International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International...

18
Numerical investigation of gravitational effects in horizontal annular liquid–gas flow Jeremy O. McCaslin , Olivier Desjardins Sibley School of Mechanical and Aerospace Engineering, Cornell University, Ithaca, NY 14853, USA article info Article history: Received 12 May 2014 Received in revised form 14 August 2014 Accepted 18 August 2014 Available online 27 August 2014 Keywords: Multiphase flow Liquid–gas flow Annular flow Stratified flow Direct numerical simulation Film sustainment abstract In this work, exploratory numerical simulations of liquid–gas flows in horizontal pipes are conducted for three different sets of conditions in the annular and stratified-annular flow regimes. Careful dimensional analysis is used to choose governing parameters in a way that yields flows that are relevant to realistic engineering applications, while remaining computationally tractable. Statistics of the velocity field and height of the liquid film are computed as a function of circumferential location in the pipe, demonstrating the existence of a viscous sublayer within the liquid film, as well as a viscous layer near the interface and a log law region within the gas core. The probability of dry-out conditions at the wall in upper regions of the pipe is shown to increase as gravitational effects increase. Circumferential motion of the liquid and gas phases within the pipe cross section are analyzed, informing possible mechanisms for sustainment of the liquid film. A simple model is developed that helps to characterize the dynamics of the liquid annu- lus and aids in understanding the effect of secondary gas flow on the circumferential motion of the film. Void fraction, film height, and film asymmetry are compared with experimental correlations available in the literature. Ó 2014 Elsevier Ltd. All rights reserved. Introduction The concurrent flow of liquid and gas inside circular pipes occurs in a vast range of engineering devices, such as chemical and nuclear reactors, pipelines, oil wells, and the receiver tubes of direct steam generation, among others. Depending on the rele- vant governing dimensionless parameters, the distribution of the phase interface can take many different forms. The way in which the phases are distributed will significantly impact hydrodynamic and thermal properties of these flows, which will in turn affect the optimized operation conditions and economical design of such sys- tems. In horizontal systems, gravitational effects lead to stratifica- tion, i.e., the buoyancy of the gas phase causes it to migrate to upper regions of the domain. This tendency can be diminished, depending on the relative importance of inertial and gravitational effects, as well as the gas volume fraction (also known as the void fraction). Within horizontal circular pipes, common flow regimes include bubbly flow, plug flow, slug flow, stratified and stratified- wavy flow, and annular and disperse–annular flow (Carey, 2008). Details regarding global flow characteristics that lead to differ- ent multiphase flow regimes are provided by Carey (2008) and are briefly conveyed here. At small void fractions, bubbly flow is char- acteristically observed. Small bubbles coalesce as the void fraction increases, forming larger bubbles. Larger bubbles intermittently flow in the upper region of the pipe, characteristic of plug flow. If the void fraction is high and inertial effects are small, the gas and liquid fully separate and stratified flow occurs. Kelvin–Helm- holtz type instabilities can occur in the stratified regime if the iner- tia of the gas is high enough, causing the interface between the phases to become wavy. The waves may reach the top of the pipe if their amplitude becomes large enough, taking a ‘‘slug-like’’ appearance, thus referred to as slug flow. When the void fraction is large and inertial effects dominate in the gas phase, an annular flow can occur in which the liquid assumes the form of a thin annular film around the interior surface of the pipe, while a gas core flows through the center. Buoyancy effects cause the film to be thinner near the top of the pipe than the bottom, leading to interfacial corrugations that vary in the circumferential direction due to waves that protrude further into the gas core near lower regions of the pipe. Shear caused by the high-inertia gas also has a tendency to entrain liquid droplets into the gas core. The afore- mentioned flow regimes for horizontal gas–liquid flows can be dif- ficult to distinguish from one another, as transitional forms of these regimes are common. In the past half-century, much work has gone into developing flow pattern maps in order to predict http://dx.doi.org/10.1016/j.ijmultiphaseflow.2014.08.006 0301-9322/Ó 2014 Elsevier Ltd. All rights reserved. Corresponding author. E-mail address: [email protected] (J.O. McCaslin). International Journal of Multiphase Flow 67 (2014) 88–105 Contents lists available at ScienceDirect International Journal of Multiphase Flow journal homepage: www.elsevier.com/locate/ijmulflow

Transcript of International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International...

Page 1: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

International Journal of Multiphase Flow 67 (2014) 88–105

Contents lists available at ScienceDirect

International Journal of Multiphase Flow

journal homepage: www.elsevier .com/locate / i jmulflow

Numerical investigation of gravitational effects in horizontal annularliquid–gas flow

http://dx.doi.org/10.1016/j.ijmultiphaseflow.2014.08.0060301-9322/� 2014 Elsevier Ltd. All rights reserved.

⇑ Corresponding author.E-mail address: [email protected] (J.O. McCaslin).

Jeremy O. McCaslin ⇑, Olivier DesjardinsSibley School of Mechanical and Aerospace Engineering, Cornell University, Ithaca, NY 14853, USA

a r t i c l e i n f o

Article history:Received 12 May 2014Received in revised form 14 August 2014Accepted 18 August 2014Available online 27 August 2014

Keywords:Multiphase flowLiquid–gas flowAnnular flowStratified flowDirect numerical simulationFilm sustainment

a b s t r a c t

In this work, exploratory numerical simulations of liquid–gas flows in horizontal pipes are conducted forthree different sets of conditions in the annular and stratified-annular flow regimes. Careful dimensionalanalysis is used to choose governing parameters in a way that yields flows that are relevant to realisticengineering applications, while remaining computationally tractable. Statistics of the velocity field andheight of the liquid film are computed as a function of circumferential location in the pipe, demonstratingthe existence of a viscous sublayer within the liquid film, as well as a viscous layer near the interface anda log law region within the gas core. The probability of dry-out conditions at the wall in upper regions ofthe pipe is shown to increase as gravitational effects increase. Circumferential motion of the liquid andgas phases within the pipe cross section are analyzed, informing possible mechanisms for sustainmentof the liquid film. A simple model is developed that helps to characterize the dynamics of the liquid annu-lus and aids in understanding the effect of secondary gas flow on the circumferential motion of the film.Void fraction, film height, and film asymmetry are compared with experimental correlations available inthe literature.

� 2014 Elsevier Ltd. All rights reserved.

Introduction

The concurrent flow of liquid and gas inside circular pipesoccurs in a vast range of engineering devices, such as chemicaland nuclear reactors, pipelines, oil wells, and the receiver tubesof direct steam generation, among others. Depending on the rele-vant governing dimensionless parameters, the distribution of thephase interface can take many different forms. The way in whichthe phases are distributed will significantly impact hydrodynamicand thermal properties of these flows, which will in turn affect theoptimized operation conditions and economical design of such sys-tems. In horizontal systems, gravitational effects lead to stratifica-tion, i.e., the buoyancy of the gas phase causes it to migrate toupper regions of the domain. This tendency can be diminished,depending on the relative importance of inertial and gravitationaleffects, as well as the gas volume fraction (also known as the voidfraction). Within horizontal circular pipes, common flow regimesinclude bubbly flow, plug flow, slug flow, stratified and stratified-wavy flow, and annular and disperse–annular flow (Carey, 2008).

Details regarding global flow characteristics that lead to differ-ent multiphase flow regimes are provided by Carey (2008) and are

briefly conveyed here. At small void fractions, bubbly flow is char-acteristically observed. Small bubbles coalesce as the void fractionincreases, forming larger bubbles. Larger bubbles intermittentlyflow in the upper region of the pipe, characteristic of plug flow. Ifthe void fraction is high and inertial effects are small, the gasand liquid fully separate and stratified flow occurs. Kelvin–Helm-holtz type instabilities can occur in the stratified regime if the iner-tia of the gas is high enough, causing the interface between thephases to become wavy. The waves may reach the top of the pipeif their amplitude becomes large enough, taking a ‘‘slug-like’’appearance, thus referred to as slug flow. When the void fractionis large and inertial effects dominate in the gas phase, an annularflow can occur in which the liquid assumes the form of a thinannular film around the interior surface of the pipe, while a gascore flows through the center. Buoyancy effects cause the film tobe thinner near the top of the pipe than the bottom, leading tointerfacial corrugations that vary in the circumferential directiondue to waves that protrude further into the gas core near lowerregions of the pipe. Shear caused by the high-inertia gas also hasa tendency to entrain liquid droplets into the gas core. The afore-mentioned flow regimes for horizontal gas–liquid flows can be dif-ficult to distinguish from one another, as transitional forms ofthese regimes are common. In the past half-century, much workhas gone into developing flow pattern maps in order to predict

Page 2: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105 89

flow regimes based on important flow parameters (Alves, 1954;Baker, 1953; Ghajar et al., 2007; Hoogendoorn, 1959; Kosterin,1949; Krasiakova, 1957; Mandhane et al., 1974; Taitel andDukler, 1976). These maps aim to categorize the two-phase flowinto the discussed regimes based on characteristics of the phases,such as superficial velocities and Reynolds numbers. Some mapshave had more success than others, but it is difficult for any twoparameters to contain enough information about the two-phaseflow to determine the resulting interfacial distribution.

Annular flows are especially common under conditions relevantto a plethora of thermo-fluid transport systems, and have thereforereceived much attention from experiments. A common goal ofmany experiments has been to extract the liquid film thickness,measured through planar laser-induced fluorescence (Schubringet al., 2010a,b; Farias et al., 2011), conductance probes (Hagiwaraet al., 1989; Paras and Karabelas, 1991a), fast response X-raytomography (Hu et al., 2013), and other techniques (Laurinatet al., 1985; Hewitt et al., 1990; Hurlburt and Newell, 1996,2000; Shedd and Newell, 2004; Schubring and Shedd, 2008). Otherworkers have measured mean circumferential velocity throughphotochromic dye activation (Sutharshan et al., 1995). A range ofother quantities relevant for practical applications have also beenmeasured extensively, such as wall shear (Hagiwara et al., 1989;Schubring and Shedd, 2009a) and pressure drop (Shedd andNewell, 2004; Schubring and Shedd, 2008). Multiphase dynamicsthat pertain to film sustainment mechanisms have been measured,such as characteristics of droplet entrainment and deposition(Paras and Karabelas, 1991b; Azzopardi, 1999; Al-Sarkhi andHanratty, 2002; Rodriguez et al., 2004; Alekseenko et al., 2013),induced secondary gas circulation (Flores et al., 1995), and surfacewave characteristics (Jayanti et al., 1990a; Hurlburt and Newell,1996; Schubring and Shedd, 2008). Despite efforts to formulatetheories and models to predict flow regime characteristics andregime transitions (Jacowitz et al., 1964; Anderson and Russel,1970; Taitel and Dukler, 1976; Kadambi, 1982; Ooms et al.,1983; Fukano and Ousaka, 1989; Serdar Kaya et al., 2000; Oomsand Poesio, 2003; Adechy, 2004; Moreno Quibén and Thome,2007; Schubring et al., 2011; Al-Sarkhi et al., 2012; Öztürk et al.,2013), there is a lack of theoretical understanding that would allowfor a description of the interface distribution from first principles.

