as implied by (6.17), but at the speed of the interfacial velocity vI and by adjusting the interfacial velocity to that of a noncontracting jet, Meister and Scheele (1969a) estimated the jet length to be: v 2s vI vA vI vA R a0

L=

+
z =5

ln
L z= R

(6.18)

where vI and vA v /a 2 are the axial interfacial and average velocity respectively, a a/R is the dimensionless interfacial radius, z z/R is the dimensionless axial distance. An iterative calculation is required to evaluate the ratio vI /vA at z = L/R since this ratio is a function of L; only two iterations are usually necessary. The evaluation of this ratio at z = 5 (which is approximately z /2) and z = L/R represents an arithmetic average at these two characteristic points. Here we have substituted Meister and Scheeles equation (21) for vI into their equation (22) for L to show that the correction to the nozzle velocity does not depend on a directly, as is implied by their jet contraction equation (15). Values of a are necessary only if values of vI are desired for comparison to experiment using equation (6.26) to be presented below. If vI = vA in equation (6.18), we recover the equation of Smith and Moss (1917) (6.17). The ratio vI /vA is obtained from an assumed velocity prole in the jet

119 evaluated at the interface:


vI = 1 + eA(z )z vA

1 eB (z

)z

(6.19)

The authors assumed velocity distributions both in the jet and in the continuous phase involving four unknown functions of the axial distance, z . They used continuity of mass in the jet, continuity of axial velocity at the interface, continuity of tangential shear stress at the interface, and an approximate axial jump momentum balance across the interface to evaluate these unknown functions. They then arrived at the following equations: 1 + e
Az

1 1 e 2

Bz

=1

(6.20)

G 1 e

Bz

dB dA 2B + 2z dz A+z dz

1 e

Bz

(2 e

Bz ) e Bz

=1

(6.21)

where G Rv (1 2 ) 1 /2 2 . Since the method of solution of these two equations was not given the following approach was taken here. Equation (6.20) is dierentiated with respect to z , and the resulting equation and equation (6.21) are solved directly for the derivatives of the unknown functions A(z ), B (z ) as follows: dA = dz 4AGe4Bz AGeAz + 11AG e3Bz + (3AG2) eAz + 11AG 2 e2Bz (3AG1) eAz + 5AG 1 eBz + AGeAz + AG 4e4Bz eAz + 11 e3Bz + 3eAz + 11 e2Bz Gz 3eAz + 5 eBz + eAz + 1

(6.22)

120 dB = dz 4BGe4Bz BGeAz + 11BG + 4 e3Bz + 3BGeAz + 11BG + 4 e2Bz Gz 3BGeAz + 5BG + 1 eBz + BGeAz + BG 4e4Bz eAz + 11 e3Bz + 3eAz + 11 e2Bz 3eAz + 5 eBz + eAz + 1

(6.23)

These ODEs are integrated numerically for z > 0 subject to initial conditions as z 0: A= 3 G 3 2G
1 3

z 3

(6.24)

1 3

B=

z 3

(6.25)

The interfacial velocity vI /v can then be calculated from: vI 1 = 2 v a vI vA (6.26)

where the dimensionless jet radius a a/R is obtained by using the macroscopic mechanical energy balance for free jets developed by Addison and Elliott (1950) (assuming an initially at velocity prole at the nozzle, and by neglecting viscous dissipation terms), which was found useful as a rough estimate close to the nozzle in chapter 5. Nj a +
4

4 4 3 4 a a + a 1 = 0 We

(6.27)

Here Nj

Re2 1 1 , the buoyancy number appropriate for a liquid-liquid F r 2 jet pointed upwards (+) or downwards () respectively, Re2 2R2v /2 , is the dispersed phase Reynolds number, F r v 2 /2Rg, is the Froude number, W e 2Rv 2 2 /, is the Weber number, and z /Re2 is the axial position downstream scaled with the dispersed phase Reynolds number. Note that equation

