international journal of heat and mass transfersashi/pub/ganesanetal_ijhmt_2014.pdf · mac-type...

18
Numerical modeling of the non-isothermal liquid droplet impact on a hot solid substrate Sashikumaar Ganesan a,, Sangeetha Rajasekaran a , Lutz Tobiska b a Numerical Mathematics and Scientific Computing, SERC, Indian Institute of Science, Bangalore 560 012, India b Institute of Analysis and Numerical Computation, OvG University, PF4120, 39016 Magdeburg, Germany article info Article history: Received 11 November 2013 Received in revised form 12 June 2014 Accepted 8 July 2014 Keywords: Finite elements ALE approach Moving mesh Impinging droplets Moving contact line Dynamic contact angle Heat transfer abstract The heat transfer from a solid phase to an impinging non-isothermal liquid droplet is studied numeri- cally. A new approach based on an arbitrary Lagrangian–Eulerian (ALE) finite element method for solving the incompressible Navier–Stokes equations in the liquid and the energy equation within the solid and the liquid is presented. The novelty of the method consists in using the ALE-formulation also in the solid phase to guarantee matching grids along the liquid–solid interface. Moreover, a new technique is developed to compute the heat flux without differentiating the numerical solution. The free surface and the liquid–solid interface of the droplet are represented by a moving mesh which can handle jumps in the material parameter and a temperature dependent surface tension. Further, the application of the Laplace–Beltrami operator technique for the curvature approximation allows a natural inclusion of the contact angle. Numerical simulation for varying Reynold, Weber, Peclét and Biot numbers are performed to demonstrate the capabilities of the new approach. Ó 2014 Elsevier Ltd. All rights reserved. 1. Introduction The interaction between sprays and hot solid objects occur in a wide variety of industrial and environmental applications. Never- theless, our understanding of the mechanisms involved in the pro- cess is far from being complete. Therefore, the study of the hydrodynamic and thermodynamic behavior of a single droplet impinging on a solid substrate is of fundamental importance. As a droplet nears a hot solid substrate, heat is transferred from the solid to the liquid phase by conduction, convection and radia- tion. This energy is used to increase the temperature of the liquid or to vaporize the liquid from the base of the droplet. In the latter case a direct contact between the solid and the liquid phase is excluded (Leidenfrost phenomenon). For a surface temperature below the Leidenfrost temperature we suppose that the direct liquid-wall contact and the kinetic of the droplet spreading domi- nate the heat transfer. Thus, we consider in this paper the coupled heat transfer process in a single deforming droplet and in the solid phase during the spreading and recoiling from the moment of its impact till losing a direct contact with the solid. Despite several advances made in the field of Computational Fluid Dynamics, modeling and simulation of these processes are still very challeng- ing. The Volume-of-Fluid [1–5], Level set [6–10], Immersed Boundary/Front Tracking [11–15] and the arbitrary Lagrangian Eulerian [16–21] are the most commonly used methods for tracking/capturing moving interfaces/boundaries. Although several interface capturing/tracking methods have been proposed in the literature, only a few numerical studies incor- porated the thermodynamic behavior. In the early study [22], MAC-type solution technique has been used for the computations of a water droplet impinging on a flat surface above the Leidenfrost temperature. However, the unsteady heat transfer computation has been neglected in this study. A considerable number of numer- ical simulations using the Volume-of-Fluid method have been reported in the literature for droplets impinging on a hot surface [23–29]. Also, numerical studies using the Level set method [30,31] and the Immersed Boundary method [32,33] have been reported in the literature. All these methods can be classified as fixed grid (Eulerian) methods. Computations using the Lagrangian approach for a liquid droplet impinging on a hot solid substrate have been presented in [34]. To the best of the authors knowledge, computations of a liquid droplet impinging on a hot solid substrate using the arbitrary Lagrangian–Eulerian (ALE) approach in both phases have not been reported in the literature. In this paper, we present an accurate and efficient sharp inter- face ALE finite element approach for the computation of a non- isothermal liquid droplet impinging on a hot solid substrate. The Marangoni force, the surface force, and the jumps in the material http://dx.doi.org/10.1016/j.ijheatmasstransfer.2014.07.019 0017-9310/Ó 2014 Elsevier Ltd. All rights reserved. Corresponding author. E-mail addresses: [email protected] (S. Ganesan), [email protected]. ernet.in (S. Rajasekaran), [email protected] (L. Tobiska). International Journal of Heat and Mass Transfer 78 (2014) 670–687 Contents lists available at ScienceDirect International Journal of Heat and Mass Transfer journal homepage: www.elsevier.com/locate/ijhmt

Upload: phungkhue

Post on 06-Jul-2018

217 views

Category:

Documents


0 download

TRANSCRIPT

International Journal of Heat and Mass Transfer 78 (2014) 670–687

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer

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

Numerical modeling of the non-isothermal liquid droplet impacton a hot solid substrate

http://dx.doi.org/10.1016/j.ijheatmasstransfer.2014.07.0190017-9310/� 2014 Elsevier Ltd. All rights reserved.

⇑ Corresponding author.E-mail addresses: [email protected] (S. Ganesan), [email protected].

ernet.in (S. Rajasekaran), [email protected] (L. Tobiska).

Sashikumaar Ganesan a,⇑, Sangeetha Rajasekaran a, Lutz Tobiska b

a Numerical Mathematics and Scientific Computing, SERC, Indian Institute of Science, Bangalore 560 012, Indiab Institute of Analysis and Numerical Computation, OvG University, PF4120, 39016 Magdeburg, Germany

a r t i c l e i n f o

Article history:Received 11 November 2013Received in revised form 12 June 2014Accepted 8 July 2014

Keywords:Finite elementsALE approachMoving meshImpinging dropletsMoving contact lineDynamic contact angleHeat transfer

a b s t r a c t

The heat transfer from a solid phase to an impinging non-isothermal liquid droplet is studied numeri-cally. A new approach based on an arbitrary Lagrangian–Eulerian (ALE) finite element method for solvingthe incompressible Navier–Stokes equations in the liquid and the energy equation within the solid andthe liquid is presented. The novelty of the method consists in using the ALE-formulation also in the solidphase to guarantee matching grids along the liquid–solid interface. Moreover, a new technique isdeveloped to compute the heat flux without differentiating the numerical solution. The free surfaceand the liquid–solid interface of the droplet are represented by a moving mesh which can handle jumpsin the material parameter and a temperature dependent surface tension. Further, the application of theLaplace–Beltrami operator technique for the curvature approximation allows a natural inclusion of thecontact angle. Numerical simulation for varying Reynold, Weber, Peclét and Biot numbers are performedto demonstrate the capabilities of the new approach.

� 2014 Elsevier Ltd. All rights reserved.

1. Introduction

The interaction between sprays and hot solid objects occur in awide variety of industrial and environmental applications. Never-theless, our understanding of the mechanisms involved in the pro-cess is far from being complete. Therefore, the study of thehydrodynamic and thermodynamic behavior of a single dropletimpinging on a solid substrate is of fundamental importance.

As a droplet nears a hot solid substrate, heat is transferred fromthe solid to the liquid phase by conduction, convection and radia-tion. This energy is used to increase the temperature of the liquidor to vaporize the liquid from the base of the droplet. In the lattercase a direct contact between the solid and the liquid phase isexcluded (Leidenfrost phenomenon). For a surface temperaturebelow the Leidenfrost temperature we suppose that the directliquid-wall contact and the kinetic of the droplet spreading domi-nate the heat transfer. Thus, we consider in this paper the coupledheat transfer process in a single deforming droplet and in the solidphase during the spreading and recoiling from the moment of itsimpact till losing a direct contact with the solid. Despite severaladvances made in the field of Computational Fluid Dynamics,modeling and simulation of these processes are still very challeng-

ing. The Volume-of-Fluid [1–5], Level set [6–10], ImmersedBoundary/Front Tracking [11–15] and the arbitrary LagrangianEulerian [16–21] are the most commonly used methods fortracking/capturing moving interfaces/boundaries.

Although several interface capturing/tracking methods havebeen proposed in the literature, only a few numerical studies incor-porated the thermodynamic behavior. In the early study [22],MAC-type solution technique has been used for the computationsof a water droplet impinging on a flat surface above the Leidenfrosttemperature. However, the unsteady heat transfer computationhas been neglected in this study. A considerable number of numer-ical simulations using the Volume-of-Fluid method have beenreported in the literature for droplets impinging on a hot surface[23–29]. Also, numerical studies using the Level set method[30,31] and the Immersed Boundary method [32,33] have beenreported in the literature. All these methods can be classified asfixed grid (Eulerian) methods. Computations using the Lagrangianapproach for a liquid droplet impinging on a hot solid substratehave been presented in [34]. To the best of the authors knowledge,computations of a liquid droplet impinging on a hot solid substrateusing the arbitrary Lagrangian–Eulerian (ALE) approach in bothphases have not been reported in the literature.

In this paper, we present an accurate and efficient sharp inter-face ALE finite element approach for the computation of a non-isothermal liquid droplet impinging on a hot solid substrate. TheMarangoni force, the surface force, and the jumps in the material

S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687 671

parameters are incorporated into the model very accurately sincethe free surface and the liquid–solid interface are resolved by thecomputational mesh in the ALE approach. Also, spurious velocitiesarising often in other methods can be suppressed by using thisapproach [35]. Further, we use the Laplace–Beltrami operator tech-nique [36,37] to treat the curvature in a semi-implicit manner [38].This technique allows us to include the equilibrium contact angleinto the model weakly [39]. Since the moving liquid–solid interfaceis resolved by the computational mesh, the energy equation inboth the liquid and the solid phases can be solved by a one-fieldformulation. This allows us also to develop a new technique forcomputing the heat flux from the solid phase to the liquid phasewithout differentiating the numerical solution.