Numerical studies of multiphase flows in transport systems arevery limited in comparison with their experimental counterparts.The vast majority of relevant liquid–gas direct numerical simula-tions (DNS) have focused on bubbly flows (Esmaeeli andTryggvason, 1998, 1999; Bunner and Tryggvason, 1999, 2002a,b;Nagrath et al., 2005), while significantly less computations of otherregimes have been performed. Those that do exist often rely on asimplified models to account for relevant physical processes, suchas two- and multi-fluid model-based simulations of slug flow(Issa and Kempf, 2003; Bonizzi and Issa, 2003a) and arising mech-anisms for bubble entrainment (Bonizzi and Issa, 2003b). A numberof numerical studies have been conducted that isolate a particularphysical process that is relevant to a liquid–gas pipe flow, such ashydrodynamic counterbalancing of the buoyancy on the gas core(Ooms et al., 2007, 2012), DNS and large-eddy simulation (LES) ofa sheared interface in turbulence (Fulgosi et al., 2003; Rebouxet al., 2006), DNS of turbulent heat transfer across a sheared inter-face (Lakehal et al., 2003), LES and Reynolds-averaged Navier–Stokes (RANS) simulations of secondary flow effects on dropletdeposition and turbophoresis (Jayanti et al., 1990b; van’tWestende et al., 2007), two-dimensional simulation of waveentrainment in vertical pipes (Han and Gabriel, 2007), and simula-tion of liquid film formation through wave pumping (Fukano andInatomi, 2003).

These types of numerical studies are very useful for gaininginsight toward the important dynamics of liquid–gas pipe flows,

yet computational limitations have prevented the formulation ofa detailed, first principles-based study of a horizontal liquid–gasflow inside a pipe that combines a fully turbulent gas phase withthe effects of a deformable interface and non-unity density and vis-cosity ratios. The present work is an exploratory investigation thatseeks to combine all of these effects together, demonstrating thatsimulations of annular and stratified flows relevant to realisticengineering applications are computationally feasible if governingparameters are chosen carefully. The aim of the present study is tofurther understand the hydrodynamic conditions that delineatethe annular and stratified flow regimes, and to test if the presentnumerical approach is capable of capturing the transition betweenthese regimes.

The computational approach used in the current study is out-lined in ‘Computational approach’, including a discussion of thegoverning equations and the methodology for interface capture.The system configuration is provided in ‘System configuration’,describing the simulation domain and the flow forcing mechanism,as well as validating the mesh resolution and immersed boundarymethod. Results of the simulations are discussed in ‘Results’, pro-viding a wide range of computed statistics and comparison withexperimental data and correlations. Mechanisms for film sustain-ment are discussed in ‘Dynamics in the pipe cross section’, fol-lowed by a simple model for the liquid film provided in ‘A modelfor the liquid film’ to further aid in understanding the interfacialdynamics as they pertain to film replenishment.

Computational approach

Mathematical formulation

The two-phase annular flows in this study are described by thecontinuity and Navier–Stokes equations. Assuming incompressibil-ity of both phases, i.e., r � u ¼ 0, the continuity equation is written

DqDt¼ @q@tþ u � rq ¼ 0; ð1Þ

where u is the velocity field and q is the fluid density. The Navier–Stokes equations are written as

@qu@tþr � qu� uð Þ ¼ �rpþr � l ruþruT

� �� �þ qg þ f b; ð2Þ

where p is the pressure, g is the gravitational acceleration, and l isthe dynamic viscosity. The body force f b used for momentum forc-ing within the periodic computational domain will be discussed in‘Flow forcing’.

The material properties are taken to be constant within eachphase, and the subscripts l and g are used to describe the densityand the viscosity in the liquid and gas, respectively. These quanti-ties are discontinuous across the interface C, and it is convenientto introduce their jump as q½ �C ¼ ql � qg and l½ �C ¼ ll � lg . Thevelocity field is continuous across C in the absence of phasechange, i.e., u½ �C ¼ 0. This is valid for the isothermal conditionssimulated herein. The surface tension force will lead to a disconti-nuity in the normal stress at the gas–liquid interface, which trans-lates into a pressure jump that can be written as

p½ �C ¼ rjþ 2 l½ �CnT � ru � n; ð3Þ

where r is the surface tension coefficient, j is the curvature of theinterface, and n is the normal vector at the interface. The discontin-uous pressure is dealt with using the ghost fluid method (GFM)(Fedkiw et al., 1999), and information regarding numerical imple-mentation of these equations is given by Desjardins et al. (2008a,b).

Page 3: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

90 J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105

Interface capturing

In this work, the accurate conservative level set approach(ACLS) of Desjardins et al. (2008b) is used to capture the phaseinterface, along with a discontinuous Galerkin discretizationdescribed by Owkes and Desjardins (2013). Within the level setframework, the interface is implicitly defined as an iso-surface ofa smooth function w, and is transported by solving

@w@tþr � uwð Þ ¼ 0: ð4Þ

Since the velocity field is solenoidal, this equation implies that theinterface undergoes material transport, in accordance with Eq. (1).The level set function is defined as the hyperbolic tangent profile

w x; tð Þ ¼ 12

tanh/ x; tð Þ

2e

� �þ 1

� �; ð5Þ

where e determines the thickness of the profile. In ACLS (Desjardinset al., 2008b), e is set to half the computational cell size. From theprevious equation, / is a standard signed distance function, i.e.,

/ x; tð Þ ¼ �kx� xCk; ð6Þ

where xC corresponds to the closest point on the interface from x,and / changes signs when C is crossed. The level set w must bere-initialized to preserve the hyperbolic tangent profile, as transportand numerical diffusion will distort it (McCaslin and Desjardins,2014). Re-initialization is achieved by solving

@w@s¼ r � e rw � nð Þnð Þ � r � w 1� wð Þnð Þ; ð7Þ

where s is a pseudo-time variable. Sequential solution of Eqs. (4)and (7) combines the accuracy of level set methods with excellentconservation properties. The details of the ACLS methodology andits coupling with the GFM are described elsewhere (Desjardinset al., 2008b; Owkes and Desjardins, 2013; Desjardins et al., 2013).

Immersed boundary method

The immersed boundary (IB) method used for the enclosingcylindrical geometry in the present study is based on the work ofMeyer et al. (2010) and Desjardins et al. (2013) so that it providesdiscrete conservation of mass and momentum. To illustrate thebasic IB methodology, consider a general conservative transportequation for a quantity x, written as

@x@tþr � FðxÞ ¼ 0; ð8Þ

where FðxÞ represents the flux of x. Second order finite volumediscretization of Eq. (8) leads to

xnþ1c ¼ xn

c �DtV

XNf

f¼1

Af Fnþ1=2f � nf

� ; ð9Þ

where xnc is the cell-mean value of x at time tn;Dt is the size of the

time step, V is the cell volume, Nf is the number of cell faces, Af isthe area of the cell face, F f is the face-mean flux, and nf is the out-ward normal to the cell face. Because computational cells contain-ing the geometry are cut by the immersed boundary, Eq. (9)should be modified to contain only the portion of the cell that con-tains fluid, i.e., is ‘‘outside’’ the immersed boundary. Replacing thecell volume and area by their wetted fractions, Eq. (9) becomes

xnþ1c ¼ xn

c �Dt

avwV

XNf

f¼1

aswAf F

nþ1=2f � nf

� þ AIBFnþ1=2

IB � nIB

!; ð10Þ

where avw ¼ Vw=V is the cell wetted volume Vw divided by the cell

volume V, and asw ¼ Aw=Af is the cell face wetted area Aw divided

by the cell face area A. The last term accounts for the flux of x atthe IB, with AIB as the cell immersed area, F IB the mean flux alongthe IB, and nIB the outward normal to the IB. Calculating Vw;Aw,and AIB requires knowing the location of the immersed boundaryxIB, which is specified implicitly through the use of the signed dis-tance level set

BðxÞ ¼ �kx� xIBk: ð11Þ

The immersed boundary then corresponds to BðxÞ ¼ 0, and a changein the sign of BðxÞ distinguishes the domain inside the boundaryfrom the domain outside. Further details regarding this immersedboundary method can be found elsewhere (Desjardins et al.,2013; Meyer et al., 2010).

System configuration

Simulation domain

Despite the appeal of a cylindrical mesh for pipe flow simula-tions, cylindrical coordinates automatically increase orthoradialresolution near the pipe centerline and decrease it near the bound-ary. This is undesirable when simulating a relatively thin annularflow that contains important interfacial dynamics near the pipewall. Combined with the presence of liquid drops entrained inthe gas core, the present study lends itself to the use of a uniformmesh so that multiphase dynamics can be well resolved every-where. To that end, we perform the simulations on a uniformCartesian mesh and account for the enclosing pipe geometry withthe IB method described in the previous section.

Flow forcing

Simulations of fully developed two-phase pipe flow areachieved by prescribing periodic boundary conditions in the axialdirection, denoted by x. Momentum lost at the pipe walls is rein-troduced in the form of the constant source term f b ¼ f b bex in Eq.(2), where bex is the axial unit vector. The value of f b is chosen tomaintain a given bulk Reynolds number of a corresponding sin-gle-phase flow with the same pipe diameter D and fluid propertiesas the gas phase, written as Re ¼ UD=m, where U is the bulk axialvelocity and m is the kinematic viscosity. Once Re is chosen, Pra-ndtl’s friction law for smooth pipes (Pope, 2000) is used to obtainthe friction factor f. The relation f b ¼ qU2f=ð2DÞ is then used toobtain the value of f b used to force the two-phase flow. For a thinannular flow that is dynamically dominated by a high inertia gascore, the superficial gas Reynolds number will be close in valueto the single-phase Reynolds number that f b would yield.