121 (6.27) diers from the one used by Meister and Scheele (1969a) who erroneously used an 8 instead of a 4 multiplying the second term. 6.3 Numerical Implementation The method discussed in chapter 3, after validation on several test problems, is used to calculate the axisymmetric dynamic ow and breakup of a liquidliquid jet. For the breakup studies, a temporal sinusoidal perturbation of amplitude a0 and angular frequency was introduced in the F function at the nozzle lip. Other disturbances such as perturbations on the inlet velocity prole (6.3) are also possible but have not been used here. Two important factors in the numerical calculation of jet breakup that must be mentioned here are the method used to calculate the surface tension volumetric forces, and the upwinding method in the nite dierence approximation of the inertial terms contribution to the momentum equations (3.2) and (3.3). The original SOLA-VOF of Nichols et al. (1980) algorithm for liquid-liquid surface tension force calculation was found to be completely unstable for liquid-liquid jet calculations. This motivated the incorporation of the CSF algorithm (Kothe et al., 1991) into the code, which resulted in a stable calculation for the steadystate region of the nozzle (Richards et al., 1993). A critical feature for numerical stability is some smoothing of the interface, especially important given the discrete character of the interface shape approximation. Spatial smoothing of the F

function for calculation of the curvature, equation (3.5), is possible in the CSF algorithm (see section 3.5), but this can lead to excessive stabilization of the jet, preventing its breakup. A minimal amount of smoothing (one pass through the spatial lter in the CSF algorithm) was used in the present work. Similarly, if the inertial terms are approximated by second-order accurate central dierencing, the algorithm is numerically unstable (Hirt, 1968) If they are approximated by rstorder accurate upwinding or donor-cell dierencing, the algorithm is stable but it might be overstable due to the presence of excessive articial viscosity

122 (see section 3.2). 6.4 Results and Discussion Several free surface test problems that have known analytical or numerical solutions were used to test the accuracy and validate the algorithm; the results for the inviscid free liquid-liquid cylinder breakup of Christiansen and Hixson (1955, 1957) discussed in section 6.2.2 are discussed in more detail in section 6.4.1 below. As previously discussed in section 6.1, Meister and Scheele (1969a) measured the jet length, L, vs. the dispersed phase Reynolds number, Re2 , for ve nozzle diameters. Their results, obtained by averaging the lengths of jets measured from four photographs per point, are summarized in Figure 7.1 which is Figure 1 of Meister and Scheele (1969a). We have used their data as well as the photograph in Figure 1 of Bright (1985) for comparison purposes. The parameters corresponding to the six base cases that we have used in this work are summarized in Table 6.1, encompassing six nozzles and two liquid-liquid systems. The systems have been allowed to become mutually saturated (to avoid mass transfer between the phases) before the experiments, and the physical parameters for the saturated systems are given in Table 6.2. In Table 6.2 note the substantial dierences between the experimentally measured and the literature values for the interfacial surface tension. The most probable cause for this dierence is the presence of (unknown) contaminants. Since the emerging jet forms a new, clean surface that under the experimental circumstances is not anticipated to have had enough time to become saturated with the contaminant, the literature values for pure components were used for the interfacial surface tension in this work. However, possible additional implications are discussed later in this section. Five independent dimensionless groups result from the dimensionless form of equations (3.1)-(6.1). /2 , Those are listed in Table 6.1 as Re2 2R2v the dispersed phase Reynolds number, Re1 2R1v /1 , the continuous phase Reynolds number, F r v 2 /2Rg, the Froude number, W e 2Rv 2 2 / , the

123 Table 6.1: Experimental systems studied. Data of Continuous, 1 Dispersed, 2 Nozzle R, cm Ro , cm v , cm/s L1 /R L3 /R L4 /R Re2 Re1 We Fr Nj 2 /1 2 /1 Ca2 ,a rad/s
a

Meister and Scheele (1969a) water n-heptane 1 0.04065 0.0635 48.7 1500 375 3001 688 412 2.577 2 0.08 0.1055 29 762 191 1525 806 482 1.799 3 0.127 0.161 21.6 480 120 961 953 570 1.584 4 0.166 0.24 24.7 367 92 735 1425 853 2.707 5 0.344 0.51 15.48 177 44 355 1851 1107 2.204 0.3552 2388.2 0.41 0.686

Bright (1985)

n-decane 6 0.08 44 521 639 4.36 12.3386 15.4 0.9 0.733 0.00838 184

29.7472 5.3599 10.6 0.41 0.686 68.9 0.41 0.686

1.8731 1.8739 233.3 0.41 0.686 348.5 0.41 0.686

0.00375 0.00223 0.00166 0.0019 0.00119 400 121 57 50 15

Perturbation angular frequency used in this work.