The paper is organized as follows. In Section 2 we introduce thegoverning equations for the fluid flow and the heat transfer in theliquid and solid phases. Then, the complex coupled problem is for-mulated in dimensionless quantities. Section 3 is devoted to theingredients of our numerical approach. First, we explain in detailhow the curvature approximation by the Laplace–Beltrami tech-nique and the contact angle condition can be implemented in theweak form of the Navier–Stokes equations. We derive a one-fieldformulation for the temperature in the liquid–solid domain. Nextthe ALE approach to handle the moving mesh is discussed. Then,under the assumption of axisymmetry, the fully weak 3D formula-tion is transferred to the weak axisymmetric formulation. The dis-cretization in space and time, in particular, the semi-discretizationof the curvature term are given. Finally, we address the mesh han-dling techniques at the liquid–solid interface and the inner meshupdate. Section 4 is concerned with numerical tests for varyingReynolds, Weber, Peclét and Biot numbers showing the capabilitiesof the new method. We shortly summarize the proposed numericalmethod and the obtained results in Section 5.

2. Mathematical model

2.1. Governing equations for the fluid flow

We consider a liquid droplet impinging on a horizontal hot sub-strate, see Fig. 1. The fluid is assumed to be incompressible; densityand viscosity are constant. The governing equations for the fluidflow describing the sequence of spreading and recoiling of animpinging droplet in the given time interval (0, I) are the time-dependent incompressible Navier–Stokes equations

@u@tþ ðu � rÞu� 1

qr � ðTðu;pÞÞ ¼ ge in XFðtÞ � ð0; IÞ; ð1Þ

r � u ¼ 0 in XFðtÞ � ð0; IÞ; ð2Þ

ΓSΩS

ΓF

ΩF

τF

τS

νFνS

ΓN

θ

Fig. 1. Computational model for an impinging droplet on a hot solid substrate.

with the initial condition

uð�;0Þ ¼ u0 in XFð0Þ ð3Þ

and the boundary conditions given below. The time-dependentdomain of the deforming droplet is denoted by XFðtÞ. The setXFðtÞ � ð0; IÞ is understood as fðx; tÞ 2 R4 : t 2 ð0; IÞ; x 2 XFðtÞg. Theboundary of the droplet satisfies @XFðtÞ ¼ CFðtÞ [ CSðtÞ, where CFðtÞand CSðtÞ are the free surface and the liquid–solid interface of thedroplet, respectively. Further, u is the velocity, p is the pressure, qis the density, g is the gravitational constant, t is the time, e is an unitvector in the opposite direction of the gravitational force,u0ð0; 0;�uimpÞ is a given initial velocity and I is a given final time.The stress tensor Tðu;pÞ for a Newtonian incompressible fluid isgiven by

Tðu;pÞ :¼ 2lDðuÞ � pI; DðuÞi;j ¼12

@ui

@xjþ @uj

@xi

� �;

with i; j ¼ 1; . . . ;3, where l denotes the dynamic viscosity, DðuÞ thevelocity deformation tensor and I the identity tensor. The kinematicand force balancing conditions on the free surface CF are given by

u � mF ¼ w � mF on CFðtÞ; ð4ÞTðu;pÞ � mF ¼ rðTFÞKmF þrrðTFÞ on CFðtÞ: ð5Þ

Here, mF denotes the unit outer normal vector on the free surface, Kdenotes the sum of the principal curvatures,r the tangential gradi-ent (defined in Section 3.1), w the velocity of the computationaldomain XFðtÞ; TF the temperature on CFðtÞ;rðTFÞ the temperature-dependent surface tension. Let si;F ; i ¼ 1;2; denote tangential vec-tors on the free surface CF(t). Then, (5) is equivalent to

mF � Tðu;pÞ � mF ¼ rðTFÞK; on CFðtÞ;si;F � Tðu;pÞ � mF ¼ si;F � rrðTFÞ on CFðtÞ;

since the tangential gradient on CFðtÞ is perpendicular to mF . Fur-ther, for simplicity we assume that the temperature-dependent sur-face tension rðTFÞ follows a simple linear law

rðTFÞ ¼ rref � C1ðTF � Tref Þ; ð6Þ

where Tref is some reference temperature, rref is the surface tensioncoefficient at the reference temperature and C1 > 0 is the negativerate of change of surface tension with temperature. The changesin the local surface temperature TFon the free surface induce varia-tions in the surface tension. Due to these variations in the surfacetension, the fluid moves away from the region of low surface ten-sion to the region of high surface tension. This effect is calledMarangoni convection. Note that for large TF the surface tension(6) could become negative and the model would be no longer cor-rect. In order to avoid this unphysical behavior we check in ournumerical computations whether the computed surface tensionvalues are in the range of validity of the law (6).

Next, for moving contact line problems the conventional no-slipboundary condition at fixed walls cannot be used along the solidsurface. Imposing the no-slip boundary condition at the movingcontact line, where the free surface and the liquid–solid interfacemeet, leads to an non-integrable force singularity at the contactline [40–43]. Therefore, we use instead the Navier–slip boundarycondition [39,44–46] on the liquid–solid interface CSðtÞ

u � mS ¼ 0; u � si;S þ �lðsi;S � Tðu; pÞ � mSÞ ¼ 0: ð7Þ

Here, mS and si;S; i ¼ 1;2 denote the outer unit normal and tangentialvectors on the liquid–solid interface. Further, �l is the slip coeffi-cient [39,47].

2.2. Heat transfer from the solid into the liquid phase

The heat transfer in the moving droplet and in the fixed solidphase are described by the energy equation. In this study, we

672 S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687

assume that the material properties such as densities, thermal con-ductivities and specific heats are constant in the liquid and thesolid phases, respectively. The solid phase will be denoted by XS

and is fixed over time. However, due to the changing wetting areaCSðtÞ the non-wetted part CNðtÞ :¼ @XS n CSðtÞ of the boundaryalong which we describe Neumann type boundary conditionsbecomes time-dependent. The heat transfer in the liquid and solidphase, respectively, is described by the energy equations

@TF

@tþ u � rTF ¼

kF

cFpq

DTF in XFðtÞ � ð0; IÞ; ð8Þ

@TS

@t¼ kS

cSpqS

DTS in XS � ð0; IÞ: ð9Þ

We complete (8) and (9) by the initial conditions

TFð�;0Þ ¼ TF;0 in XFð0Þ; TSð�;0Þ ¼ TS;0 in XS; ð10Þ

the boundary conditions at the free surface CFðtÞ and along the non-wetting part of the solid CNðtÞ

�kF@TF

@mF¼ aFðTF � T1Þ on CFðtÞ;

@TS

@n¼ 0 on CNðtÞ ð11Þ

and the transition conditions (continuity of temperature and heatflux)

kF@TF

@mF¼ �kS

@TS

@mS; TF ¼ TS on CSðtÞ: ð12Þ

Here, TF and TS are the temperatures in XFðtÞ and XS, respectively,kF ; kS the conductivities, q;qS the densities, cF

p; cSp the specific heat

of the fluid and the solid. Finally, T1 denotes the temperature ofthe surrounding gas, aF is the convection heat transfer coefficienton the liquid–gas interface, and n is the outer normal on CNðtÞ.

2.3. Dimensionless form of the model equations

In this section we derive the dimensionless form of the coupledNavier–Stokes and energy equations. We introduce the scaling fac-tors L and U as characteristic length and velocity, respectively. Fur-thermore, we define the dimensionless variables

~x ¼ xL; ~u ¼ u

U; ~w ¼ w

U; ~t ¼ tU

L; ~I ¼ IU

L; ~p ¼ p

qU2

and the dimensionless numbers (Reynolds, Weber, Froude and slip,respectively)

Re ¼ qULl ; We ¼ qU2L

rref; Fr ¼ U2

Lg; b� ¼ �lqU:

We introduce a dimensionless stress tensor Sð~u; ~pÞ by

Sð~u; ~pÞ ¼ 1qU2 Tð~u; ~pÞ ¼ 1

Re@~ui

@~xjþ @

~uj

@~xi

� �� pdij

� �i;j¼1;...;3

and the dimensionless temperature in the liquid phase

eT F ¼TF � T1Tref � T1

:

Writing the coupled system in the new dimensionless variables andomitting the tilde afterwards we end up with the dimensionlessNavier–Stokes problem

@u@tþ ðu � rÞu�r � Sðu;pÞ ¼ 1

Fre in XFðtÞ � ð0; IÞ; ð13Þ

r � u ¼ 0 in XFðtÞ � ð0; IÞ; ð14Þu � mS ¼ 0 on CSðtÞ � ð0; IÞ; ð15Þb�ðsi;S � Sðu;pÞ � mSÞ ¼ �u � si;S on CSðtÞ � ð0; IÞ; ð16Þu � mF ¼ w � mF on CFðtÞ � ð0; IÞ; ð17Þ

with the initial condition

uð�;0Þ ¼ u0=U:

The dimensionless form of the normal stress boundary condition onCF becomes

mF � Sðu; pÞ � mF ¼1

We1� C0ðTF � 1Þð ÞK ð18Þ

and the shear stress boundary condition reads

si;F � Sðu;pÞ � mF ¼ �C0

Wesi;F � rTF ; ð19Þ

where

C0 ¼C1

rref ðTref � T1Þ:

In order to write the energy equations in a dimensionless form,we define in addition the dimensionless temperature in the solidphase as

eT S ¼TS � T1

Tref � T1:

After transforming the energy equations as well as the initial,boundary and transition conditions in dimensionless form we omitthe tilde and obtain the dimensionless energy equations in theliquid and solid phase

@TF

@tþ u � rTF �

1PeF

DTF ¼ 0 in XFðtÞ � ð0; IÞ; ð20Þ

@TS

@t� 1

PeSDTS ¼ 0 in XS � ð0; IÞ; ð21Þ

respectively. The dimensionless initial conditions become

TFð�;0Þ ¼TF;0 � T1Tref � T1

in XFð0Þ; ð22Þ

TSð�;0Þ ¼TS;0 � T1Tref � T1

in XS: ð23Þ

Finally, the dimensionless thermal boundary and transition condi-tions can be written as

� @TF

@mF¼ BiTF on CFðtÞ � ð0; IÞ; ð24Þ

@TF

@mF¼ � kS

kF

@TS

@mSon CSðtÞ � ð0; IÞ; ð25Þ

TF ¼ TS on CSðtÞ � ð0; IÞ; ð26Þ@TS

@n¼ 0 on CNðtÞ � ð0; IÞ: ð27Þ

In the above equations, PeF ;PeS, and Bi denote the dimensionlessPeclet numbers in the fluid and solid, respectively, and Biot number,given by

PeF ¼LUcF

pqkF

; PeS ¼LUcS

pqS

kS; Bi ¼ aFL

kF:

3. Numerical scheme

We start with the weak form of the Navier–Stokes equationsand derive a weak one-field formulation of the energy equation.Next, we describe the arbitrary Langrangian–Eulerian approachto handle the equations on time-dependent domains. Then the dis-cretization in space and time will be given. Finally in this section,we discuss aspects of the mesh handling at the time-dependentliquid–solid interface and within the domains XFðtÞ and XS.

S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687 673

3.1. Weak form of the Navier–Stokes equations

We use the standard notations L2;H1 and ð�; �Þx for Sobolevspaces and the inner product in L2ðxÞ and its vector-valued ver-sions, respectively. We look for velocity and pressure in the spaces

V :¼ fv 2 H1ðXFðtÞÞ3 : v � mS ¼ 0 on CSðtÞg; ð28ÞQ :¼ L2ðXFðtÞÞ; ð29Þ

thus the no penetration boundary condition u � mS ¼ 0 on liquid–solid interface CSðtÞwill be implemented in both the ansatz and testvelocity spaces. Multiplying the momentum and mass balanceequations. (13) and (14) by test functions v 2 V and q 2 Q respec-tively, integrating over XFðtÞ and applying the Gaussian theoremfor the part including the stress tensor lead us to the weak formu-lation. In particular, the term including the symmetric stress tensorcan be written as

�Z

XF

r�Sðu;pÞ �vdxþZ@XF

v �Sðu;pÞ �mdc

¼Z

XF

Sðu;pÞ :rvdx¼Z

XF

12

Sðu;pÞ :rvdxþZ

XF

12

STðu;pÞ :rvdx

¼ 2Re

ZXF

DðuÞ : DðvÞdx�Z

XF

pr�vdx: ð30Þ

Now, the free surface and liquid–solid boundary conditions will beincluded in the weak form as follows. We move the boundary inte-gral term in (30) to the right hand side and split it into integralsover CS and CF

�Z@XF

v � Sðu;pÞ � mdc ¼ �Z

CS

v � Sðu; pÞ � mSdcS

�Z

CF

v � Sðu; pÞ � mF dcF : ð31Þ

Then, using the orthonormal decomposition

v ¼ ðv � mSÞmS þX2

i¼1

ðv � si;SÞsi;S on CS;

the no penetration condition (15) incorporated in (28) and the slipwith friction boundary condition (16), the integral over the liquid–solid interface can be written as

�Z

CS

v � Sðu;pÞ � mSdcS ¼ �X2

i¼1

ZCS

ðv � si;SÞsi;S � Sðu; pÞ � mSdcS

¼ 1b�

X2

i¼1

ZCS

ðu � si;SÞðv � si;SÞdcS: ð32Þ

Similarly, using the orthonormal decomposition

v ¼ ðv � mFÞmF þX2

i¼1

ðv � si;FÞsi;F on CF

and the boundary conditions (18) and (19), the integral over theliquid–gas interface becomes

�Z

CF

v � Sðu; pÞ � mF dcF ¼ �Z

CF

X2

i¼1

ðv � si;FÞsi;F � Sðu;pÞ

� mFdcF �Z

CF

ðv � mFÞ � mF � Sðu;pÞ

� mFdcF

¼ C0

We

ZCF

X2

i¼1

ðv � si;FÞðsi;F � rTFÞdcF

� 1We

ZCF

MTF ðv � mFÞKdcF ; ð33Þ

where MTF ¼ ð1� C0ðTF � 1ÞÞ. The surface tension force given by thesecond term in (33) contains the temperature dependent surfacetension coefficient and the curvature. Hence, the approximationsof the temperature and the curvature have to be very accurate inorder to avoid unphysical flows in the droplet. One of the populartechniques is based on the replacement of the curvature vectorKmF by the Laplace–Beltrami operator applied to the identity idCF

followed by an integration by parts [36,37,48]. The main advantagesof this technique are

- the surface tension force can be computed for piecewise smoothsurfaces,

- only first derivatives of the basis functions are needed,- the surface tension force can be treated semi-implicitly giving

additional stability,- the contact angle at the moving contact line can be incorpo-

rated directly.

In order to explain this technique in more detail we need some factsfrom differential geometry. Let U � R3 be an open set with CF � U.For a scalar function f given on U, we define the components of thetangential gradient by

ðrf Þi :¼ dif ¼ @if � ðmF � rf ÞmF;i; i ¼ 1;2;3:

It turns out that the restriction ofrf onto CFðtÞ depends only on thevalues of f on CFðtÞ. The Laplace–Beltrami operator applied to f isdefined as

Df :¼ r � ðrf Þ ¼X3

i¼1

diðdif Þ: ð34Þ

Its application to vector-valued function is understood in a compo-nent-wise manner. We replace the curvature vector KmF in (33) bythe Laplace–Beltrami operator applied to the identity DidCF andintegrate by parts [49,50] to get

� 1We

ZCF

MTF v � mFKdcF ¼ �1

We

ZCF

MTF DidCF � v dcF

¼ 1We

ZCF

ridCF

: rðMTF vÞdcF �1

We

Zfðmf � ridCF Þ

�MTF vdf

¼ 1We

ZCF

ridCF

: ½MTFrv þ v �rMTF �dcF �1

We

�Z

fmf �MTF vdf: ð35Þ

Here, we used mf � mF ¼ 0 such that mf � ridCF ¼ mf � ridCF ¼ mf. Note,that the last term in (35), which does not appear for closed surfacesCF , is an integral over the contact line f and mf is the co-normal vectorat the contact line which is normal to f and tangent to CF . Let sS be thescaled projection of mF onto the plane CS such that jsSj ¼ 1 that means

sS :¼ mF � ðmF � mSÞmS

jmF � ðmF � mSÞmSj:

Then, mf � sS ¼ cosðhÞwhere h is the contact angle at the moving con-tact line. In order to incorporate the contact angle into the model,we decompose the test function v into

v ¼ ðv � mSÞmS þ ðv � sSÞsS þ ðv � s?S Þs?S ;

where s?S is a unit vector perpendicular to mS and sS. Now, computingthe inner product mf � v in the last term of (35) we use v � mS ¼ 0 on CS

for all v 2 V and the orthogonality mf � s?S ¼ 0 of the co-normal to get

674 S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687

Zfmf �MTF vdf ¼

Zfðmf � sSÞðv � sSÞMTF df

¼Z

fcosðhÞMTF v � sSdf: ð36Þ

Using the derivations (30)–(33), (35) and (36) the weak form ofthe Navier–Stokes equation reads:

For given uð0Þ, find ðu; pÞ 2 V � Q such that

@u@t;v

� �XF

þ aðu; u;vÞ � bðp;vÞ þ bðq;uÞ ¼ f ðK; vÞ ð37Þ

for all ðv; qÞ 2 V � Q . Here, the forms a; b and f are given by

aðu; u;vÞ ¼Z

XF

2Re

DðuÞ : DðvÞ þ ðu � rÞu � vdx

þ 1b�

ZCS

X2

i¼1

ðu � si;SÞðv � si;SÞdcS;

bðq;vÞ ¼Z

XF

qr � vdx;

f ðK; vÞ ¼ 1Fr

ZXF

e � vdxþ 1We

Zf

cosðhÞMTF v � sSdf

� 1We

ZCF

ridCF : ½MTFrv þ v �rMTF �dcF

� C0

We

ZCF

X2

i¼1

ðv � si;FÞðsi;F � rTFÞdcF :

3.2. Weak form of the energy equation

We start with multiplying (20) with a test function wF 2 H1ðXFÞ,integrating by parts, incorporating the boundary condition (24)and using the transition condition (25) to getZ

XF

@TF

@twF dxþ

ZXF

ðu � rÞTFwF dxþ 1PeF

ZXF

rTF � rwF dx

þZ

CF

BiPeF

TFwF dcF

¼ � 1PeF

kS

kF

ZCS

@TS

@mSwFdcF : ð38Þ

Similarly, we multiply (21) with a test function wS 2 H1ðXSÞ, inte-grate by parts and incorporate the boundary condition (27) toobtainZ