Mesh resolution

The computational domain is of size 5D� D� D and is com-posed of Nx � Ny � Nz ¼ 1280� 256� 256 uniform grid cells. Toensure sufficient resolution for the two-phase simulations, we val-idate that the mesh, together with the IB method, is capable ofresolving a single-phase turbulent flow that is forced by the samef b used in the two-phase simulations. Fig. 1 shows a single-phaseturbulent flow with Re ¼ 5310 on the same uniform mesh usedin the two-phase simulations. Results compare favorably withthe well-established DNS results of Fukagata and Kasagi (2002).Standard plus units uþ and yþ are used for the mean profile onthe left. The rms velocity urms is shown on the right as a functionof yþ, and the x; r, and h components agree very well with the cylin-drical DNS of Fukagata and Kasagi (2002).

The smallest interfacial length scales that arise are limited bysurface tension, which is a controlled parameter in the simulations.We set r to be as large as possible to allow for droplet entrainment

Page 4: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

Fig. 1. Validation of the Cartesian mesh resolution with the immersed boundary method for a single-phase flow with Re ¼ 5310. NGA (Desjardins et al., 2008a) (lines), DNS ofFukagata and Kasagi (2002) on a cylindrical mesh (symbols).

Table 1Frsp values for the three cases. Each case is runwith Resp ¼ 5000;Wesp ¼ 2000; f ¼ 0:037;ql=qg ¼ 16;ll=lg ¼ 4:62, and eg ¼ 0:85. Gasphase properties correspond to air.

Case Frsp

A 1B 6.56C 1.64

J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105 91

into the gas core, but small enough to satisfy the condition that themesh-based Weber number WeDx ¼ qgu2

relDx=rK 5, where Dx isthe uniform mesh size. The relative velocity urel is estimated asthe difference between the gas and liquid superficial velocities.This criterion corresponds to resolving vibrational breakup of aliquid droplet on at least two computational cells and should guar-antee that most interfacial length scales are captured by the mesh.

Cases considered

The gaseous core in a horizontal annular flow must carry a sig-nificant amount of momentum in order to sustain a liquid film atthe top of the pipe, since the liquid is much more dense. Thus, itsvelocity must be quite high. Indeed, very high gas Reynolds num-bers are required to achieve an annular flow under realisticair–water type conditions, and such conditions are not computa-tionally tractable with our approach. However, it is possible toobtain an annular flow in a simulation by modifying the densityratio, gas velocity, or gravitational acceleration. We do this care-fully by defining the gas Froude number as

Frg ¼

ffiffiffiffiffiffiffiffiffiffiffiqgj2

g

qlgD

s; ð12Þ

where jg is the superficial gas velocity (described further below) andg is gravitational acceleration. The Froude number as defined in Eq.(12) is a measure of the ratio of aerodynamic forces from the gasphase to gravitational effects on the liquid phase. Simulations allowus to verify the idea that, assuming a large void fraction eg ¼ Vg=V ,which is the ratio of volume occupied by the gas phase Vg to thetotal volume V, a value of Frg � 1 governs ‘‘annularity’’. Values ofFrg > 1 lead to the presence of a contiguous film, while Frg K 1means that the film is not contiguous around the pipe wall. Theinterfacial distribution converges to a stratified flow as Frg ! 0.

Once the source term f b is specified as described in ‘Flow forc-ing’, the other inputs to the simulations are the density and viscos-ity ratios, gravity, surface tension, and void fraction. Note that thewetted fraction el ¼ 1� eg is known when eg is known.

Since we do not know a priori what the value of jg will be beforethe flow becomes statistically stationary, it is convenient to definethe case parameters in terms of dimensionless variables basedon the corresponding single-phase bulk velocity U:

Reynolds number : Resp ¼qgUDlg

; ð13Þ

Weber number : Wesp ¼qgU2D

r; ð14Þ

Froude number : Frsp ¼

ffiffiffiffiffiffiffiffiffiffiffiqgU2

qlgD

s; ð15Þ

Friction factor : f ¼ 2Dfb

qgU2 : ð16Þ

In the present work we conduct three simulations with differentvalues of Frsp to study the balance between inertia and gravityand test the capability of the computational approach to reproducedifferent flow regimes. Each case is simulated with Resp ¼ 5000;Wesp ¼ 2000; f ¼ 0:037; ql=qg ¼ 16; ll=lg ¼ 4:62, and eg ¼ 0:85.The value of Frsp for each case is provided in Table 1. The simulateddensity and viscosity ratios correspond to high pressure conditionsinside the receiver tubes of direct steam generation loops, which isone application of interest for the present study.

Results

Flow characterization

Assuming an isothermal annular flow that is dynamically dom-inated by the high-inertia gas core, we use the following 3 dimen-sionless groups, in addition to Frg ;ql=qg ;ll=lg , and eg , tocharacterize the flow:

gas Reynolds number : Reg ¼qgjgDlg

; ð17Þ

gas Weber number : Weg ¼qgj2

g Dr

; ð18Þ

flow quality : x ¼_mg

_mg þ _ml; ð19Þ

where _ml;g are the mass flow rates of the liquid and gas. The super-ficial gas velocity is defined as jg ¼ egug , where the bulk gas velocity

ug ¼R

V agudVRV ag dV

¼ 1Vg

ZVg

udV ð20Þ

Page 5: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

Table 2Dimensionless groups for the three cases.

Case x Reg Weg Frg

A 0.66 3:37� 103 9:06� 102 1B 0.70 3:48� 103 9:67� 102 4.56

C 0.65 3:65� 103 1:07� 103 1.20

92 J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105

is the axial velocity u averaged over the volume occupied by gas,and ag is the local void fraction. Due to the presence of the liquidfilm, jg is not equal to U, the bulk velocity in the single-phase case.

The flow quality, gas Reynolds number, gas Weber numer, andgas Froude number that the flow converges to depend on the bal-ance between inertia and gravity and the resulting spatial distribu-tion of the phase interface. Fig. 2 shows the convergence history ofReg ;Weg ; Frg , and x for the three simulations, and their convergedvalues are listed in Table 2.

The nature of gravitational effects on the flow is clearly visiblethrough inspection of Fig. 3. A large number of disperse drops existwithin the gas core for both cases A and B, as is visible in Fig. 3(g)and (h). Drops are entrained and persist for a long time within thegas core before they are deposited on the liquid film, as gravita-tional effects are relatively limited compared to inertial effectsfor these cases. For case C, however, gravitational effects decreasethe lifespan of drops within the gas core, returning them to thebase film in lower regions of the pipe. This observation is in qual-itative agreement with the experimental findings of Simmons andHanratty (2001), who found that stratification of drops in a hori-zontal annular pipe diminished as gas velocity was increased, i.e.,as inertia became more significant relative to gravitational effects.It is also evident in Fig. 3(a)–(f) that the circumferential bias of theliquid film thickness increases with decreasing Frg .

A first step to compare the simulations with experiments ismade by plotting the simulations on the flow regime maps ofGhajar et al. (2007) and Taitel and Dukler (1976), as shown inFig. 4(a) and (b), respectively. The regime map of Ghajar et al.(2007) shown in Fig. 4(a) is based on superficial phase Reynoldsnumbers and does not explicitly incorporate gravitational effects.The map was derived from air–water experiments (Ghajar et al.,

Fig. 2. Convergence history of Reg ;Weg ; Frg , and x. Simulation time t is normalized by T flo

line), Frg ¼ 4:56 (dashed line), Frg ¼ 1:20 (dash-dotted line). The thin dashed lines show

2007) and therefore assumes a gravitational accelerationg0 ¼ 9:81 m=s2. An annular flow requires a very large superficialgas velocity to yield a large Froude number based on g0. Conse-quently, Reg for cases B and C is not large enough to fall withinthe annular regime of the Ghajar et al. (2007) map, as seen inFig. 4(a).

The Taitel and Dukler (1976) map, however, does explicitlyaccount for the effect of gravity. It is based on the Martinelli FactorX (Martinelli and Nelson, 1948) and the transition parameter F,defined as

X ¼ ðdP=dxÞlðdP=dxÞg

!1=2

ð21Þ

and

F ¼qgj2

g

ðql � qgÞDg

!1=2

; ð22Þ

where ðdP=dxÞl;g are the frictional pressure gradients caused if theliquid or gas were flowing alone in the pipe. The values ofðdP=dxÞl;g are computed according to the formulation of Carey

w, which is the flow through time of the gas core for the Frg ¼ 1 case. Frg ¼ 1 (solidthe mean value computed over the assumed statistically stationary period.

Page 6: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

Fig. 3. Instantaneous results. In (a)–(f), the interface is shown by the white line, and the grayscale indicates normalized axial velocity, ranging from 0 (black) to 1.7 (white).The phase interface is shown in (g)–(i).

J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105 93

(2008) (see page 484 of Carey (2008)). Although both cases B and Cfall within the disperse-annular regime of the Taitel and Dukler(1976) map in Fig. 4(b), the increased stratification of the film forcase C is quantified by the fact that F is closer to the stratified-wavyregime for case C than for case B.

Liquid volume fraction statistics

To display radial profiles of the results, it is convenient to intro-duce the orthoradial angle h, as defined in Fig. 5. For cases B and Cin which g – 0, data is averaged about the plane of symmetryshown by the dashed line in Fig. 5. Fig. 6 compares the temporallyand axially averaged liquid volume fraction al ¼ 1� ag for cases Band C, along with the mean location of the interface hðhÞ. While

only slightly thicker near h ¼ 0 for case B, the liquid film becomessignificantly thicker for 0 6 h 6 p=4 in case C. This is also displayedin the profiles of alðy; hÞ in Fig. 7, where y ¼ R� r for a pipe ofradius R with radial coordinate r. Although orthoradially symmet-ric, alðy; hÞ is shown for case A without averaging in h for reference.

In industrial applications, much can be inferred about the two-phase dynamics of the pipe flow based on the void fraction. As aconsequence, much effort has been put into developing experi-mental correlations and models for the void fraction(Butterworth, 1975; Harms et al., 2003; Serdar Kaya et al., 2000;Tandon et al., 1985) based on parameters such as quality and phasesuperficial velocities. A well-known correlation based solely onflow quality x and fluid densities is provided by Chisholm (1973),written as

Page 7: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

Fig. 4. Cases B and C on flow pattern maps found in the literature. Case B (�), case C (j).