Weber number, and Nj =

Re2 Fr

1 2

1 , the buoyancy number appropriate for a

liquid-liquid jet pointed upwards. The Re2 , Re1 numbers represent the ratio of inertial to viscous forces, F r represents the ratio of inertial to gravitational forces, W e represents the ratio of inertial to surface forces, and Nj represents the ratio

124 Table 6.2: Physical properties. Continuous, 1 Dispersed, 2 , pure,a dyne/cm , exp.,b dyne/cm 2 , g/cm3 1 , g/cm3 2 , g/(cm s) 1 , g/(cm s)
a b c d e

water n-heptane 51.1c 36.2d 0.683d 0.996d 0.00393d 0.00958d n-decane 52c 22.5e 0.732e 0.999e 0.0099e 0.011e

Pure value used in this work. Measured value used in MS model. Johnson (1966) Scheele and Meister (1968) Bright (1985)

of buoyancy to viscous forces. Alternatively, one or more of these dimensionless groups can be replaced by other dimensionless parameter combinations, such as the viscosity ratio 2 /1 , or the capillary number, Ca2 W e/Re2 , representing the ratio of viscous to surface forces. In the numerical simulations, the ow rate at the nozzle was instantaneously increased from zero to its nal value with the alkane lling the nozzle and water lling the tank, in much the same way as the physical system was started. Then the simulation followed the time evolution of the ow until the jet length vs. time prole reached pseudo-steady behavior. A typical (coarse) mesh is shown in Figure 6.2, with the increased mesh renement conned, as illustrated, to the dispersed phase. A locally rened, but uniform, mesh is used inside the jet region, whereas a variable, coarser, mesh is employed in the much slower moving outer

125 region. The mesh size that was used for the liquid-liquid jet runs involved 55 cells stacked in the radial direction (with 40 cells being conned to the interval 0 r 2) and 255 cells in the axial direction, with the exception of nozzle 1 and nozzle 4, which involved 65 cells in the radial direction. This cell distribution was optimized through mesh sensitivity studies. In addition, the eects of the euent boundary distance L1 , the bottom distance, L2 , the fully developed velocity prole distance, L5 , and the outer wall distance L3 were investigated separately by sensitivity studies. The actual tank dimensions L1 /R and L3 /R are given in Table 6.1. The bottom distance was not given in the reference, but was estimated as at least L2 /R > 10. In any case, Yu and Scheele (1975) have found experimentally that placing an annular disk ush with the nozzle tip (L2 = 0) had little eect on the measured jet radii, so this dimension appears to be of little importance. For the conditions of Table 6.1, it was found that L1 /R = 50, L2 /R = 1, L3 /R = 12, and L5 /R = 1 were large enough to establish insensitivity of the results to the actual values of L1 /R, L2 /R, and L3 /R. For example, comparing results obtained with L3 /R = 6 to 12, no signicant dierence was found in the breakup length. Likewise, because of the high speed of the jet, comparing results corresponding to L1 /R = 25 to 50 for nozzle 5, no signicant dierence was found in the dispersed phase solution as long as the jet breakup occurs far from the outow, as was mostly the case. For the high velocity jet cases it was found necessary to modify the outow condition (6.4) by forcing only ow out of the mesh at the outow boundary z = L5 , i.e., by requiring v 0, a method suggested by Hirt (1993). This was found to alleviate a problem with the jet initially hitting the outow reective boundary and causing a signicant backow and consequent jet disruption. Further, it was found that it was preferable to use the boundary conditions (6.5) at the far outer radial region rather than the no-slip condition (e.g., (6.1)) used in chapter 5. More specically, it was found that since the axial domain is longer in this study than the previous one, use of a no-slip condition led to the development of a recirculation

126

50

Symmetry Axis

z*

Jet Interface

0 -1

1 r*

12 Nozzle

Figure 6.2: Coarse mesh layout.

127 in the outer continuous phase. Since this outer boundary at L3 is actually closer to the nozzle than the actual tank wall at L3 , the recirculation does not exist in the physical system. Thus, a more reasonable physical choice, which does not produce this recirculation (i.e., (6.5)), was to use the fact that the radial gradients of velocity diminish rapidly with distance from the nozzle. The breakup of the jets occurs experimentally as a result of naturally occurring perturbations. Given the uncertainty of its exact form in the reported experiments, we added a numerical perturbation to the free surface at the nozzle lip as discussed in section 6.3. The angular frequencies of the perturbations applied are given in Table 6.1. As was rst pointed out by Rayleigh (1879), the most dangerous frequency, corresponding to the fastest growing disturbance on the jet, dominates all others, and it becomes the one observed in experiments. We have estimated this frequency in the following way. We calculated from the linear theory equation (6.10) the most dangerous wavenumber = kR. Then assuming that the average interfacial velocity vI was about half the nozzle velocity (Bright, 1985), the angular frequency was estimated as = k vI . This estimate is close enough to the most dangerous frequency since a sensitivity study on the nozzle 3 case revealed that the jet length decreases by only 19% as is varied from 28 to 114 rad/s, with 57 rad/s being the base angular frequency. The amplitude of the perturbation, a0 /R = 0.01, was xed for all the present simulation cases. Its variation has a varying inuence on the results, depending on the Reynolds number. In fact, for the high Reynolds number, Re2 , cases such as nozzle 5, no perturbation was necessary, there being enough numerical noise to break up the jet. However, the low Re2 jets, such as nozzle 1, required a nite perturbation to break within L/R 50; the larger the amplitude of the perturbation, the shorter the jet length, in agreement with the logarithmic dependence of the perturbation given in equation (6.18). In calculating the

predicted jet lengths from the linear theory of Meister and Scheele (1969a), we have used their recommended value, ln(R/a0) = 6, corresponding to a value