XS

@TS

@twSdxþ 1

PeS

ZXS

rTS � rwSdx ¼ 1PeS

ZCS

@TS

@mSwSdcS: ð39Þ

In order to write the energy equations (38) and (39) in a one-fieldformulation in the domain XðtÞ :¼ XFðtÞ [XS [ CSðtÞ, we define

uTðx; tÞ ¼uðx; tÞ if x 2 XFðtÞ;0 if x 2 XS;

�Tðx; tÞ ¼

TFðx; tÞ if x 2 XFðtÞ;TSðx; tÞ if x 2 XS;

�PeðxÞ ¼

PeF if x 2 XFðtÞ;kFkS

PeF if x 2 XS;

(

T0ðxÞ ¼TF;0ðxÞ if x 2 XFðtÞ;TS;0ðxÞ if x 2 XS:

�Moreover, let gðxÞ ¼ 1 in XFðtÞ and gðxÞ ¼ kSPeS=ðkFPeFÞ in XS. Then,multiplying (39) with kSPeS=ðkFPeFÞ and adding to (38) we end upwith the weak one-field formulation of the energy equation:

For given Xð0Þ;uT and T0, find T 2 H1ðXÞ such that for allw 2 H1ðXÞ

g@T@t;w

� �X

þ aTðuT ; T;wÞ þ bTðT;wÞ ¼ 0: ð40Þ

The forms aT and bT are given by

aTðuT ; T;wÞ ¼Z

X

1PerT � rwdxþ

ZXðu � rÞTwdx; bTðT;wÞ

¼ BiZ

CF

1Pe

Twdx:

3.3. Arbitrary Lagrangian Eulerian (ALE) approach

We consider the impinging droplet problem on a micro scaleand exclude topological changes in XFðtÞ. Thus, the ArbitraryLagrangian–Eulerian (ALE) approach is an appropriate tool to han-dle the free surface of the droplet. In the ALE approach, the freesurface is resolved by the computational mesh. It allows furtherto suppress spurious velocities in free surface flows [35].

For each t 2 ð0; IÞ, we define a family of ALE mappings

At : bXF ! XFðtÞ; AtðYÞ ¼ XðY; tÞ; Y 2 bXF ;

where bXF is a reference domain of XFðtÞ. In practical computations,the initial domain or the previous time-step domain is often takenas the reference domain. We assume that the ALE mapping At is anhomeomorphism, i.e. continuous with continuous inverse A�1

t . Fur-ther, we assume that the mapping

t ! XðY; tÞ; Y 2 bXF

is differentiable almost everywhere in the time interval ð0; IÞ. Forany scalar function v 2 C0ðXFðtÞÞ, we define their correspondingfunction v 2 C0ðbXFÞ as follows:

v :¼ v � At ; with vðY; tÞ ¼ vðAtðYÞ; tÞ:

The time derivative of v 2 C0ðXFÞ on the reference frame

@v@t

����bXF

: XFðtÞ � ð0; IÞ ! R

is defined as

@v@t

����bXF

ðX; tÞ ¼ @v@tðY; tÞ; Y ¼ A�1

t ðXÞ:

We define the domain velocity w as

wðX; tÞ ¼ @X@t

����bXF

:

Now, applying the chain rule to the time derivative of v � At on thereference frame we get

@v@t

����bXF

¼ @v@t

����Xþ @X@t

����bXF

� rXv ¼ @v@t

����Xþw � rXv: ð41Þ

Thus, to rewrite any equation into the non-conservative ALE form,the time derivative has to be replaced with the time derivative onthe reference frame and a convective domain velocity term has tobe added. Now, using (41) the variational form of the Navier–Stokesequation (37) in the ALE frame reads:

For given Xð0Þ;uð0Þ, find ðu; p;wÞ 2 V � Q � V such that for allðv; qÞ 2 V � Q

@u@t

� ����bXF

;v

!XF

þ aðu�w; u;vÞ � bðp;vÞ þ bðq;uÞ ¼ f ðK; vÞ

Now, to rewrite the energy equation in an ALE forms, we need anextension of the ALE mapping on bXF onto bX with the associateddomain velocity wT where wT jXF

¼ w. Here, bX is a reference domainof XðtÞ. The variational form of the energy equation (40) in the ALEframe reads:

S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687 675

For given uT ;wT ; Tð0Þ, find T 2 H1ðXÞ such that for all w 2 H1ðXÞ

g@T@t

� ����bX ;w�

X

þ aTðuT �wT ; T;wÞ þ bTðT;wÞ ¼ 0:

Fig. 2. Mesh adaptive method.

3.4. 3D-axisymmetric formulation

The computational cost can be drastically reduced when werestrict to a 3D-axisymmetric model. There is no need to start againwith the partial differential equation in cylindrical coordinates andderive a suitable variational formulation. Instead we generate the3D-axisymmetric weak form in the 2D-meridian domain directlyfrom the weak form in 3D-Cartesian coordinates. Introducing cylin-drical coordinates, the volume, surface and line integrals in 3D aretransformed into area integrals, line integrals and a functional inthe contact point as described for interface flows in [21]. In thisway the space dimension is reduced by one. Note that this approachleads naturally to boundary conditions along the ’artificial rota-tional axis’ which are already partially included in the weak form.

3.5. Discretization in space and time

We use the fractional-step-H scheme as time-discretization[51, Section 3.2] which is strongly A-stable and second order accu-rate on fixed domains. One step in the fractional-step-H schemeconsists of a clever combination of three H schemes with differentstep sizes and different H. In order not to overload the representa-tion, in the following we will denote by ðtn; tnþ1Þ a subinterval inthe fractional-step-H scheme. We advect the free surface andinterface points Xn on CFðtnÞ [ CSðtnÞ solving

dXdt¼ uðX; tÞ

with the implicit Euler scheme

Xnþ1 ¼ Xn þ ðtnþ1 � tnÞunþ1;

to obtain the new position of the free surface and interface points,respectively. More details on the mesh handling are given in thenext two subsections.

Investigations in [52] show that an implicit handling of the cur-vature term seem to be needed to hope for unconditional stability.Thus, as in [48,53], we use a semi-implicit approximation of thecurvature term (first term in (35))

1We

ZCF ðtnþ1Þ

MTFridCF ðtnþ1Þ : rvdcF

1We

ZCF ðtnÞ

MTF ridCF ðtnÞ þ dtnunþ1� �

: rvdcF ;

with the time step dtn ¼ tnþ1 � tn. Consequently, the curvature termis splitted into an explicit term on the right hand side of the weakformulation (37)

� 1We

ZCF ðtnÞ

MTnFridCF ðtnÞ : rvdcF ;

where TnF :¼ TFðtnÞ, and an implicit term on the left hand side

dtn

We

ZCF ðtnÞ

MTnFrunþ1 : rvdcF :

Note that this term on the left hand side is symmetric and positivesemi-definite and – compared with a fully explicit approach –improves the stability of the discrete system.

Since in the ALE approach the computed velocities are used tomove the surface and interface points, high accuracy is required.

Further, mass should be conserved and spurious velocities – ifthere are some – should be suppressed. Therefore, we usesecond–order inf–sup stable finite element approximationsavoiding spurious pressure oscillations [54]. Continuous pressureapproximations often generate spurious velocities. In order tosuppress them discontinuous pressure approximations are a goodchoice [35]. But to guarantee the inf–sup stability for discontinu-ous, piecewise linear pressure approximations on arbitrary shaperegular families of meshes we have to enrich the space ofcontinuous, piecewise quadratic functions by cubic bubblefunctions. We denote this pair shortly as ðPþ2 ; P

disc1 Þ and use it for

the approximation of the velocity components and the pressure.The good mass conservation properties are lying on the fact thatup to the first integral moments the divergence is elementwisevanishing. The temperature is approximated in X by the standardcontinuous, piecewise quadratic P2 element.

3.6. Mesh handling at the liquid–solid interface

In the droplet deformation problem, the free surface verticesadjacent to the contact line may reach the solid surface in the nexttime step during the droplet spreading, see Fig. 2. Further, the ver-tices on the liquid–solid interface can move in the tangential direc-tion due to the slip with friction boundary condition. This leads tonon-matching grid along the liquid–solid interface if the mesh inthe solid domain would be fixed. The liquid–solid interface canbe handled by either a non-matching or a matching grid. In thenon-matching grid, it is enough to find the displacement of theliquid domain and the domain velocity in the solid phase will be

676 S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687

zero. In the matching grid, the displacement of the solid domainalso has to be calculated due to the movement of the vertices onthe liquid–solid interface. Thus, the domain velocity will benon-zero in solid phase which will induce the domain velocity con-vection term in the energy equation of the solid phase. In order tohandle the non-matching grid at the liquid–solid interface, the mor-tar finite element method could be used. However, then we wouldget mortar spaces of time-dependent dimension which are not sim-ple to realize, whereas no special method is needed in the case of amatching grid. That is why we prefer to use matching grids.

Let us explain in more detail the implemented mesh manipula-tion technique, which maintains a matching grid when a free sur-face vertex reaches the solid surface. Every time step, after movingthe mesh, we calculate the distance between the solid surface andthe free surface vertex adjacent to the solid surface. If the distanceis less than 10�8 or if the free surface vertex is penetrated into thesolid surface, we do the following steps to maintain a matchinggrid before continuing with the next time step:

- Move the free surface vertex perpendicularly to the solid sur-face and change its type to a contact line (point) vertex. Here,we could move the free surface vertex using the fluid velocitybut the vertex is already very close (10�8) to the solid surfaceand for simplicity we move it perpendicularly to the solid sur-face, see Fig. 2, top and middle.

- Delete the free surface edge and the solid edge on which thefree surface vertex incident, see Fig. 2, middle.

- Reconnect all edges associated with the right side vertex of thesolid edge to the moving contact line vertex.