θ = 0

θ = π

Fig. 5. Definition of orthoradial angle h for flow statistics.

94 J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105

ecorrg;1 ¼ 1þ 1� x 1� ql

qg

! !1=2qgð1� xÞ

qlx

24 35�1

; ð23Þ

where the superscript ‘‘corr’’ denotes a correlated value. Table 3shows the error between the prescribed void fraction eg ¼ 0:85compared to the correlated value for cases A–C, computed as

error ¼ecorr

g;1 � eg

��� ���eg

; ð24Þ

and good agreement is observed.A comprehensive assessment of 68 different experimental void

fraction correlations proposed in the literature is provided byWoldesemayat and Ghajar (2007). Including the term proposedby Woldesemayat and Ghajar (2007) that accounts for pipe

Fig. 6. Mean al for cases B and C, varying from 0 (white) to 1 (gray).

inclination angle b, the correlation of Coddington and Macian(2002) becomes

ecorrg;2 ¼ jg jg 1þ jl

jg

!ðqg=qlÞ0:10@ 1Aþ2:9

gDrð1þcosbÞðql�qgÞq2

l

� �0:2524 35�1

: ð25Þ

Table 4 shows the correlation error for each case, computed inthe same way as Eq. (24). The correlation agrees reasonably wellwith the simulations, considering that 85.6% of the 2845 datapoints tested by Woldesemayat and Ghajar (2007) fall within15% of the correlated value.

Velocity statistics

For case A in which the liquid film is not a function of h, it is ofinterest to directly compare the single- and two-phase turbulentvelocity profiles. Fig. 8(a) shows both profiles along with their cor-responding viscous sublayers, which follow uþ ¼ yþ, and logarith-mic regions, which follow

uþ ¼ 1j

ln yþ þ c; ð26Þ

where j is the von Kármán constant and c is a constant. The single-phase viscous sublayer roughly corresponds to the region yþ < 5,while the logarithmic region corresponds to yþ > 30. For the sin-gle-phase profile, Eq. (26) is plotted with j ¼ 0:35 and c ¼ 6 inFig. 8(a). Since the phase properties change across the interface,

The mean location of the interface hðhÞ is given by the black line.

Page 8: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

Fig. 7. Mean al as a function of y=R for different values of h. h=p ¼ 0 (thick solid line), h=p ¼ 1=4 (dashed line), h=p ¼ 1=2 (dash-dotted line), h=p ¼ 3=4 (dotted line), h=p ¼ 1(thin solid line).

Table 3Chisholm (1973) void fraction correlation, Eq. (23).

Case A Case B Case C

Error 6.34% 7.85% 5.95%

Table 4Woldesemayat and Ghajar (2007) void fraction correla-tion, Eq. (25).

Case A Case B Case C

Error 7.31% 10.51% 17.32%

J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105 95

there is no single value of m that can be used to perform the viscousnormalization for the two-phase profile. For comparison, us corre-sponding to the single-phase simulation is used. In order to plotthe liquid viscous sublayer uþl ¼ yþl in the same figure, the equation

uþ ¼ yþus;l

usð27Þ

is used, where us;l ¼ffiffiffiffiffiffiffiffiffiffiffisl=ql

pis the liquid friction velocity and sl is

the wall shear stress in the film. Eq. (26) within the gas core ofthe two-phase flow is plotted using j ¼ 0:29 and c ¼ �3:2. Thelocation of the interface hþ ¼ hus=m is also shown. There is clearlya viscous sublayer within the liquid film, as one would expectwithin a viscously dominated near-wall region. Interestingly, a log-arithmic region is observed in Fig. 8(a) within the gas core despitethe presence of disperse liquid droplets. The axial, radial, and ortho-radial rms velocities appear similar to the single-phase case withmaximum values occurring on the gas side of the interface, asshown in Fig. 8(b). The decrease in peak rms relative to the

single-phase flow is consistent with the fact that us;C=us ¼ 0:715,where us;C ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffisC=qg

qis the friction velocity at the interface and

sC is the mean interfacial shear stress. In a DNS of a sheared air–water interface, Fulgosi et al. (2003) attribute decreased interfacialfriction (relative to a solid wall) to transfer of energy from the flowinto form drag as the interface deforms. This idea is consistent withstudies that have shown reduced skin friction in turbulent channelswhen injecting bubbles near walls (Lu et al., 2005; Murai et al.,2007) or by using deformable walls (Kang and Choi, 2000).

We observe that the mean flow of the gas core is qualitativelysimilar to a single-phase turbulent pipe flow subject to a slip-wallboundary condition. This idea is in agreement with previousnumerical results of turbulence near an interface for simplifiedconfigurations (Lombardi et al., 1996; Solbakken and Andersson,2005). Based on such findings, the modified law of the wall forthe gas core becomes

u� uC

us;C¼ ðy� hÞus;C

mð28Þ

within the ‘‘viscous sublayer’’, where uC is the mean axial velocity atthe interface. We write Eq. (28) compactly as ðu� uCÞþC ¼ ðy� hÞþC,and the gas core log law is written as

ðu� uCÞþC ¼ 1j

ln ðy� hÞþCh i

þ c0; ð29Þ

where the value of the von Kármán constant j ¼ 0:35 for the single-phase log law is recovered, and c0 is a constant found to be equal to�2.5. This result is seen in Fig. 9(a) compared with the single-phaseprofile, plotted as ðu� uCÞþ ¼ f ððy� hÞþÞ for the sake of compari-son. In order to plot the interfacial viscous sublayer in terms ofquantities normalized by us, the lower dashed line in Fig. 9(a) isgiven by the equation

Page 9: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

Fig. 8. Mean and rms velocities for both case A and the corresponding single-phase flow.

96 J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105

ðu� uCÞþ ¼us;C

us

� �2

ðy� hÞþ: ð30Þ

The gas core profile follows this narrow interfacial viscous regionfor ðy� hÞþ K 2, and the gas core log law appears to be valid inthe region ðy� hÞþ > 30, which corresponds well with the outerportion of the single-phase buffer region. When plotting the shiftedaxial velocity rms within the gas core and normalizing by us;C as inFig. 9(b), we find that the peak value is roughly equal to the single-phase value normalized by us. It also occurs at roughly the samedistance from the interface as does the peak single-phase rms fromthe wall. Further analysis of annular flows in the formðu� uCÞþC ¼ f ððy� hÞþCÞ could lead to future modeling efforts builton ideas of existing law of the wall models.

Using liquid properties to normalize by us;l and ml;uþl ðyþl Þ is plot-ted in Fig. 10, along with the interface location hþl ¼ 8:83. The pro-file deviates from uþl ¼ yþl at a value of yþl � 5 on the liquid side ofthe interface, indicating that the interface lies outside the near-wall liquid viscous sublayer. This suggests that the outermostregion of the liquid film is not completely dominated by viscousaffects, which is in agreement with the presence of waves gener-ated through interfacial shear from the turbulent gas core, as is

Fig. 9. Comparison between the mean and rms velocities for case A gas core and the sin(lower solid line), single-phase viscous sublayer (upper dashed line), gas core interfacialgas core log law (lower dash-dotted line). In (b): single-phase profile (solid line), shifted

evident in Fig. 3(g). This deviation from laminar behavior in theoutermost region of the film will become relevant to the modelintroduced in ‘A model for the liquid film’.

Profiles of uþðyþÞ at different values of h are shown in Fig. 11 forcases B and C. Case A is also shown for reference, averaged onlyabout the h ¼ 0 plane. Similar to case A, viscous sublayers and log-arithmic regions are observed for cases B and C. The extent of theregions depends on the h location, since the film thickness varieswith h. The difference in uþ from case A is most significant for caseC, as the film is much thicker at h=p ¼ 0 than at h=p ¼ 1.

For h=p ¼ 0 and 1=4, it is difficult to tell from Fig. 11(b) if a log-arithmic region exists in the gas core. In Fig. 3(i), it is evident thatwaves protrude far into the gas core due to the increased filmthickness near the bottom of the pipe. This protrusion of waves,together with the wall-normal extent of the base film, may preventthe possibility of a logarithmic region for lowers values of h. A sim-ilar argument could be made at h ¼ 0 for case B, as shown inFig. 11(a). Modulation of turbulence by a rough interface andslowly moving drops after creation from the film has been shownpreviously (Azzopardi, 1999), thus it is reasonable to speculate thatinterfacial motions protruding into the gas core affect the logarith-mic region. This is evidenced by the axial velocity rms for cases B

gle-phase flow. In (a): single-phase profile (upper solid line), shifted gas core profileviscous sublayer (lower dashed line), single-phase log law (upper dash-dotted line),

gas core profile (dashed line).

Page 10: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

Fig. 10. Velocity profile within the liquid film. uþl ðyþl Þ (solid line), uþl ¼ yþl (dashedline), hþl (dash-dotted line).

J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105 97

and C shown in Fig. 12, which is largest near the interface for lowvalues of h.

Orthoradial symmetry in case A leads to a clean comparisonbetween radial gas core statistics and the single-phase case. How-ever, the presence of the liquid film imposes a new length scale onthe physical system, making the analysis more complicated thanthe single-phase case. The complexity is further increased for casesB and C, due to mean interfacial shear and film height that are afunction of h. An interesting question is whether or not the asym-metric film effects the radial gas core velocity profile in a qualita-tively similar way, regardless of h. This does seem to be the case,observed by normalizing the shifted gas core profiles by their localhðhÞ. The result is shown in Fig. 13, where ðu� uCÞ=us is plotted as a

Fig. 11. Mean uþðyþÞ for different values of h. h=p ¼ 0 (thick solid line), h=p ¼ 1=4 (dasline). The dashed lines give the location of hþ for each value of h.

function of ðy� hÞ=h. The single-phase log law with j ¼ 0:35 andc ¼ 6 is shown for comparison. The observed collapse onto a singleprofile for case B in Fig. 13(a) indicates that the local film height isa good measure of the characteristic viscous length scale for thatparticular case. Other possible length scales that do not lead to col-lapse are ml=us;l and mg=us;C. The gas core collapse of case C onlyseems to work for the three largest h values in which h does notvary too much. This is reasonable, as the heavily stratified film islikely not dominated by viscous effects near h ¼ 0. In any case,the idea of collapsing gas core profiles at different h onto a singleprofile, assuming hðhÞ does not vary too rapidly, could prove tobe useful in future modeling efforts of horizontal annular flows.