128 of a0 /R = 0.0025. The use of a0 /R = 0.01 would have resulted in a wider disagreement of the MS model and the data. The variation of this parameter in the present simulation is discussed in section 6.4.4. The value of interfacial surface tension used in the present simulations was the pure value taken from the literature and listed in Table 6.2. Meister and Scheele (1968) used the Harkins and Brown (1919) drop volume technique to measure the interfacial surface tension, which is an equilibrium or static measurement. However, Bohr (1909) recognized that a dynamic measurement of surface tension was needed to describe new, freshly formed surfaces, which are presumably free from surface active contaminants, and developed a relationship in terms of the undulations on a jet exiting from a circular nozzle into a gas. Later workers (Addison and Elliott, 1950; Garner and Mina, 1959) used a contracting jet technique to measure the surface tension as a function of surface age for liquidliquid jets. Garner and Mina (1959), in a study of aqueous solutions of n-hexyl alcohol jetting into mutually saturated medicinal paran, showed in their Figure 9 that when the surface age approached 0.4 s for low concentrations of alcohol the dynamic interfacial tension decreased from near the water/paran value to a lower value as the surface active agent (alcohol here) adsorbed on the surface. Thus, instead of the lower interfacial tension values measured by Meister and Scheele and Bright, and listed in Table 6.2, which are probably due to slight impurities, we have used the pure values, since the surface ages of the breaking jets in this study are all less than about 0.1 s. However, to calculate the predicted jet lengths of the linear theory of Meister and Scheele (1969a), we have used their measured surface tension values. Again, use of the pure surface tension values would have resulted in a higher discrepancy between their predictions and the data. Further discussion of dynamic surface tension can be found in the work of Edwards et al. (1991).

129 6.4.1 Inviscid Free Liquid-Liquid Cylinder Breakup The problem of the capillary breakup of an initially stationary innitely long free inviscid liquid cylinder in another inviscid liquid with an outer freeslip wall (Christiansen and Hixson, 1957) was chosen as a validation test of the numerical method developed in the present work. This problem involves a transient ow, a non-constant surface curvature, and admits an analytical solution at least for small times obtained from a linear analysis as given in section 6.2.2. Figure 6.3 shows the complete breakup time sequence as obtained with the VOF numerical calculation for one axial wavelength using an initial cosine amplitude perturbation with a0 /R = 0.001 on the interface and the most dangerous wavenumber kR = 0.6773 as calculated from equation (6.9). A half wavelength was used for the actual computational domain with free-slip conditions at z z/R = 0 and z = 4.6384, and the distance from the centerline to the free-slip wall is L3 /R = 3; similar results were obtained for two wavelengths. Figure 6.4 shows the early time sequence for a half wavelength calculated using the linear theory of Christiansen and Hixson (1957) (CH) (equations (6.6) and (6.9)) compared to the present nonlinear simulation. The densities used were those of n-heptane in water given in Table 6.2, 1 = 2 = 0, and = 51.1 dyne/cm. The uniform mesh used for the simulation comprised 75 cells in the radial and 120 cells in the axial directions. This cell distribution was optimized through mesh sensitivity studies. The breakup time of 0.023 s predicted by the simulation compares well with that predicted by the linear theory of 0.024 s (when a = R in equation (6.6)). This result is consistent with the calculations of Mansour and Lundgren (1990), who also found good agreement with breakup times between their nonlinear model and Rayleighs (1879) linear theory, but, in addition, found signicant interface deviation from the linear theory near breakup time, leading to the formation of the satellite drop. The excellent agreement of the simulation results with the analytical theory validates the numerical code under conditions that share many factors with the actual physical problems studied in this work.

130

Figure 6.3: Inviscid liquid-liquid cylinder breakup. Heptane in water, t = 0.002 s.