- Generate a new liquid–solid interface edge with the old andnew contact line vertex, see Fig. 2, bottom.

- Construct new finite element spaces for the velocity, pressureand temperature on the manipulated mesh.

- Map the nodal values of old finite element functions on the nodalvalues of the new finite element functions without interpolation.

- Impose the no penetration condition by setting the normalcomponent of the velocity to zero for the degrees of freedomassociated with the newly generated liquid–solid interfaceedge.

- Take the average value from the free surface temperature val-ues and the solid surface temperature values for the tempera-ture degrees of freedom on the newly generated liquid–solidinterface edge.

Due to the above mesh manipulation approach, the mesh at theliquid–solid interface remains fitted and thus no mortar (or other)technique is needed. But, due to the tangential movement of theliquid–solid vertices, the displacement of inner points in the solidphase has to be recalculated and this induces a domain velocityconvection term in the energy equation of the solid phase. Note,that the mesh manipulation also guarantees a fixed number of cellsbut only the number of degrees of freedom along the liquid–solidinterface and along the free surface may change.

3.7. Linear elastic mesh update

In the elastic mesh update, we calculate the displacement vec-tors subject to the displacement of the boundary in each time step.The displacement vector Wnþ1 in the domains bXFðtnÞ and bXS areobtained by solving linear elasticity equations. For a given bound-ary displacement !nþ1, find the displacement Wnþ1such that

r � TðWnþ1Þ ¼ 0 in bXðtnÞ [ bXS

Wnþ1 ¼ !nþ1 on CFðtnÞ [ CSðtnÞ [ CNðtnÞð42Þ

where

Tð/Þ ¼ k1ðr � /ÞIþ 2k2Dð/Þ:

Here, k1 and k2 are Lame constants. To solve problem (42), we usethe same triangulation T h of the domain which has been takenfor the flow and temperature variables and apply a finite elementmethod. The discrete form of Eq. (42) is derived in the usual way:

Find Wnþ1h 2Wn

h :¼ H1ðbXðtnÞ [ bXSÞ such that Wnh ¼ !n

h onCFðtnÞ [ CSðtnÞ [ CNðtnÞ and for all / 2Wn

h;0 the relationZbXhðtnÞ

DðWnhÞ : Dð/Þ þ C1

ZbXhðtnÞ

r �Wnhr � / ¼ 0 ð43Þ

holds. Here C1 ¼ k1=k2 is a positive constant, which in all our calcu-lations equals 1. The solution of (43) is approximated by continu-ous, piecewise linear P1 triangular finite elements since we haveto find the displacements of the vertices only and want to havelow costs. Also, note that we used the previous time step domainas the reference domain to calculate the mesh velocity. We referto [55] for a priori bounds and the regularity requirements of map-pings in (43).

3.8. Calculating the heat flux across the liquid–solid interface

One of the main objectives of this modeling is to study the heattransfer from the solid phase into the liquid phase during thespreading and recoiling of the droplet for better understanding ofthe cooling effects. Thus, the crucial parameter in this study isthe heat flux across the liquid–solid interface CS from the solidphase into the liquid phase. The heat flux (Q) quantifies the coolingeffect in the solid phase, and one of the main objective in spraycooling is to maximize Q. Therefore, the heat flux has to be calcu-lated very accurately. In this section, we present an accurate tech-nique to calculate the heat flux based on the ideas given in [56] fora stationary domain. In the time interval ðtn; tnþ1Þ the heat flux overCS is given by

Q :¼Z tnþ1

tn

ZCS

1PeS

@TS

@mSdcS ¼

Z tnþ1

tn

Z@XS

1PeS

@TS

@mSdcS; ð44Þ

since @TS@n

��@XSnCS

= 0, see (11). Let us consider the Dirichlet problemwritten in ALE frame

@TS

@t

����bXS

�w � rTS �r �1

PeSrTS

� �¼ 0 in XS � ðtn; tnþ1Þ;

TSðX; tnÞ ¼ TnjXSin XS;

TS ¼ Tj@XSon @XS;

ð45Þ

where TS on @XS is known as the trace of T. Now, multiplying (45)by v 2 H1ðXSÞ and integrating by parts we get

@TS

@t

� ����bXS

�w � rTS; v!

XS

þ 1PeSrTS;rvð ÞXS

¼ 1PeS

Z@XS

@TS

@mSvdc:

Since the left hand side is equal to zero for all test functionsv 2 H1

0ðXSÞ (which is just the weak formulation of the Dirichletproblem (45)), the quantity of interest (right hand side for v ¼ 1on @XS) is independent of the special choice of

v 2 fw 2 H1ðXSÞ : w ¼ 1 on @XSg:

Now integrating over one substep ðtn; tnþ1Þ of the fractional-step-Hscheme, we get for all v 2 H1ðXSÞ with v ¼ 1 on @XS

Q :¼ 1PeS

Z tnþ1

tn

Z@XS

@TS

@mSdcdt

¼ Tnþ1S � Tn

S ; v�

XS

þZ tnþ1

tn

1PeSðrTS;rvÞXS

� ðw � rTS;vÞXS

�dt:

In the discrete case, we define

S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687 677

Q h : ¼ Tnþ1S;h � Tn

S;h; vh

� XS

þHdtn1

PeSrTnþ1

S;h ;rvh

� XS

� wnþ1 � rTnþ1

S;h ;vh

� XS

�þ ð1�HÞdtn

1PeS

rTnS;h;rvh

� XS

� wn � rTn

S;h; vh

� XS

�;

where H is the corresponding value from the substep in the frac-tional-step-H scheme. Note that, as in the continuous case, Qh isindependent from the special choice of the finite element functionvh 2 H1ðXSÞwith vh ¼ 1 on @XS. Let u1;u2; . . . ;uN be the basis func-tion associated with the inner nodes of XS and indices from N þ 1 toN þM be the basis functions associated with the boundary @XS

nodes. Then, with vh as above and

TS;h ¼XNþM

j¼1

Tjuj vh ¼XNþM

i¼Nþ1

ui:

Note that vh 1 at all boundary nodes on @XS and vh 0 at allinner nodes in XS. Thus, we get

Q h ¼XNþM

i¼Nþ1

XNþM

j¼1

ðTnþ1j � Tn

j Þðuj;uiÞXSþHdtn Tnþ1

j1

PeSðruj;ruiÞXS

��ðwnþ1 � ruj;uiÞXS

iþ ð1�HÞdtn Tn

j1

PeSðruj;ruiÞXS

� wn � ruj;ui

� XS

��:

Therefore, to calculate the heat flux, we need only the mass and thestiffness matrices, but these matrices have been generated alreadyin the assembling process. From the generated matricesMij ¼ ðuj;uiÞXS

and

A‘ij ¼

1PeSðruj;ruiÞXS

� w‘ � ruj;ui

� XS

for ‘ 2 fn;nþ 1g;

we compute the vectors

mðjÞ ¼ uj;XNþM

i¼Nþ1

ui

!XS

¼XNþM

i¼Nþ1

ðuj;uiÞXSa‘ðjÞ

¼XNþM

i¼Nþ1

1PeSðruj;ruiÞXS

� w‘ � ruj;ui

� XS

�:

More precisely, we first initialize the vectors mðjÞ ¼ 0 and a‘ðjÞ ¼ 0for j ¼ 1; . . . ;N þM. Then, for each nodal value Tj on @XS, that is, foreach Tj; j ¼ N þ 1; . . . ;N þM, we identify the jth column in the matri-ces M and A‘. After that we set

mðjÞ ¼XNþM

i¼Nþ1

Mi;j; a‘ðjÞ ¼XNþM

i¼Nþ1

A‘i;j:

Note that these vectors have to be calculated before doing anymanipulation for imposing boundary conditions. Finally, we com-pute the heat flux as

Q h ¼ ðTnþ1 � TnÞ �mþHdtnTnþ1 � anþ1 þ ð1�HÞdtnTn � an: ð46Þ

Note that the basis functions on the liquid–solid interface have sup-port on both liquid and solid phases. However, to evaluate Qh weneed the contribution only from one side, and we always considerthe support of the basis functions from the solid phase in ourcalculations.

4. Numerical results

In this section, we present an array of numerical results for a3D-axisymmetric non-isothermal liquid droplet impinging on a

horizontal hot solid substrate. Simulations of isothermal impingingdroplets have been presented in [21,38,39]. To demonstrate theaccuracy of the proposed numerical model, we first compare thesimulated results with experimental data given in [57,58], and thenthe mesh convergence study is performed. After that, we study theinfluence of the Reynolds number on the flow dynamics of thedeforming non-isothermal droplet on a hot substrate. Finally, werun an array of computations to study the thermal effects on the flowdynamics of the droplet by varying the Weber, Peclet, and Biot num-bers. Unless specified, we used a constant time step 0:0005; b� ¼ 1and C0 ¼ 0:002 in all computations. The total number of degrees offreedom of an impinging droplet problem varies during the reme-shing, however we have around 14,000 degrees of freedom initiallyin all test cases with h ¼ 0:01557859. Further, a typical simulation ofimpinging droplet takes approximately nine hours for computinguntil the dimensionless time T ¼ 12:5 using the Intel(R) Core(TM)i7–2600 CPU at 3.40 GHz with 8 GB memory.