Film height and dry-out statistics

The probability of encountering dry-out conditions, i.e., the pipewall being non-wetted, depends on the average thickness of theliquid film at a particular orthoradial angle. Note that our compu-tational approach does not properly account for the physics at thetriple point if the wall becomes non-wetted. The Neumann bound-ary condition on our level set enforces a 90� contact angle with asolid wall, resulting in a simplified model for contact line dynamicsand contact line migration. Although we do not account for properphysics once the wall becomes non-wetted, the present resultsshould provide insight on the onset of dry-out.

As seen in Fig. 14, both case A and case B remain sufficientlywetted for all h and thus have a zero probability of dry-out. Onlysymmetry with respect to the h ¼ 0 plane is exploited for case Afor the sake of comparison. The film height hðhÞ is nearly flat forcase A and would eventually be perfectly flat if averaged over a

hed line), h=p ¼ 1=2 (dash-dotted line), h=p ¼ 3=4 (dotted line), h=p ¼ 1 (thin solid

Page 11: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

Fig. 12. Axial velocity rms urms normalized by the bulk axial velocity in case A, ubulk. Values vary from 0 (gray) to maxðurms=ubulkÞ (white), given by the legend. The meanlocation of the interface hðhÞ is given by the black line.

Fig. 13. Collapse of multiple ðu� uCÞþððy� hÞþÞ onto ðu� uCÞþððy� hÞ=hÞ. h=p ¼ 0 (thick solid line), h=p ¼ 1=4 (dashed line), h=p ¼ 1=2 (dash-dotted line), h=p ¼ 3=4 (dottedline), h=p ¼ 1 (thin solid line).

98 J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105

long enough period. For case B, h is larger near h ¼ 0 than h ¼ p, asexpected. The slight local maximum in hðhÞ near h=p � 3=4 couldbe due to insufficient averaging. For case C, the film thickness issignificantly biased toward h ¼ 0 and becomes very thin at thetop of the pipe. As a result, Pdry varies between 0 and nearly 8%for values of h=p > 0:4.

Film thickness calculations from the present simulations arecompared to experiments in the literature. Schubring and Shedd(2009b) provide two correlations for the average film thickness �h.In our simulations, we define �h as

�h ¼ 1p

Z p

0hðhÞdh: ð31Þ

In their experiments, Schubring and Shedd (2009b) only measuredthe film thickness at h=p ¼ 0, 1/2, and 1, so Eq. (31) becomes

�h ¼ 14

h0 þ 2hp=2 þ hp� �

; ð32Þ

where the subscript indicates the value of h corresponding to h. Asimple correlation for �h provided by Schubring and Shedd (2009b) is

�hcorr

D¼ 12:5 Re�2=3

g : ð33Þ

They measured the correlation error

error ¼�hcorr � �h�� ��

�hð34Þ

for 206 annular data points and reported an average value of 11%.The correlation does not fit the present simulations, as shown inTable 5. The discrepancy is possibly due to high Reynolds andWeber number effects, as most of the experiments were conductedover a range of Reg and Weg more than an order of magnitude largerthan the simulations. Increased inertia could lead to different filmcharacteristics that are not accounted for in this correlation, as wellas a higher percentage of liquid taking the form of drops. This wouldreduce the film height, which is consistent with Eq. (33) underpre-dicting the simulation mean film height for all three cases.

The previous correlation is based entirely on the gas phase. Inorder to account for the presence of the liquid film, Schubringand Shedd (2009b) also proposed

�hcorr

D¼ 4:7

1x

qg

ql

� �1=3

Re�2=3g ; ð35Þ

where

ReG ¼GDll

ð36Þ

is a Reynolds number based on the mass velocity and liquid viscos-ity. We find that the inclusion of liquid film processes causes thecorrelation to fit better with the current simulations than the previ-ous one, as seen in Table 6. Note that the improvement is onlyobserved for cases A and B, which is reasonable, since they fall withinthe annular regime that Eq. (35) is based on. The experimental

Page 12: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

Fig. 14. Mean film thickness hðhÞ and dry-out probability PdryðhÞ for the three cases. hðhÞ (solid line), PdryðhÞ ().

Table 5Film height correlation error (computed from Eq. (34))when compared to Eq. (33) (Schubring and Shedd,2009b).

Case A Case B Case C

Error 48.77% 49.69% 26.18%

J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105 99

film heights from the data bank of Schubring and Shedd (2009b)have been normalized by pipe diameter in order to compare themwith the present simulations, as shown in Fig. 15. Good agreementrelative to the spread of the experimental data is observed.

In addition to the mean film thickness, Schubring and Shedd(2009b) also provide a correlation for the asymmetry of the liquidfilm, defined as

A ¼ h0

hp: ð37Þ

The correlation is

Acorr ¼ 1� e�0:63Frh� ��1

; ð38Þ

where

Frh ¼qgjg

ql g�h� �1=2 ð39Þ

Table 6Film height correlation error (computed from Eq. (34))when compared to Eq. (35) (Schubring and Shedd,2009b).

Case A Case B Case C

Error 21.16% 22.49% 32.95%

is a Froude number based on the superficial gas velocity and themean film thickness. Asymmetry for case A is unity, since Frg ¼ 1and h0 ¼ hp. Eq. (39) significantly overpredicts asymmetry for casesB and C, and the errors jAcorr �Aj=A are shown in Table 7. Eq. (38) isbased on annular experiments with g ¼ 9:81 m=s2 and may there-fore encompass physical processes not present in the simulations.Also, Eq. (38) is based on experiments with A < 3:5, and thus maynot incorporate physical processes for heavily stratified films.

A more detailed asymmetry correlation is provided by Hurlburtand Newell (2000), which accounts for asymmetry by defining

eA ¼ �hh0; ð40Þ

Fig. 15. Film height correlation of Schubring and Shedd (2009b), Eq. (35). Case A (),case B (r), case C (j), �25% (dashed lines). Open symbols denote different pipediameters from the experiments: 8.8 mm (), 15.1 mm (M), 26.3 mm (�).

Page 13: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

Fig. 16. Film asymmetry eA compared to experimental data. Eq. (41) with h0 as themean h0 of the experiments (solid line) and as the mean of the simulations (dashedline). Case A (), case B (r), case C (j), Dallman (1978) (�), Fukano and Ousaka(1989) (), Hurlburt and Newell (1996) (M), Jayanti et al. (1990a) (�), Laurinat and

100 J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105

where �h is defined by Eq. (31). Fig. 16 shows eA as a function ofð _mg= _mlÞ1=2Frsg for all three cases, compared to a wide range ofexperimental data. The curve fit to the data in Fig. 16 is given as

eA ¼ 43p

h0

D

!1=2

þ 0:9 1� exp�ð _mg= _mlÞ1=2 Frsg

90

! !; ð41Þ

where h0 is a value of ho representative of the data, andFrsg ¼ jg=ðgDÞ1=2 is a Froude number based on the superficial gasvelocity and pipe diameter. Specifying h0 as the mean h0 of theexperimental data leads to the solid line in Fig. 16, while usingthe mean h0 of the simulations as h0 leads to the dashed line. Eitherway, eA from the simulations agrees well with Eq. (41) with respectto the spread of the experimental data. Note that the abscissa ofFig. 16 is1 for case A, and the value of eA ¼ 1 is shown at the rightof the figure (with an arrow pointing to the right) for the sake ofcomparison.

Hanratty (1982) (+), Paras and Karabelas (1991a) (}), Williams (1990) (O).

Table 8

Dynamics in the pipe cross section

In the absence of mechanisms for replenishing the liquid filmnear h=p ¼ 1, all of the liquid would drain due to gravity and theflow would become stratified. Much debate has centered aroundthe mechanisms for film sustainment, possible explanations beingthat drops are entrained near lower values of h and deposited nearhigher values (Russell and Lamb, 1965), the action of surface wavesmoves liquid in the circumferential direction (Butterworth, 1968;Butterworth, 1972; Fukano and Inatomi, 2003; Fukano andOusaka, 1989; Jayanti et al., 1990b; Sutharshan et al., 1995), andcircumferential variations in interfacial roughness cause secondarygas flows which slow film drainage due to gravity (Butterworth,1972; Darling and McManus, 1968; Flores et al., 1995; Laurinatet al., 1985; Lin et al., 1985). While it is not the goal of this workto offer a definitive explanation for liquid film sustainment, thestatistics reported here do provide some insight toward under-standing this phenomenon.

Cross-sectional liquid motion

In the present study, a band-growth algorithm was used to sep-arate the liquid in the simulation domain into contiguous regions.This allows us to separate the droplets from the base film and tracktheir displacement in time in order to compute droplet velocities,analogous to particle image velocimetry (PIV). The algorithm wasoriginally developed by Herrmann (2010) for tracking droplets dur-ing primary atomization and has been applied to bubble tracking indense fluidized beds (Pepiot and Desjardins, 2012; Capecelatro andDesjardins, 2013; Capecelatro et al., 2014). Isolating droplets fromthe base film allows us to compute their contribution to the totalliquid volume. Defining the operator h�ii as an average with respectto coordinate i, Table 8 shows the temporal mean of the ratio ofliquid droplet volume Vd

l to total liquid volume Vl for the threecases, and it is clear that the amount of liquid in the form of dropsdecreases as Frg increases. It follows naturally that axial mass trans-port of liquid in the form of drops, denoted by _md

l , also decreaseswith increasing Frg , as shown by the mean ratio of _md

l to the totalliquid flow rate in Table 8. This is in keeping with qualitative anal-ysis of Fig. 3 as well as previous experimental results (Simmons and

Table 7Error of Schubring and Shedd (2009b) film asymmetrycorrelation (Eq. (35)) for the simulations.

Case A Case B Case C

Error 0% 68.09% 93.30%

Hanratty, 2001), which suggest that increased stratification leads tosignificantly fewer disperse drops within the gas core.