131

CH This work 4

Interface at t = 0 s 3

z*

t = 0.014 s 2 t = 0.018 s t = 0.022 s 1 Free Slip Wall Axis of Symmetry

r* Figure 6.4: Inviscid liquid-liquid cylinder. Heptane in water with interface positions at four times before breakup. Linear theory of Christiansen and Hixson (1957) (CH) compared to the present nonlinear model.

6.4.2 Long-Time Free Surface Dynamics The main body of the results presented here are comparisons between the experiments of Meister and Scheele (Meister, 1966; Meister and Scheele, 1969a) and Bright (1985) and our numerical simulations. Figures 6.5, 6.6, and 6.7 represent free surface comparisons of nozzle 5, 3, and 6 respectively (see Table 6.1). The experimental interface data, traced from photographs of the interface, are located in the lower right corner frame of each gure (labeled with the experimenters name). Each frame of the gures is a contour of the interface

132 (F = 1/2) at equal time intervals, starting from startup at t = 0. Figure 6.5, nozzle 5, represents a short jet at high Re2 with only one undulation before breakup into drops. Our numerical simulation also shows one undulation before breakup, and about the same jet length before breakup as the data. Notice the cap-like drop formation and the general trend to oscillating ellipsoid-like drops in both the numerical and experimental data. There is some evidence of non-axisymmetric drops in the experimental frame on Figure 6.5; this is obviously not captured by the simulations, but appears to be a minor feature in the overall picture. Qualitatively the drop sizes predicted by our simulation can be seen to be about the same as the data, but the quantitative comparison of drop sizes is the subject of chapter 7, and is not discussed here. Our simulation also indicates a signicant jet interface contraction before breakup, also in agreement with the data. Figure 6.6, nozzle 3, represents a jet of intermediate length, obtained at intermediate Re2 with again only one undulation before breakup into drops. The numerical simulation also shows one undulation before breakup, and about the same jet length before breakup as the experimental data. The oscillating ellipsoidlike drops can also be seen in some of the simulation frames. The drop sizes predicted by the simulation can again be seen to be about the same as the data, with more shape variety in the simulation results. The simulation also indicates some jet interface contraction before breakup, in agreement with the data. Figure 6.7, nozzle 6, represents the dynamic behavior of a long jet, obtained at the lowest Re2 with, in general, three undulations observed before breakup. The numerical simulation shows about two undulations before breakup, and again about the same jet length before breakup as the data. Notice the pear-like drops formed, which can also be seen in the numerical simulations. The present simulation indicates very little jet interface contraction before breakup, in agreement with the data.

133

Figure 6.5: Free surface comparison. Experimental data of Meister (1966) compared to the present simulation. Heptane into water, nozzle 5, R = 0.344 cm, v = 15.48 cm/s.

134

Figure 6.6: Free surface comparison. Experimental data of Meister (1966) compared to the present simulation. Heptane into water, nozzle 3, R = 0.127 cm, v = 21.6 cm/s.

135

Figure 6.7: Free surface comparison. Experimental data of Bright (1985) compared to the present simulation. Decane into water, nozzle 6, R = 0.08 cm, v = 44 cm/s.

136 6.4.3 Short-Time Free Surface Shape Sequence Meister and Scheele (1969a) presented a short-time sequence of frames for the liquid jet obtained with nozzle 3 to illustrate that alternating larger and smaller drops could be produced from a liquid-liquid jet. Their tracings of the free surface taken from a high-speed movie (their Figure 6) are shown in Figure 6.8. For comparison, a jet case with the same conditions, corresponding to nozzle 3 and Figure 6.6, was simulated numerically and is illustrated in Figure 6.9. In Figure 6.8 the data indicate a pattern corresponding to a large drop breaking o, a small drop breaking o about ve frames later, followed by another large drop breaking o about 18 frames later. The numerical simulation indicates about the same pattern and interval between frames being produced. It turns out that in the simulation this pattern did not always recur, with several large drops being produced before the large-small-large pattern appeared again. It is interesting also that most of the simulation frames do not contain satellite drops which would be formed from the neck of the uid drawn out during drop break o. This appears to be consistent with the experiments of Meister and Scheele (1969a) who found that only in some instances were satellite drops formed. This is in contrast to single phase jets where satellite drops are frequently formed (Donnelly and Glaberson, 1966). 6.4.4 Jet Length Comparisons and Discussion The jet length, L/R, was calculated as a function of time for each simulation case, and a typical result is shown in Figure 6.10. As can be seen in that gure, the general pattern is a rise to a certain length, detachment of a drop, with a buildup in length to form the next drop. This results in a sawtooth pattern as seen in the gure, a result that was conrmed by Fourier analysis of the time series data. Superimposed on that, a longer-term trend appears in that the mean height between drops rises from zero at startup and levels o to a pseudo-steadystate (limit cycle) value. However, it can be seen that at long times the length is

137

Figure 6.8: Free surface sequence. Experimental data of Meister and Scheele (1969a) Heptane into water, nozzle 3, R = 0.127 cm, v = 21.6 cm/s, t = 0.00416 s.