4.1. Validation

We first compare the numerically obtained wetting diameterand the apex height of an impinging water droplet with the exper-imental results given in [57]. We consider a water droplet of diam-eter d0 ¼ 1:29� 10�3 m impinging on a polished silicon surfacewith the pre-impact speed of uimp ¼ 1:18 m/s, polished test casein Table 1 of [57]. We consider the non-heated case with the equi-librium contact angle he ¼ 46�. Using L ¼ d0 and U = 1.18 m/s ascharacteristic values, we get Re ¼ 1522;We ¼ 25 and Fr ¼ 110.Computations are performed with b� ¼ 150, and the initial meshsize is h ¼ 0:01557859. The computationally obtained wettingdiameter and the apex height of the droplet are compared inFig. 3 with the experimental results. The droplet attains a maxi-mum dimensionless wetting diameter of 2.1728, and then startsto recoil before attaining the equilibrium state. The wetting diam-eter fits very well, both qualitatively and quantitatively, with theexperimental result even in the recoiling stage. In experiments,an oscillating behavior in the apex height of the droplet has beenobserved, and it could be due to the capillary waves on the freesurface. The numerically obtained apex height curve is moredamped, even though the qualitative behavior is similar toexperiments. The damping effect has also been observed in the Vol-ume–of–Fluid simulations of 3D–axisymmetric droplets with adynamic contact angle model, see Fig. 10 in [28]. The axisymmetricassumption in the numerical model could be the reason for thisdamping effect, and thus a full 3D simulation has to be used in orderto capture the surface capillary waves more accurately. To supportour argument, we consider another experiment reported in [58]with no capillary waves on the free surface. A water droplet of diam-eter d0 ¼ 2:7� 10�3 m with We ¼ 90 and he ¼ 100� is considered.The numerically obtained wetting diameter and the apex height ofthe droplet are depicted in Fig. 4. The obtained wetting diameterand the apex height are in very good agreement, both qualitativelyand quantitatively during the spreading and recoiling stages, withthe experimental results reported in Fig. 6 and 7 of [58].

We next perform the mesh convergence study for the proposednumerical scheme. Consider a hemispherical water droplet ofdiameter d0 ¼ 1:29� 10�3 m with the equilibrium contact anglehe ¼ 80� on a cylindrical substrate. We assume that the droplet isin rest initially. Further, the height and diameter of the solid cylin-der are 5� 10�4 m and 3:8� 10�2 m respectively. Using L ¼ d0 andU = 1.18 m/s as characteristic values, we get Re ¼ 1522;We ¼ 25and Fr ¼ 110. We run an array of simulations with a fixed surfacetension by assuming TF ¼ TS ¼ 0 K to study the mesh convergenceof the flow solver. The initial mesh (Level 0) consists of 25 verticeson the free surface with h ¼ 0:06282152, and the successivemesh levels are generated by uniformly refining the initial mesh.

0 5 10 15 200

1

2

2.5

time

Wet

ting

diam

eter

experimental numerical

0 5 10 15 200

0.2

0.4

0.6

0.8

1

time

apex

hei

ght

experimental numerical

(a)

(b)

Fig. 3. Comparison with experimental results. (a) wetting diameter, and (b) apexheight of a liquid droplet with he ¼ 46� , Re = 1522, We = 25 and Fr = 110.

0.01 0.1 1 50

0.5

1

1.5

2

2.5

3

time

Wet

ting

diam

eter

experimental numerical

0.01 1 50

0.2

0.4

0.6

0.8

1

time

apex

hei

ght

experimental numerical

(a)

(b)

Fig. 4. Comparison with experimental results. (a) wetting diameter, and (b) apexheight of a water droplet with he ¼ 100� , We = 90.

678 S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687

All computations are performed with the slip coefficientb� ¼ 0:55=h, where h is the mesh size, and with the time step0.0005 (Level 0 to 3) and 0.0001 (Level 4), respectively.

Since the initially contact angle (90� due to hemisphericalshape) and wetting diameter are not in equilibrium, the dropletslowly starts to oscillate and attains it equilibrium value. Theobtained dynamics of the wetting diameter and the dynamic con-tact angle in all five mesh levels look similar in Fig. 5(a) and (c). Aclose-up look, however, at the maximal wetting diameter valueFig. 5(b) clearly shows the convergence behavior. A similar meshconvergence behavior has been observed in the dynamic contactangle, see Fig. 5(c) and (d).

We finally perform the mesh convergence study on the heattransfer using the same example with different initial temperaturein the solid substrate. In this study we consider the first three meshlevels with the time step 0.0005 and b� ¼ 0:55=h. Further, we usedPeF ¼ PeS ¼ 1. The obtained heat flux and the total heat flux overtime in different mesh levels are presented in Fig. 6. The conver-gence (in space) of the heat transfer are clearly seen in Fig. 6.

4.2. Influence of the Reynolds number on the flow dynamics

To study the influence of the Reynolds number on the flowdynamics of a non-isothermal water droplet impinging on a hot solidsubstrate, we consider a droplet of diameter d0 ¼ 2:7� 10�3 m. Fur-ther, the following material parameters are used: the densityq = 1000 kg/m3, the dynamic viscosity l = 0.001 Ns/m2, the surfacetension r = 0.073 N/m, the equilibrium contact angle he ¼ 30�. Also,we assume that uimp = 1.54 m/s, TF;0 = 298 K, TS;0 = 328 K, T1 = 298 K,Tref = 323 K (temperature of the fluid surface, the solid surface, theatmospheric and the reference, respectively). Using L = d0=2 andU = uimp as characteristic values, we get Re ¼ 2079;We ¼ 43; Fr

¼ 179;PeF ¼ 10;PeS ¼ 100;Bi ¼ 0:000058 for the considered mate-rial parameters. In order to study the influence of the Reynolds num-ber, we considered four variants: (i) Re = 260, (ii) Re ¼ 520, (iii)Re ¼ 1040 and (iv) Re ¼ 2079 by varying the viscosity l.

We first present the contours of the pressure, velocity, vorticity,and the temperature distribution for the high Reynolds numbercase Var (iv). Fig. 7 depicts the pressure contours in the dropletof Var (iv) at different instances (dimensionless time)t = 0:1;1:0;2:0;8:0;16;25. Initially, the pressure variation is largenear the contact line. However, the distribution of pressurebecomes almost uniform toward the end of simulation. Next, thecontour lines of the magnitude of the velocity for the same variantare depicted in Fig. 8. The arrows in each picture represent the flowdirection. The vorticity contours in the 3D-axisymmetric imping-ing droplet during the sequence of spreading and recoiling aredepicted in Fig. 9. The observed range of vorticity is high in thehigh Reynolds number variant in comparison with the low Rey-nolds number Var (i).

Next, the computationally obtained sequence of droplets in Var(i) and Var (iv) at different instances (dimensionless time)t = 1:0;8:0;16;25, are depicted in Figs. 10 and 11 with the solidsubstrate. The colors in the pictures represent the temperature dis-tribution during the droplet deformation over time. Initially, thedimensionless temperature in the droplet is zero, whereas the tem-perature in the solid is 1.2143. In the high Reynolds number case,Var (iv), the wetting diameter of the droplet is larger, and thus thedroplet covers more solid substrate and absorb more heat from thesolid substrate. The total heat flux from the solid phase into theliquid phase will be studies in the Sections 4.4 and 4.5.

Next, the obtained wetting diameter and the dynamic contactangle in all four variants over time are presented in Fig. 12. Asexpected, the wetting diameter increases when the Reynoldsnumber increase. The maximum wetting diameter and the time

0 5 101

1.05

1.1

1.15

time

Wet

ting

diam

eter

Level 0Level 1Level 2Level 3Level 4

1.9 2.1 2.3

1.115

1.117

1.119

time

Wet

ting

diam

eter

Level 0Level 1Level 2Level 3Level 4

0 5 1075

80

85

90

time

Dyn

amic

con

tact

ang

le

Level 0Level 1Level 2Level 3Level 4

10 11 12

78

79

80

time

Dyn

amic

con

tact

ang

le

Level 0Level 1Level 2Level 3Level 4

(a) (b)

(c) (d)

Fig. 5. Mesh convergence study on the flow dynamics of the hemispherical water droplet with Re = 1522, We = 25 and Fr = 110 and he = 80�. The obtained dynamics of thewetting diameter (a) and their close-up at the maximal value (b), the dynamics contact angle (c) and their close-up at the end of the equilibrium stage in five mesh levels.

0.01 1−20

−15

−10

−5

0

5 x 10−4

time

Hea

t flu

x

Level 0Level 1Level 2

4 8 12−0.8

−0.7

−0.6

time

Tot

al h

eat f

lux

Level 0Level 1Level 2

(a)

(b)

Fig. 6. Mesh convergence study on the heat transfer in the hemispherical waterdroplet with Re = 1522, We = 25 and Fr = 110 and he ¼ 80� . The obtained heat flux(a) and the total heat flux (b) over time in different mesh levels.

S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687 679

taken to obtain it are presented in Fig. 13. We can observe that themaximum wetting diameter is nonlinear with respect to the Rey-nolds number, see Fig. 13. A similar behavior is observed in thetime taken to attain the maximum wetting diameter. The dynamiccontact angle in all four variants reach the equilibrium valuealmost after the dimensionless time t = 10, see Fig. 12. Further,the maximum relative mass loss in the liquid droplet occurs in

Var (iv), and it is 0.42% which shows the accuracy of the numericalscheme.

Remark 1. Since the contact point and the free surface pointsmove with the discrete fluid velocity, a small change in the relativevelocity between the contact point and the free surface point nextto it will induce wiggles in the dynamics contact angle calculations.Note that the free surface point of the droplet may reach the solidsubstrate (rolling motion) during spreading. These are the reasonsfor the wiggles in the curves of Fig. 12. Also, note that in ourcalculations, the angle between the solid substrate and the firstfree surface point is calculated as the dynamic contact angle atevery time-step.