We use the orthoradial phase-averaged liquid velocity to exam-ine the liquid motion in the pipe cross section, defined as

ulðr; hÞ ¼alðx; tÞurhðx; tÞh ix;t

alðx; tÞh ix;t: ð42Þ

Vectors of ul in the r � h cross-section for cases B and C are shown inFig. 17, normalized by ul, the bulk axial liquid velocity for case A.We compute ul for halix;t > 0:05 in order to obtain clean statistics,since converged drop statistics in the gas core require longer runtimes than are tractable in the present study. Relative to the meanfilm height, the direction of the arrows for case B in Fig. 17(a) indi-cates that the film is predominantly draining on the wall side of themean interface location. Just across the mean interface on the gasside, however, there is clearly flow of liquid in the positive h direc-tion until h � p=2, indicating the circumferential propagation ofwaves that protrude into the gas core to a height larger than themean film height. The termination of this wave propagation nearh � p=2 suggests that wave pumping is not solely responsible fortransporting liquid to the top of the pipe. Given that 26% of theliquid flux through the cross section is in the form of drops, it is pos-sible that drop deposition could account for sustainment in upperregions of the pipe.

The scenario is very different for the heavily stratified film ofcase C shown in Fig. 17(b), which depicts circumferential motionof liquid until h � p=4, before the flow returns along the walltoward h ¼ 0. The flows within the film from the opposite sidesof the pipe meet at h ¼ 0 and are forced to the surface, leading tothe establishment of a counter-rotating vortex pair (CVP). In qual-itative agreement with Fig. 3(i), negligible transport of liquid in theform of drops indicates that droplets do not act to replenish thefilm.

Ratio of liquid drop volume to total liquid volume andratio of axial drop flow rate to total liquid flow rate.

Case Frg Vdl =Vl

D Et

_mdl = _ml

� t

A 1 0.063 0.37B 4.56 0.039 0.26C 1.20 0.0005 0.0024

Page 14: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

Fig. 17. Liquid motion in the r—h plane. Arrows indicate direction and the color palettes give magnitude. The solid black line shows the mean location of the interface. (a) and(c) are normalized by the bulk axial liquid velocity for case A, ul . (For interpretation of the references to color in this figure legend, the reader is referred to the web version ofthis article.)

J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105 101

Cross-sectional gas motion

The idea that circumferential variations in pipe surface rough-ness induce orthoradial circulation was initially proposed in theexperimental work of Darling and McManus (1968), and thismechanism is believed to be relevant to horizontal annular flowsdue to the increased waviness of the interface in lower regions ofthe pipe relative to upper regions. Numerical simulations ofJayanti et al. (1990b) showed secondary gas flows by using wallfunctions to represent differential wall roughness, and possiblestreamlines of secondary flows in horizontal annular flows wereproposed based on the experiments of Jayanti et al. (1990a). Lateron, the experiments of Flores et al. (1995) confirmed the existenceof secondary gas flows in an air–water annular flow. Recently, van’tWestende et al. (2007) performed large-eddy simulations to studythe effects of gravitational settling, turbophoresis, and secondarygas flows on droplets by modeling the film as a wall with circum-ferentially varying roughness and modeling the droplets as solidspheres. Their conclusion regarding secondary flows was that gascirculation alters deposition of the droplets, and secondary flowcentrifugal effects can lead to increased deposition in regions ofhigh orthoradial gas velocity. Secondary gas flows induced by hvariations in interfacial topology are physically constrained tomaintain a net-zero circulation, explaining the CVP that has beenobserved (Jayanti et al., 1990b; van’t Westende et al., 2007).

We write the phase-averaged gas velocity in the cross section as

ugðr; hÞ ¼agðx; tÞurhðx; tÞ�

x;t

agðx; tÞ�

x;t

: ð43Þ

Fig. 18 shows ug for cases B and C, normalized by ug , the bulkaxial gas phase-averaged velocity for case A. A CVP is clearly pres-ent within the gas core. The strength of the vortices for case C rel-ative to case B lead to a more clearly defined vortex center, likelydue to the absence of disperse droplets. Bringing the gas phase vor-tices together with the motion of liquid film in Fig. 17 leads tooppositely signed circulation within each quadrant of the pipewhen the average velocity in the r—h plane is computed, defined as

�uðr; hÞ ¼ urhðx; tÞh ix;t : ð44Þ

Fig. 19 shows streamlines of �u and their magnitude. Vortices inthe stratified film appear much more coherent for case C than forcase B. This is due to the increased film thickness and faster gas cir-culation speed, which is nearly 10% of ubulk, compared to less than4% for case B.

Fig. 20 shows the h component of �u as a function of y=R and h=pfor cases B and C. Fig. 20(a) shows that �uh < 0 for y < h, then �uh

increases to a local maximum on the gas side of the interface.Essentially the same behavior is observed for case C, as shown inFig. 20(b).

A model for the liquid film

The concepts of secondary gas flows and wave pumping lead toa simple model that aids in characterizing the behavior of theliquid film. Assuming the mean film thickness h is small comparedto the radius of curvature of the pipe, the ðy; hÞ polar coordinates(recall that y ¼ R� r) are approximated by a local Cartesian coordi-nate system, i.e., ðy; hÞ ! ðY;XÞ, where X is the wall-parallel coor-dinate and Y is the wall-normal coordinate, as shown in Fig. 21. Inthe model, the liquid film is characterized by gravitational acceler-ation and shear at the interface location h, due to the h componentof the secondary gas flow �uh. Assuming that the velocity profileacross the liquid film is laminar, is in the X-direction, and variesonly with Y, the momentum equation in the simplified coordinatesystem gives

@2uX@Y2 ¼

gXmlþ 1

ll

@p@X ; ð45Þ

where gX is the component of gravity in the negative X-direction,and @p=@X is the pressure gradient in the X-direction. The bound-ary conditions satisfied by uX ðYÞ are

uX ðY ¼ 0Þ ¼ 0uX ðY ¼ hÞ ¼ �uhðy ¼ hÞ ¼ uh;C:

ð46Þ

Utilizing these boundary conditions and acknowledging thatgX ¼ g sin h, the velocity across the film is given as

uXðYÞ ¼uh;C

hY þ 1

2g sin h

mþ 1

ll

@p@X

� �Y2 � Yh� �

: ð47Þ

Although the circumferential pressure gradient could be computedfrom the simulation, the value of @p=@X provided to Eq. (47) is opti-mized in order to minimize jh�uhih � huXihj, where the operator

ð�Þh ih ¼1h

Z h

0ð�Þdy ð48Þ

has been used to denote a film-averaged quantity, and h�uhih andhuX ih are the bulk velocities within the film for the simulationand the model, respectively. The resulting values of ð@p=@XÞ=f b vary

Page 15: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

Fig. 18. Gas motion in the r—h plane. Arrows indicate direction and the color palettes give magnitude. The solid black line shows the mean location of the interface. ug isnormalized by the bulk axial gas velocity for case A, ug . (For interpretation of the references to color in this figure legend, the reader is referred to the web version of thisarticle.)

Fig. 19. Mean cross-sectional velocity �uðr; hÞ normalized by ubulk, the bulk axial velocity for case A. The thick black line shows the mean location of the interface, color palettesgive velocity magnitude, and thin black lines with arrows show streamlines. (For interpretation of the references to color in this figure legend, the reader is referred to the webversion of this article.)

Fig. 20. �uh Normalized by ubulk (the bulk axial velocity for case A) as a function of y=R : h=p ¼ 1=4 (solid line), h=p ¼ 1=2 (dashed line), h=p ¼ 3=4 (dash-dotted line), hðhÞ=R(vertical lines).

102 J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105

within 30–70% (recall that f b is the axial source term described in‘Flow forcing’).

Fig. 22 compares the model velocity uX with the computed hvelocity �uh at different values of h for case B. The orthoradial veloc-ity is well-approximated by the simple model inside the liquidfilm, which is the intended goal. Table 9 gives the L2 error of thevelocity profile between the model and simulation within theregion 0 < Y < h for each value of h. Low values of the L2 profileerror indicate that the orthoradial velocity in the film is well-approximated by a laminar parabolic profile.

An important conclusion is that the bulk velocity of the film inthe X-direction is negative for all values of h. This implies thatinterfacial shear induced by secondary gas flow is not strongenough to overcome gravitational draining of the film, an argu-ment supported by previous findings (Jayanti et al., 1990b). Thisis further elucidated by using Eqs. (47) and (48) to write outhuXih, which yields

uXh ih ¼uh;C

2� h2

12g sin h

mlþ 1

ll

@p@X

� �: ð49Þ

Page 16: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

Fig. 21. Schematic of the liquid film model.

Fig. 22. Comparison between computed velocity �uh (solid line) and model velocity ujh�uhih � huX ihj (dash-dotted line). The location of the interface is shown by the vertical d

J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105 103

This implies that huX ih > 0 if the condition

uh;C >h2

6g sin h

mlþ 1

ll

@p@X

� �ð50Þ

is satisfied. Since sin h > 0 for 0 < h < p; @p=@X is positive for all h inthe model, and uh;C < 0 in Fig. 22, this cannot occur. However, thefact that the bulk film velocity is largest at the minimum value ofh is in keeping with Eq. (50). This further strengthens the idea thatsecondary flows can slow down gravitational drainage throughshear, but the shear is not strong enough to move liquid up the wall.Essentially the same conclusion is drawn when applying the modelto case C, but the lack of resolution in the liquid film for large h dueto heavy stratification makes the analysis less informative, and it istherefore not shown.

X , with mt computed from the mixing length (dashed line) and by minimizingotted line.

Page 17: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

Table 9L2 Profile error between the simulation and themodel.

h L2 Error

p=6 9:195� 10�5

p=3 4:983� 10�5

p=2 4:697� 10�5

2p=3 3:691� 10�5

5p=6 1:188� 10�5

104 J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105

Conclusion

In this work a general strategy for numerical simulation ofliquid–gas annular pipe flows is outlined. A turbulent gas core, afully deformable phase interface, and non-unity density and vis-cosity ratios are all accounted for, allowing for an original numer-ical exploration of gravitational effects in horizontal liquid–gasflows from first principles. We provide descriptions of the phaseinterface capture scheme, the immersed boundary method usedto represent the enclosing pipe, and the mechanism used to forcethe two-phase flow. The mesh and immersed boundary used inthe two-phase simulations are validated against a correspondingsingle-phase turbulent pipe flow, showing excellent agreementfor both first and second order velocity statistics.

Governing parameters of the system are shown through dimen-sional analysis and used to characterize the global behavior ofthree simulations with increasing importance of gravitationaleffects relative to inertial effects. By increasing the Froude numberthrough numerically decreasing gravitational acceleration, simula-tions within the annular and stratified-annular regime are per-formed. Results indicate that the present approach is capable ofcapturing the transition between these regimes, which was a prin-cipal directive of this exploratory study.