138

Figure 6.9: Free surface sequence the present simulation. Heptane into water, nozzle 3, R = 0.127 cm, v = 21.6 cm/s, t = 0.00416 s.

139 not periodic, but rather chaotic. We have found that the variation in amplitude and period generally increases with increasing Reynolds number, Re2 , with the uctuations becoming more random. We thus conjecture that the aperiodicity is due to the increasing importance of the nonlinear terms in the momentum equations, causing a chaotic response. 30

20

L/R
10 0 0

1 t (s)

Figure 6.10: Jet length vs. time. Heptane into water, nozzle 3, R = 0.127 cm, v = 21.6 cm/s.

The average, minimum, and maximum lengths at the pseudo-steady-state predicted by the numerical simulation are shown in Figures 6.11 to 6.13 for all six cases. Also shown are the experimental length data of Meister and Scheele (1969a)

140 and Bright (1985) and the numerical results of Meister and Scheele (1969a) (MS) from equations (6.18) (6.25). Figure 6.11 shows the predicted lengths for nozzles 1, 3, and 5 as a function of Reynolds number, Re2 . Figure 6.12 shows the predicted lengths for nozzles 2 and 4 as a function of Reynolds number, Re2 . Figure 6.12 also shows the predicted lengths of the current simulation for nozzle 2 at two perturbation levels, a0 /R = 0.01 and 0.001. Figure 6.13 shows predicted length vs. experimental length for all six nozzles. As pointed out by Meister (1966), since the average measured height of the jet is increased by /2 due to the forming and detaching of the drops (see, e.g., Figure 6.10), a value of /2 was added to the MS model length predictions to be comparable to the photographic data. Several points regarding the origin and nature of the experimental data should be noted assessing the agreement of theory with experiment. One is

that Meister and Scheele (1969a) took four photographs of the jet and averaged them to obtain their experimental length (see Figure 1 of the reference). No information is available on the accuracy (or long-term reproducibility) of this approach, except the regularity and frequency of the data points themselves, but it can be conjectured by considering the numerically simulated length in Figure 6.10 that only four points of this curve taken at random times would not be a very representative sample. Only one photograph was available for the data point of Bright (1985) which was used in Figure 6.13. One can then imagine minimum and maximum bars on the experimental data about as large as those included in Figures 6.11 to 6.13 for the numerical simulations. Another aspect of the data inadequately accounted for in the simulations is the fact the magnitude and frequency of the perturbations is experimentally unknown. We have kept the amplitude constant for all cases for both the MS model (a0 /R = 0.0025) and the numerical simulations (a0 /R = 0.01). A better t of the data could be obtained by varying this parameter for each nozzle but little is to be gained from this approach in terms of improving the predictive capability of the simulations. As previously mentioned, the frequency of the perturbation was also

141

50 MS model This work MS Data, R = 0.04065 cm MS Data, R = 0.127 cm MS Data, R = 0.344 cm

40

30

L/R
20 10 0 500

1000 Re2

1500

2000

Figure 6.11: Jet length vs. Reynolds number data of Meister and Scheele (1969a) (heptane into water) and numerical results of Meister and Scheele (1969a) (MS) compared to the present simulation for nozzles 1, 3, and 5.

varied in the numerical simulations to evaluate the sensitivity of the results, which proved in this case to be small. The MS model used the maximum in the growth rate according to equation (6.10) to estimate the most dangerous wavelength. To eliminate this ambiguity a future experimental study could be undertaken with forced amplitude and frequency perturbations for liquid-liquid jets analogous to the experiments of Donnelly and Glaberson (1966) who investigated liquid jets in air. Recently we became aware of an experimental study by Tadrist et al. (1991),

142

50 MS model This work, ao/R=0.001 This work, ao/R=0.01 MS Data, R = 0.08 cm MS Data, R = 0.166

40

30

L/R
20 10 0 500

1000 Re2

1500

2000

Figure 6.12: Jet length vs. Reynolds number data of Meister and Scheele (1969a) (heptane into water) and numerical results of Meister and Scheele (1969a) (MS) compared to the present simulation for nozzles 2 and 4.