4.3. Influence of the Weber number on the flow dynamics

To study the influence of the surface tension on the flowdynamics, we vary the Weber number and fix the other dimension-less numbers. We consider the following four variants: (i)We ¼ 43, (ii) We ¼ 90, (iii) We ¼ 140, and (iv) We ¼ 190 by vary-ing the surface tension. Further, we used the fixed numbersRe ¼ 2079; Fr ¼ 179, PeF ¼ 10;PeS ¼ 100, and Bi ¼ 0:000058 in allvariants. Also, we set the equilibrium contact anglehe ¼ 30�; TF;0 ¼ 298 K; TS;0 ¼ 328 K, T1 = 298 K, Tref = 323 K.

The computationally obtained wetting diameter and dynamiccontact angle over time for all four variants are presented inFig. 14. Since the surface tension will be large in the Var (i), themaximum wetting diameter is small. As expected, the maximumwetting diameter increases when the Weber number increase.The dynamic contact angle in all four variants reach the equilib-rium value almost after the dimensionless time t = 1. However,more oscillations have been observed in Var (iv). Since the surfacetension coefficient is very small (less rigid) in this variant, morerolling motion is observed while spreading. Next, the maximumwetting diameter and the time taken to obtain it are presented inFig. 15. As in the Reynolds number case, the maximum wettingdiameter is nonlinear with respect to the Weber number, see

Fig. 7. Pressure contours in the impinging droplet with he ¼ 30� , We = 43, Fr = 179 during spreading and recoiling. Var (iv) Re = 2079. The dimensionless timing from the top:t ¼ 0:1;1:0;2:0;8:0;16;25.

680 S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687

Fig. 8. Velocity (magnitude) contours in the impinging droplet with he ¼ 30�;We ¼ 43; Fr ¼ 179 during spreading and recoiling. Var (iv) Re = 2079. The dimensionless timingfrom the top: t ¼ 0:1;1:0;2:0;8:0;16;25.

S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687 681

Fig. 9. Vorticity contours in the impinging droplet with he ¼ 30� ;We ¼ 43; Fr ¼ 179 during spreading and recoiling. Var (iv) Re=2079. The dimensionless timing from the top:t ¼ 0:1;1:0;2:0;8:0;16;25.

682 S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687

Fig. 10. Sequence of images of a liquid droplet impinging on a hot solid substratewith he ¼ 30�;We ¼ 43; Fr ¼ 179. Var (i) Re = 260. The dimensionless timing fromthe top: t ¼ 1:0;8:0;16;25.

Fig. 11. Sequence of images of a liquid droplet impinging on a hot solid substratewith he ¼ 30�;We ¼ 43; Fr ¼ 179. Var (iv) Re = 2079. The dimensionless timing fromthe top: t ¼ 1:0;8:0;16;25.

S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687 683

Fig. 15. Further, the maximum relative mass loss in the liquid drop-let is 0.4% and it occurred in Var (iv).

4.4. Influence of the solid Peclet number on the flow dynamics

In this section we study the influence of the solid Peclet number onthe heat transfer. We consider five variants (i) PeS = 50, (ii) PeS = 100,(iii) PeS = 250, (iv) PeS = 350 and (v) PeS = 400 by varying the heat

conductivity of the solid. Further, we use fixed dimensionlessnumbers Re ¼ 2079; Fr ¼ 179;We ¼ 43;Bi ¼ 0:000058;PeF ¼ 10.

0.01 1 100

2

4

6

time

Wet

ting

diam

eter

i

ii

iii

iv

0.01 1 100

30

120

180

time

Dyn

amic

con

tact

ang

le i

ii

iii

iv

(a)

(b)

Fig. 12. Wetting diameter (a) and the dynamic contact angle (b) of a liquid dropletwith he ¼ 30�;We ¼ 43 and Fr = 179, impinging on a hot solid substrate. (i)Re = 260, (ii) Re = 520, (iii) Re = 1040, (iv) Re = 2079.

500 1500 25003

4

5

Re

max

Wd

500 1,500 2,5007

8

Re

t at m

ax W

d

(a)

(b)

Fig. 13. Maximal wetting diameter (a) and the time to attain it (b) for differentReynolds numbers with fixed We = 43 for a liquid droplet.

0.01 1 100

2

4

6

time

Wet

ting

diam

eter

i

ii

iii

iv

0.01 1 100

30

120

180

time

Dyn

amic

con

tact

ang

le

i

ii

iii

iv

(a)

(b)

Fig. 14. Wetting diameter (a) and the dynamic contact angle (b) of a liquid dropletwith he ¼ 30�;Re ¼ 2079 and Fr = 179, impinging on a hot solid substrate. (i)We = 43, (ii) We = 90, (iii) We = 140 and (iv) We = 190.

0 100 2004.6

4.7

4.8

4.9

We

max

Wd

0 100 2008

10

12

We

t at m

ax W

d(a)

(b)

Fig. 15. Maximal wetting diameter (a) and the time to attain it (b) for differentWeber numbers with fixed Re = 2079, Fr = 179 and B = 0.000058.

684 S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687

The fluid flow indirectly depends on the temperature thoroughthe surface tension. A change in the local surface temperature T onthe free surface induce Marangoni convection. However, in allvariants the surface temperature variation is not high. Thus, theflow dynamics is same in all five variants of the solid Peclet number,see the wetting diameter and the dynamic contact angle over time inFig. 16.

Next, we use the new method in Section 3.8 for calculating theheat flux ðQhÞ and the total heat flux (sum over all time intervals)across the liquid–solid interface. The results for all five variants arepresented in Fig. 17. Though, the wetting area is same in all vari-ants, the heat flux varies. The heat flux is larger in the small solid

Peclet number Var (i), whereas the heat flux is smaller in the largesolid Peclet number Var (v). It is interesting to note that the heatflux is same in all five variants when the wetting diameter is large,say around the dimensionless time t = 9, see Fig. 17. During therecoiling, we can see the reverse effect, that is, the heat flux is lessin the small solid Peclet number case Var (i), and it approaches zerofaster than the other variants. This behavior can clearly be seen inthe total heat flux, Fig. 17 (bottom), where there is almost no

0.01 1 100

1

3

5

time

Wet

ting

diam

eter

i

ii

iii

iv

v

0.01 1 100

30

120

180

time

Dyn

amic

con

tact

ang

le

i

ii

iii

iv

v

(a)

(b)

Fig. 16. Wetting diameter (a) and dynamic contact angle (b) of a liquid droplet withhe ¼ 30�;Re ¼ 2079;We ¼ 43 and Fr = 179, impinging on a hot solid substrate.Variants of solid Peclet number (i) PeS = 50, (ii) PeS = 100, (iii) PeS = 250, (iv)PeS = 350 and (v) PeS = 400.

0 10 20−3

−2

−1

0

1 x 10−5

time

Hea

t flu

x

i

ii

iii

iv

v

0 10 20−0.8

−0.4

0

time

Tot

al h

eat f

lux

i

ii

iii

iv

v

(a)

(b)

Fig. 17. Heat flux (a) and total heat flux (b) over a liquid–solid interface withhe ¼ 30�;Re ¼ 2079;We ¼ 43 and Fr = 179, impinging on a hot solid substrate.Variants of solid Peclet number (i) PeS = 50, (ii) PeS = 100, (iii) PeS = 250, (iv)PeS = 350 and (v) PeS = 400.

0.01 1 100

2

4

6

time

Wet

ting

diam

eter i

ii

iii

0 0.01 1 100

30

120

180

time

Dyn

amic

con

tact

ang

le

i

ii

iii

(a)

(b)

Fig. 18. Wetting diameter (a) and dynamic contact angle (b) of a liquid droplet withhe ¼ 30�;Re ¼ 3750;We ¼ 128; Fr ¼ 425, impinging on a hot solid substrate. Vari-ants of Biot number (i) Bi = 0.001, (ii) Bi = 0.1, (iii) Bi = 1.

0 10 20−10

−5

0

x 10−4

time

Hea

t flu

x

i iiiii

0 10 20−100

−50

0

time

Tot

al h

eat f

lux

i iiiii

(a)

(b)

Fig. 19. Influence of the Biot number on the heat flux (a) and on the total heat flux(b) with Re = 3750, We = 128, Fr = 425 (i) Bi = 0.001, (ii) Bi = 0.1, (iii) Bi = 1.

S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687 685

variation in the total heat flux of Var (i) after the dimensionlesstime t = 10.

4.5. Influence of the Biot number on the flow dynamics

In this section we study the influence of the Biot number on theheat flux. We consider a droplet of diameter d0 ¼ 3:0� 10�3 m.Further, the following material parameters are used: the density

q = 1000 kg/m3, the dynamic viscosity l = 0.001 Ns/m2, the surfacetension r = 0.073 N/m, the equilibrium contact angle he = 30�. Also,we assume that uimp ¼ 2:5 m/s, TF;0=300 K, TS;0 = 700 K, T1 = 298 K,Tref = 323 K (temperature of the fluid surface, the solid surface, theatmospheric and the reference, respectively). Using L = d0=2 andU = uimp as characteristic values, we get Re ¼ 3750; Fr ¼ 425;We ¼ 128;PeF ¼ 10;PeS ¼ 100. In this study, we consider threevariants: (i) Bi = 0.001, (ii) Bi = 0.1 and (iii) Bi = 1. The obtainedwetting diameter and the dynamic contact angle over time for all

686 S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687

three variants are presented in Fig. 18. Though the flow dynamicsis similar in first two variants, there are some oscillations in thedynamic contact angle of Var (iii). Recall that the Biot number con-trols the heat flux across the free surface, that is, a large value ofBiot number induces more heat flux. Since we use the same liquidPeclet number in all three variants, the temperature on the freesurface will be less in case of a large Biot number, Var (iii), in com-parison with the Var (i). Therefore, the surface tension force will belarger in Var (iii), see the relation (18), and it induces more rollingmotion while spreading.