Detailed statistics of the flows are provided, including howstratification of liquid droplets and the base film modifies liquidvolume fraction and axial velocity profiles and increases the prob-ability of encountering dry-out conditions. Modification to the tur-bulent law of the wall by the liquid film is investigated, and it isseen that a viscous sublayer is observed within the liquid film,and both viscous and logarithmic regions exist within the gas core.The void fraction, liquid film height, and film asymmetry in thesimulations are all shown to agree within reason with experimen-tal data and correlations.

Statistical flow features within the pipe cross section are ana-lyzed, and results suggest that, for the case in which the film isnot significantly drained, the action of surface waves drives liquidroughly halfway up the pipe walls. Droplets account for nearly 30%of liquid transport for this case, compared to less than 1% for theheavily stratified case, leading to the possibility that preferentialdroplet deposition could play a role in film replenishment in upperregions of the pipe. Secondary gas flows are observed within thepipe cross section, likely induced by circumferential gradients ininterfacial roughness. Shearing of the liquid film imposed by sec-ondary gas flow acts to slow the gravitational drainage of the film,an idea that is further substantiated by a simple model for the film.

Previous computations have been performed that simulate sim-plified aspects of a horizontal liquid–gas annular flow, such asmodeling the base film through wall roughness functions, model-ing entrained droplets with solid particles, or detailed simulationof a sheared phase interface. This exploratory study demonstratesthat it is feasible to perform simulations that combine the majorityof the physical processes that occur in realistic annular flows andcapture global features like flow regime transition. The hope of thiswork is to promote future similar studies, as much can be learned

from a direct simulation approach that may not be attainedthrough reduced order models or experimental correlation andobservation alone.

Acknowledgements

This work is funded in part by Abengoa Research, Contract No.OCG 5478 B. Support from the National Institute for ComputationalSciences, made possible through an NSF TeraGrid allocation, isgratefully acknowledged.

References

Adechy, D., 2004. Modelling of annular flow through pipes and T-junctions. Comput.Fluids 33 (2), 289–313.

Alekseenko, S., Cherdantsev, A., Markovich, D., Rabusov, A., 2013. Dynamics ofheavy droplets and large bubbles in annular gas–liquid flow. In: 8thInternational Conference on Multiphase Flow, Jeju, Korea, pp. 1–8.

Al-Sarkhi, A., Hanratty, T., 2002. Effect of pipe diameter on the drop size in ahorizontal annular gas–liquid flow. Int. J. Multiphase Flow 28 (10), 1617–1629.

Al-Sarkhi, A., Sarica, C., Qureshi, B., 2012. Modeling of droplet entrainment in co-current annular two-phase flow: a new approach. Int. J. Multiphase Flow 39,21–28.

Alves, G., 1954. Cocurrent liquid–gas flow in a pipe-line contactor. Chem. Eng. Prog.50 (9), 449–456.

Anderson, R.J., Russel, T.W.F., 1970. Circumferential variation of interchange inhorizontal annular two-phase flow. Ind. Eng. Chem. Fundam. 9 (3), 340–344.

Azzopardi, B., 1999. Turbulence modification in annular gas/liquid flow. Int. J.Multiphase Flow 25 (6-7), 945–955.

Baker, O., 1953. Design of pipelines for the simultaneous flow of oil and gas. In: FallMeeting of the Petroleum Branch of AIME.

Bonizzi, M., Issa, R., 2003a. On the simulation of three-phase slug flow in nearlyhorizontal pipes using the multi-fluid model. Int. J. Multiphase Flow 29 (11),1719–1747.

Bonizzi, M., Issa, R., 2003b. A model for simulating gas bubble entrainment in two-phase horizontal slug flow. Int. J. Multiphase Flow 29 (11), 1685–1717.

Bunner, B., Tryggvason, G., 1999. Direct numerical simulations of three-dimensionalbubbly flows. Phys. Fluids 11 (8), 1967–1969.

Bunner, B., Tryggvason, G., 2002a. Dynamics of homogeneous bubbly flows – Part 1:Rise velocity and microstructure of the bubbles. J. Fluid Mech. 466, 17–52.

Bunner, B., Tryggvason, G., 2002b. Dynamics of homogeneous bubbly flows – Part 2:Velocity fluctuations. J. Fluid Mech. 466, 53–84.

Butterworth, D., 1968. Air–water climbing film flow in an eccentric annulus. In: Int.Symposium on Research on C-Current Gas–Liquid Flow, Waterloo, Ontario, pp.145–201.

Butterworth, D., 1972. Air–water annular flow in a horizontal tube. Prog. Heat MassTransfer 6, 235–251.

Butterworth, D., 1975. A comparison of some void-fraction relationships for co-current gas–liquid flow. Int. J. Multiphase Flow 1 (6), 845–850.

Capecelatro, J., Desjardins, O., 2013. An Euler–Lagrange strategy for simulatingparticle-laden flows. J. Comput. Phys. 238, 1–31.

Capecelatro, J., Pepiot, P., Desjardins, O., 2014. Numerical characterization andmodeling of particle clustering in wall-bounded vertical risers. Chem. Eng. J.245, 295–310.

Carey, V.P., 2008. Liquid–Vapor Phase-Change Phenomena: An Introduction to theThermophysics of Vaporization and Condensation Processes in Heat TransferEquipment. Taylor & Francis Group.

Chisholm, D., 1973. Pressure gradients due to friction during the flow of evaporatingtwo-phase mixtures in smooth tubes and channels. Int. J. Heat Mass Transfer 16(2), 347–358.

Coddington, P., Macian, R., 2002. A study of the performance of void fractioncorrelations used in the context of drift-flux two-phase flow models. Nucl. Eng.Des. 215 (3), 199–216.

Dallman, J.C., 1978. Investigation of Separated Flow Model in Annular Gas–LiquidTwo-Phase Flows. Ph.D. Thesis, University of Illinois, Urbana.

Darling, R., McManus, H., 1968. Flow patterns in circular ducts with circumferentialvariation of roughness: a two-phase flow analog. In: Proceedings of the 11thMidwestern Mechanics Conference, vol. 5, pp. 153–170.

Desjardins, O., Blanquart, G., Balarac, G., Pitsch, H., 2008a. High order conservativefinite difference scheme for variable density low Mach number turbulent flows.J. Comput. Phys. 227 (15), 7125–7159.

Desjardins, O., Moureau, V., Pitsch, H., 2008b. An accurate conservative level set/ghost fluid method for simulating turbulent atomization. J. Comput. Phys. 227(18), 8395–8416.

Desjardins, O., McCaslin, J.O., Owkes, M., Brady, P., 2013. Direct numerical and large-eddy simulation of primary atomization in complex geometries. Atomizat.Sprays 23 (11), 1001–1048.

Esmaeeli, A., Tryggvason, G., 1998. Direct numerical simulations of bubbly flows –Part 1: Low Reynolds number arrays. J. Fluid Mech. 377, 313–345.

Esmaeeli, A., Tryggvason, G., 1999. Direct numerical simulations of bubbly flows –Part 2: Moderate Reynolds number arrays. J. Fluid Mech. 385, 325–358.

Page 18: International Journal of Multiphase Flow - Cornell University Desjardins. 2014... · International Journal of Multiphase Flow ... first principles-based study of a horizontal liquid–gas

J.O. McCaslin, O. Desjardins / International Journal of Multiphase Flow 67 (2014) 88–105 105

Farias, P.S.C. Martins, F.J.W.a., Sampaio, L.E.B., Serfaty, R., Azevedo, L.F.a. Liquid filmcharacterization in horizontal, annular, two-phase, gas–liquid flow using time-resolved laser-induced fluorescence. Exp. Fluids.

Fedkiw, R.P., Aslam, T.D., Merriman, B., Osher, S., 1999. A non-oscillatory Eulerianapproach to interfaces in multimaterial flows (the ghost fluid method). J.Comput. Phys. 152 (2), 457–492.

Flores, A.G., Crowe, K.E., Griffith, P., 1995. Gas-phase secondary flow in horizontal,stratified and annular two-phase flow. Int. J. Multiphase Flow 21 (2), 207–221.

Fukagata, K., Kasagi, N., 2002. Highly energy-conservative finite difference methodfor the cylindrical coordinate system. J. Comput. Phys. 181 (2), 478–498.

Fukano, T., Inatomi, T., 2003. Analysis of liquid film formation in a horizontalannular flow by DNS. Int. J. Multiphase Flow 29 (9), 1413–1430.

Fukano, T., Ousaka, A., 1989. Prediction of the circumferential distribution of filmthickness in horizontal and near-horizontal gas–liquid annular flows. Int. J.Multiphase Flow 15 (3), 403–419.

Fulgosi, M., Lakehal, D., Banerjee, S., De Angelis, V., 2003. Direct numericalsimulation of turbulence in a sheared air–water flow with a deformableinterface. J. Fluid Mech. 482, 319–345.

Ghajar, A.J., Tang, C.C., 2007. Heat transfer measurements, flow pattern maps, andflow visualization for non-boiling two-phase flow in horizontal and slightlyinclined pipe. Heat Transfer Eng. 28 (6), 525–540.

Hagiwara, Y., Esmaeilzadeh, E., Tsutsui, H., Suzuki, K., 1989. Simultaneousmeasurement of liquid film thickness, wall shear stress and gas flowturbulence of horizontal wavy two-phase flow. Int. J. Multiphase Flow 15 (3),421–431.

Han, H., Gabriel, K., 2007. A numerical study of entrainment mechanism inaxisymmetric annular gas–liquid flow. J. Fluids Eng. 129 (3), 293.

Harms, T.M., Li, D., Groll, E.a., Braun, J.E., 2003. A void fraction model for annularflow in horizontal tubes. Int. J. Heat Mass Transfer 46 (21), 4051–4057.

Herrmann, M., 2010. A parallel Eulerian interface tracking/Lagrangian point particlemulti-scale coupling procedure. J. Comput. Phys. 229 (3), 745–759.

Hewitt, G.F., Jayanti, S., Hope, C.B., 1990. Structure of thin liquid films in gas–liquidhorizontal flow. Int. J. Multiphase Flow 16 (6), 951–957.

Hoogendoorn, C., 1959. Gas–liquid flow in horizontal pipes. Chem. Eng. Sci. 9 (4),205–217.