in an attempt to characterize the instability as completely as possible, measured uctuations of the interface of a liquid-liquid jet and found them to be characteristic of a weak-bad random signal with the power spectra densities of typical signals presenting a band of a given width around a characteristic frequency. Near the nozzle the jet amplitude uctuations were slight, but uctuations increased rapidly with distance from the nozzle. This study also found that the measured amplitude of the disturbance on the jet was (a0 /R 0.01), a value that we found

143

50

40

L/R Predicted

30

20

10

MS model, MS Data MS model, Bright Data This work, MS Data This work, Bright Data 0 10 20 30 L/R Experimental 40 50

Figure 6.13: Jet length predicted vs. experimental data of Meister and Scheele (1969a) (heptane into water) and Bright (1985) (decane into water) and numerical results of Meister and Scheele (1969a) (MS) compared to the present simulation for nozzles 1 6.

independently by our simulations that best t the data. As was mentioned previously in this section, a surface contaminant was probably present in the experiments as evidenced by the discrepancy between the pure and measured surface tension values listed in Table 6.2. Since the jet is generating a clean surface at the nozzle tip, the contaminant concentration will be low. As surface age increases, the contaminant will diuse to the surface, therefore resulting in an axial concentration gradient with a subsequent surface

144 tension gradient resulting. This can lead to surface forces along the axial direction resulting in a Marangoni ow (Edwards et al., 1991). However, in the present cases, the magnitude of this gradient is probably small since the jet breaks up rather quickly, and the measured value for the interfacial tension of n-heptane is only 21% lower than the pure value. Although the predictions from the MS model come close to the data at L/R 15, it can be seen from Figure 6.13 that their slope is much less than the 45 line (which corresponds to perfect agreement). On the other hand, while the results from the simulations show the same general trend of the 45 line, they also show a small bias to over-prediction of the length. This should be attributed to the presence of numerical viscosity (Hirt, 1968) which tends to make the numerical solution more stable than the physical one. For example, in numerical calculations where HN was set to unity (full donor-cell dierencing), the nozzle 5 case would not break in the region L/R 20. As was mentioned previously, HN was set to 1.2 times the minimum required for numerical stability to minimize this viscosity eect. As a smaller value for HN would have made the code numerically unstable, it appears that the only remedy is mesh renement that the present availability of computer resources has limited to about 65 255 cells. A stringent numerical stability criterion (Hirt and Nichols, 1981) is used to adjust the time step automatically to a small value (typically 1 105 s), and combined with the large number of mesh cells, this results in about 10 teraop of work for a typical calculation such as those reported in Figures 6.11 to 6.13. An exception to the above observation for overprediction of jet-length are the results obtained for nozzle 2, which are substantially under-predicted. A possible explanation is that this nozzle may correspond to an exceptionally low value of a0 /R experimentally. To test this we simulated the nozzle 2 case with a ten-fold lower value of a0 /R = 0.001 and found the results to be more consistent with the data as expected from linear stability analysis. As can be seen in Figure 6.12, the simulation approaches the data as the perturbation is lowered.

145 Keeping the preceding discussion in mind, we conclude that, in most cases, the present simulation agrees satisfactorily with the data, within experimental error. To be more quantitative, a more careful control of the noise present during the experiments appears to be necessary. The MS model has many assumptions, as mentioned in section 6.2.4. As can be seen by examining the jet length equation (6.18), two of the critical steps are the calculation of the most dangerous growth rate s and the velocity at which the disturbance is propagated, the interfacial velocity vI . As discussed in section 6.1, Kitamura et al. (1982) experimentally varied the relative motion of the continuous phase. They found that the jet length shortened as the absolute value of the continuous phase velocity relative to the jet increased from zero with condition of zero relative phase velocity giving the best experimental agreement with Tomotikas stationary jet solution translated at the average velocity of the jet. They concluded that the discrepancy between experiment and Tomotikas analysis in the nonzero relative velocity cases was due primarily to the relative motion of the continuous phase, and not to other factors such as velocity prole relaxation. In our case, the outer phase is relatively stagnant away from the interface, and this could result in a shearing type instability (Russo and Steen, 1989) that would result in an incorrect s, since the Tomotika equation (6.10) assumes that the two phases have the same axial velocity. To correct this, an analysis similar to that of Sterling and Sleicher (1975), Bright (1985), or Russo and Steen (1989) would have to be performed, but since gravity (buoyancy) and the outer phase viscosity would have to be included, the modications would be nontrivial and, if feasible, would result in a rather complex solution procedure. The value of the Tomotika analysis is that it is currently the only approach that incorporates in the outer phase the physical properties of density and viscosity. Another source of error in the MS model is the calculation of the axial interfacial velocity, vI . It is noteworthy that Kitamura et al. obtained good results at zero relative phase velocity, using the average velocity of the jet in their jet

146 length equation, the same equation used by Smith and Moss (1917) (6.17), with the maximum growth rate s calculated using Tomotikas equation (6.10). This can be explained by the fact that if the jet is long (their shortest heptane into water jet was L/R = 16), the velocity prole will have relaxed so that the interfacial velocity will be comparable to the average velocity. However, if the velocity prole has not had a chance to relax, this would not be true. Bright (1985) found experimentally that the surface wave velocity was about half the nozzle velocity for decane into water jets at v = 150 cm/s. Meister and Scheele (1969a) reported only two experimental point checks of liquid-liquid interfacial velocities predicted by equations (6.19) (6.27) as measured by observing disturbance node velocities on the surface of the jet. To test these ideas we compared the MS model calculation of jet radii a/R (equation (6.26)) with the current (equation (6.27)) and interfacial velocities vI /v simulation for nozzle 3; the results are shown in Table 6.3. The comparison is quite good until just before jet breakup, which is consistent with the experimental checks by Meister and Scheele (1969a). Therefore, we conclude that the error in the calculation of the interfacial velocities is not a major contributor to the overall discrepancy between the MS model and the data. Thus, given this fact, coupled with the experimental observations by Kitamura et al., and by examining the MS jet length equation (6.18), the most likely reasons for the MS disagreement are (1) the maximum growth rate s is assumed independent of the jet velocity v , and (2) the entire analysis is a local one, not taking into account the velocity prole relaxation in the axial direction away from the nozzle exit. Still, the MS analysis can be used as a rst estimate, its major drawback being its relative insensitivity towards Reynolds number changes. 6.5 Conclusions The robust and stable VOF-based numerical method developed in this dissertation has been used here successfully to simulate high Reynolds number, high buoyancy number, and low Weber number ows, up to and beyond the

147 Table 6.3: Jet radii a/R and interfacial velocity vI /v as a function of axial position z/R. Comparison between the MS model and this work at two nozzle velocities v for nozzle 3. v cm/s z/R a/R MS model Equation (6.27) 21.6 21.6 21.6 28.0 28.0 28.0 28.0 28.0 5 10 15 5 10 15 20 25 0.851 0.766 0.710 0.890 0.819 0.769 0.730 0.699 0.870 0.750 0.634 0.880 0.841 0.807 0.765 0.690 This work vI /v MS model Equation (6.26) 0.528 0.773 0.990 0.451 0.635 0.795 0.942 1.08 0.525 0.790 1.29 0.487 0.632 0.742 0.865 0.956 This work

jet breakup into drops. We have compared the predicted interface and drop shapes and average jet lengths with experimental data of Meister and Scheele (1969a) and Bright (1985). We conclude that, given the stochastic nature of the results, and their sensitivity to initial conditions, the present simulation agrees satisfactorily with the data, within experimental error. The use of this direct numerical simulation allowed us to ascertain the accuracy of the approximate numerical model of Meister and Scheele (1969a) and to examine the validity of the key underlying assumptions. The experimental system displays many aspects of an actual liquid-liquid contactor but it is especially dicult to simulate. There is also a non-deterministic character of the outcome, particularly at high Reynolds numbers, due to unknown disturbances as well as the nonlinearity of the governing equations. Additional issues include such aspects as impure materials, resulting in the possible need

148 to include dynamic surface tension, which has an inuence on the predicted jet lengths. Predominant among the issues that the present study revealed is the sensitivity to the external disturbances. An experimental study further characterizing these ambiguities could be undertaken with pure materials and forced amplitude and frequency perturbations for liquid-liquid jets, analogous to the experiments of Donnelly and Glaberson (1966), who investigated liquid jets in air. However, a broader implication for real liquidliquid contactors is that totally predictive simulations will remain dicult, even with the great improvements in computational power that continue to become available. The physical problem studied here is a challenging one because of the existence of moving complex free surfaces and the signicant roles of surface tension, inertia, and gravity, as well as experimental disturbances of unknown frequency and amplitude. It particularly places stringent requirements of stability and accuracy on our numerical technique in fact, any numerical technique aiming to handle these types of high Reynolds number free surface problems. As in all numerical computations, our current calculations are able to provide an approximate answer only. Much more extensive mesh renement and/or

use of adaptive (local) mesh renement is needed for more accurate answers. This additional computational eort is not currently justied, however, given the existing uncertainty in the magnitude of the experimental disturbances driving the jet breakup.

Vous aimerez peut-être aussi