Hu, B., Nuland, S., Lawrence, C., 2013. Phase distribution and interface structure ofgas–liquid stratified wavy flows in a large-diameter high-pressure pipeline. In:8th International Conference on Multiphase Flow, Jeju, Korea.

Hurlburt, E., Newell, T., 1996. Optical measurement of liquid film thickness andwave velocity in liquid film flows. Exp. Fluids 21 (5), 357–362.

Hurlburt, E.T., Newell, T.a., 2000. Prediction of the circumferential film thicknessdistribution in horizontal annular gas–liquid flow. J. Fluids Eng. 122 (2), 396.

Issa, R.I., Kempf, M.H.W., 2003. Simulation of slug flow in horizontal and nearlyhorizontal pipes with the two-fluid model. Int. J. Multiphase Flow 29 (1), 69–95.

Jacowitz, L., Brodkey, R.S., 1964. An analysis of geometry and pressure drop for thehorizontal, annular, two-phase flow of water and air in the entrance region of apipe. Chem. Eng. Sci. 19 (4), 261–274.

Jayanti, S., Hewitt, G., White, S., 1990a. Time-dependent behaviour of the liquid filmin horizontal annular flow. Int. J. Multiphase Flow 16 (6), 1097–1116.

Jayanti, S., Wilkes, N.S., Clarke, D.S., Hewitt, G.F., 1990b. The prediction of turbulentflows over roughened surfaces and its application to interpretation ofmechanisms of horizontal annular flow. Proc. Roy. Soc. A: Math. Phys. Eng.Sci. 431 (1881), 71–88.

Kadambi, V., 1982. Stability of annular flow in horizontal tubes. Int. J. MultiphaseFlow 8 (4), 311–328.

Kang, S., Choi, H., 2000. Active wall motions for skin-friction drag reduction. Phys.Fluids 12 (12), 3301.

Kosterin, S., 1949. An investigation of the influence of the diameter and inclinationof a tube on the hydraulic resistance and flow structure of gas–liquid mixtures.Izvest. Akad. Nauk. SSSR, Otdel Tekh Nauk 12, 1824–1830.

Krasiakova, L., 1957. Some characteristics of the flow of a two phase mixture in ahorizontal pipe. Atom. Energy Res. Establish.

Lakehal, D., Fulgosi, M., Yadigaroglu, G., Banerjee, S., 2003. Direct numericalsimulation of turbulent heat transfer across a mobile, sheared gas–liquidinterface. J. Heat Transfer 125 (6), 1129–1139.

Laurinat, J., Hanratty, T., 1982.Studies of the Effects of Pipe Size on HorizontalAnnular Two-Phase Flows, Ph.D. Thesis, University of Illinois, Urbana.

Laurinat, J., Hanratty, T., Jepson, W., 1985. Film thickness distribution for gas–liquidannular flow in a horizontal pipe. PhysicoChem. Hydrodynam. 6, 179–195.

Lin, T., Jones, O., Lahey, R., Block, R., Murase, M., 1985. Film thickness measurementsand modelling in horizontal annular flows. PhysicoChem. Hydrodynam. 6, 197–206.

Lombardi, P., De Angelis, V., Banerjee, S., 1996. Direct numerical simulation of near-interface turbulence in coupled gas–liquid flow. Phys. Fluids 8 (6), 1643.

Lu, J., Fernandez, A., Tryggvason, G., 2005. The effect of bubbles on the wall drag in aturbulent channel flow. Phys. Fluids 17 (9), 095102.

Mandhane, J., Gregory, G., Aziz, K., 1974. A flow pattern map for gas–liquid flow inhorizontal pipes. Int. J. Multiphase Flow 1 (4), 537–553.

Martinelli, R., Nelson, D., 1948. Prediction of pressure drop during forced-circulationboiling of water. Trans. Asme 70 (6), 695–702.

McCaslin, J.O., Desjardins, O., 2014. A localized re-initialization equation for theconservative level set method. J. Comput. Phys. 262, 408–426.

Meyer, M., Devesa, A., Hickel, S., Hu, X., Adams, N., 2010. A conservative immersedinterface method for Large-Eddy Simulation of incompressible flows. J. Comput.Phys. 229 (18), 6300–6317.

Moreno Quibén, J., Thome, J.R., 2007. Flow pattern based two-phase frictionalpressure drop model for horizontal tubes – Part II: New phenomenologicalmodel. Int. J. Heat Fluid Flow 28 (5), 1060–1072.

Murai, Y., Fukuda, H., Oishi, Y., Kodama, Y., Yamamoto, F., 2007. Skin frictionreduction by large air bubbles in a horizontal channel flow. Int. J. MultiphaseFlow 33 (2), 147–163.

Nagrath, S., Jansen, K.E., Lahey, R.T., 2005. Computation of incompressible bubbledynamics with a stabilized finite element level set method. Comput. MethodsAppl. Mech. Eng. 194 (42–44), 4565–4587.

Ooms, G., Poesio, P., 2003. Stationary core-annular flow through a horizontal pipe.Phys. Rev. E 68 (6), 066301.

Ooms, G., Segal, A., Vanderwees, A., Meerhoff, R., Oliemans, R., 1983. A theoreticalmodel for core-annular flow of a very viscous oil core and a water annulusthrough a horizontal pipe. Int. J. Multiphase Flow 10 (1), 41–60.

Ooms, G., Vuik, C., Poesio, P., 2007. Core-annular flow through a horizontal pipe:hydrodynamic counterbalancing of buoyancy force on core. Phys. Fluids 19 (9),092103.

Ooms, G., Pourquie, M., Poesio, P., 2012. Numerical study of eccentric core-annularflow. Int. J. Multiphase Flow 42, 74–79.

Owkes, M., Desjardins, O., 2013. A discontinuous Galerkin conservative level setscheme for interface capturing in multiphase flows. J. Comput. Phys. 249 (1),275–302.

Öztürk, O.C., Marchand, M., Berlandis, J.-p., Lance, M., 2013. Experimental andanalytical investigation of various horizontal two-phase flow regimes. In: 8thInternational Conference on Multiphase Flow, Jeju, Korea, pp. 1–7.

Paras, S., Karabelas, A., 1991a. Properties of the liquid layer in horizontal annularflow. Int. J. Multiphase Flow 17 (4), 439–454.

Paras, S.V., Karabelas, A.J., 1991b. Droplet entrainment and deposition in horizontalannular flow. Int. J. Multiphase Flow 17 (4), 455–468.

Pepiot, P., Desjardins, O., 2012. Numerical analysis of the dynamics of two- andthree-dimensional fluidized bed reactors using an Euler–Lagrange approach.Powder Technol. 220, 104–121.

Pope, S.B., 2000. Turbulent Flows. Cambridge University Press.Reboux, S., Sagaut, P., Lakehal, D., 2006. Large-eddy simulation of sheared interfacial

flow. Phys. Fluids 18 (10), 105105.Rodriguez, D., Shedd, T.A., 2004. Entrainment of gas in the liquid film of horizontal,

annular, two-phase flow. Int. J. Multiphase Flow 30 (6), 565–583.Russell, T.W.F., Lamb, D.E., 1965. Flow mechanism of two-phase annular flow.

Canad. J. Chem. Eng. 43 (5), 237–245.Schubring, D., Shedd, T., 2008. Wave behavior in horizontal annular air–water flow.

Int. J. Multiphase Flow 34 (7), 636–646.Schubring, D., Shedd, T., 2009a. Prediction of wall shear for horizontal annular air–

water flow. Int. J. Heat Mass Transfer 52 (1–2), 200–209.Schubring, D., Shedd, T., 2009b. Critical friction factor modeling of horizontal

annular base film thickness. Int. J. Multiphase Flow 35 (4), 389–397.Schubring, D., Shedd, T., 2011. A model for pressure loss, film thickness, and

entrained fraction for gas–liquid annular flow. Int. J. Heat Fluid Flow 32 (3),730–739.

Schubring, D., Ashwood, A., Shedd, T., Hurlburt, E., 2010a. Planar laser-inducedfluorescence (PLIF) measurements of liquid film thickness in annular flow. PartI: Methods and data. Int. J. Multiphase Flow 36 (10), 815–824.

Schubring, D., Shedd, T., Hurlburt, E., 2010b. Planar laser-induced fluorescence(PLIF) measurements of liquid film thickness in annular flow. Part II: Analysisand comparison to models. Int. J. Multiphase Flow 36 (10), 825–835.

Serdar Kaya, A., Tom Chen, X., Sarica, C., Brill, J.P., 2000. Investigation of transitionfrom annular to intermittent flow in pipes. J. Energy Resour. Technol. 122 (1),22.

Shedd, T.a., Newell, T.a., 2004. Characteristics of the liquid film and pressure drop inhorizontal, annular, two-phase flow through round, square and triangulartubes. J. Fluids Eng. 126 (5), 807.

Simmons, M.J.H., Hanratty, T.J., 2001. Droplet size measurements in horizontalannular gas–liquid flow. Int. J. Multiphase Flow 27 (5), 861–883.

Solbakken, S., Andersson, H.I., 2005. Direct numerical simulation of lubricatedchannel flow. Fluid Dynam. Res. 37 (3), 203–230.

Sutharshan, B., Kawaji, M., Ousaka, A., 1995. Measurement of circumferential andaxial liquid film velocities in horizontal annular flow. Int. J. Multiphase Flow 21(2), 193–206.

Taitel, Y., Dukler, A.E., 1976. A model for predicting flow regime transitions inhorizontal and near horizontal gas–liquid flow. AIChE J. 22 (1), 47–55.

Tandon, T.N., Varma, H.K., Gupta, C.P., 1985. A void fraction model for annular two-phase flow. Int. J. Heat Mass Transfer 28 (I), 191–198.

van’t Westende, J., Belt, R., Portela, L., Mudde, R., Oliemans, R., 2007. Effect ofsecondary flow on droplet distribution and deposition in horizontal annularpipe flow. Int. J. Multiphase Flow 33 (1), 67–85.

Williams, L., 1990. Effect of Pipe Diameter on Horizontal Annular Two-Phase Flow.Ph.D. Thesis, University of Illinois, Urbana.

Woldesemayat, M.a., Ghajar, A.J., 2007. Comparison of void fraction correlations fordifferent flow patterns in horizontal and upward inclined pipes. Int. J.Multiphase Flow 33 (4), 347–370.