Next, the obtained heat flux and the total heat flux across theliquid–solid interface over time are presented in Fig. 19. Asexpected, the heat flux is large in the high Biot number case Var(iii). The heat flux in Var (i) approaches zero when the dropletattains it maximum wetting diameter, whereas the heat transferfrom the solid phase to the liquid phase occurs even during therecoiling stage in the Var (iii), see Fig. 19.

5. Summary

A finite element scheme using the arbitrary Lagrangian–Euleri-an (ALE) approach for computations of a non-isothermal liquiddroplet impinging on a horizontal hot solid substrate is presentedin this paper. The coupled Navier–Stokes and the energy equationsare solved using this numerical scheme. The derivation of theenergy equations in solid and liquid phases into a single variationalequation makes the scheme robust in the ALE approach. The high-lights of the numerical scheme are the Laplace–Beltrami operatortechnique for the curvature approximation and the contact angleinclusion, the ALE approach with moving meshes in both the liquidand the solid phase to track the free surface and to guaranteematching grids at the liquid–solid interface, and the special algo-rithm for calculating the heat flux without differentiating thenumerical solution. The proposed numerical scheme works wellwithout loss of accuracy even for problems with large jumps inthe material parameters across the liquid–solid interface. An arrayof numerical experiments by varying the Reynolds, Weber, Peclet,and Biot numbers are performed for a 3D–axisymmetric non-iso-thermal liquid droplet impinging on a horizontal hot solidsubstrate.

Conflict of interest

None declared.

Acknowledgement

The authors would like to thank the German research founda-tion (DFG, GRK 1554) and the National Board for Higher Mathe-matics (NBHM), India for partially supporting this research.

References

[1] C. Hirt, B. Nichols, Volume of Fluid (VOF) method for the dynamics of freeboundaries, J. Comput. Phys. 39 (1981) 201–225.

[2] J.E. Pilliod, E.G. Puckett, Second-order accuarate volume-of-fluid algorithms fortracking material interfaces, J. Comput. Phys. 199 (2) (2004) 465–502.

[3] M. Renardy, Y. Renardy, J. Li, Numerical simulation of moving contact lineproblems using a volume-of-fluid method, J. Comput. Phys. 171 (2001) 243–263.

[4] W. Rider, D. Kothe, Reconstructing volume tracking, J. Comput. Phys. 141(1998) 112–152.

[5] M. van Sint Annaland, N. Deen, J. Kuipers, Numerical simulation of gas bubblesbehaviour using a three-dimensional volume of fluid method, Chem. Eng. Sci.60 (11) (2005) 2999–3011.

[6] S. Osher, J.A. Sethian, Fronts propagating with curvature dependent speed:algorithms based on Hamilton–Jacobi forumlations, J. Comput. Phys. 79 (1988)12–49.

[7] M. Sussman, P. Smereka, S. Osher, A level set approach for computing solutionsto incompressible two-phase flow, J. Comput. Phys. 114 (1) (1994) 146–159.

[8] J.A. Sethian, Level Set Methods, Cambridge University Press, 1996.[9] M. Sussman, E.G. Puckett, A coupled level set and volume-of-fluid method for

computing 3d axisymmetric incompressible two-phase flows, J. Comput. Phys.162 (2000) 301–337.

[10] S. Gross, V. Reichelt, A. Reusken, A finite element based level set method fortwo-phase incompressible flows, Comput. Visual. Sci. 9 (4) (2006) 239–257.

[11] C.S. Peskin, Flow patterns around heart valves: a numerical method, J. Comput.Phys. 10 (1972) 252–271.

[12] S. Unverdi, G. Tryggvason, A front-tracking method for viscous, incompressiblemulti-fluid flows, J. Comput. Phys. 100 (1992) 25–37.

[13] K.S. Sheth, C. Pozrikidis, Effects of inertia on the deformation of liquid drops insimple shear flow, Comput. Fluids 24 (2) (1995) 101–119.

[14] G. Tryggvason, B. Bunner, A. Esmaeeli, D. Juric, N. Al-Rawahi, W. Tauber, J. Han,S. Nas, Y.-J. Jan, A front-tracking method for the computations of multiphaseflow, J. Comput. Phys. 169 (2) (2001) 708–759.

[15] C.S. Peskin, The immersed boundary method, Acta Numer. (2002) 1–36.[16] C.W. Hirt, A.A. Amsden, J.L. Cook, An arbitrary Lagrangian Eulerian computing

method for all flow speeds, J. Comput. Phys. 14 (3) (1974) 227–253.[17] J. Donéa, Arbitrary Lagrangian–Eulerian finite element methods, in: T.

Belytschko, T.R.J. Hughes (Eds.), Computational Methods for TransientAnalysis, Elsevier scientific publishing co, Amsterdam, 1983, pp. 473–516.

[18] F. Nobile, Numerical approximation of fluid-structure interaction problemswith application to haemodynamics (Ph.D. thesis), École PolytechniqueFédérale de Lausanne, 2001.

[19] P. Nithiarasu, An arbitrary Lagrangian Eulerian (ALE) formulation for freesurface flows using the characteristic-based split (CBS) scheme, Int. J. Numer.Methods Fluids 12 (2005) 1415–1428.

[20] S. Ganesan, L. Tobiska, Computations of flows with interfaces using arbitraryLagrangian Eulerian method, in: Proceedings of ECCOMAS CFD 2006, Egmondaan Zee, The Netherlands, ISBN 90-9020970-0.

[21] S. Ganesan, L. Tobiska, An accurate finite element scheme with moving meshesfor computing 3D-axisymmetric interface flows, Int. J. Numer. Methods Fluids57 (2) (2008) 119–138.

[22] H. Fujimoto, N. Hatta, Deformation and rebounding processes of a waterdroplet impinging on a flat surface above leidenfrost temperature, J. FluidsEng. 118 (1996) 142–149.

[23] M. Pasandideh-Fard, R. Bhola, S. Chandra, J. Mostaghimi, Deposition of tindroplets on a steel plate: simulations and experiments, Int. J. Heat MassTransfer 41 (1998) 2929–2945.

[24] D.J.E. Harvie, D.F. Fletcher, A hydrodynamic and thermodynamic simulation ofdroplet impacts on hot surfaces. Part II: validation and applications, Int. J. HeatMass Transfer 44 (14) (2001) 2643–2659.

[25] R. Ghafouri-Azar, S. Shakeri, S. Chandra, J. Mostaghimi, Interactions betweenmolten metal droplets impinging on a solid surface, Int. J. Heat Mass Transfer46 (8) (2003) 1395–1407.

[26] G. Strotos, M. Gavaises, A. Theodorakakos, G. Bergeles, Numerical investigationon the evaporation of droplets depositing on heated surfaces at low webernumbers, Int. J. Heat Mass Transfer 51 (7–8) (2008) 1516–1529.

[27] G. Strotos, M. Gavaises, A. Theodorakakos, G. Bergeles, Numerical investigationof the cooling effectiveness of a droplet impinging on a heated surface, Int. J.Heat Mass Transfer 51 (19–20) (2008) 4728–4742.

[28] A.M. Briones, J.S. Ervin, S.A. Putnam, L.W. Byrd, L. Gschwender, Micrometer-sized water droplet impingement dynamics and evaporation on a flat drysurface, Langmuir 26 (16) (2010) 13272–13286.

[29] S.A. Putnam, A.M. Briones, L.W. Byrd, J.S. Ervin, M.S. Hanchak, A. White, J.G.Jones, Microdroplet evaporation on superheated surfaces, Int. J. Heat MassTransfer 55 (21–22) (2012) 5793–5807.

[30] L.S.F.Y. Ge, Three-dimensional simulation of impingement of a liquid dropleton a flat surface in the leidenfrost regime, Phys. Fluids 17 (2005) 027104.

[31] L.S.F.Y. Ge, 3-D modeling of the dynamics and heat transfer characteristics ofsubcooled droplet impact on a surface with film boiling, Int. J. Heat MassTransfer 49 (21–22) (2006) 4231–4249.

[32] M. Francois, W. Shyy, Computations of drop dynamics with the immersedboundary method, Part 1: Numerical algorithm and buoyancy-induced effect,Numer. Heat Transfer, Part B: Fundam. 44 (2) (2003) 101–118.

[33] M. Francois, W. Shyy, Computations of drop dynamics with the immersedboundary method, Part 2: Drop impact and heat transfer, Numer. HeatTransfer, Part B: Fundam. 44 (2) (2003) 119–143.

[34] J.F.Z. Zhao, D. Poulikakos, Heat transfer and fluid dynamics during the collisionof a liquid droplet on a substrate. 1. Modeling, Int. J. Heat Mass Transfer 39 (13)(1996) 2771–2789.

[35] S. Ganesan, G. Matthies, L. Tobiska, On spurious velocities in incompressibleflow problems with interfaces, Comput. Methods Appl. Mech. Eng. 196 (7)(2007) 1193–1202.

[36] G. Dziuk, Finite elements for the Beltrami operator on arbitrary surfaces, in: S.Hildebrandt, R. Leis (Eds.), Partial Differential Equations and Calculus ofVariations, Springer, Berlin, 1988, pp. 142–155.

[37] G. Dziuk, An algorithm for evolutionary surfaces, Numer. Math. 58 (1991)603–611.

[38] S. Ganesan, Finite element methods on moving meshes for free surface andinterface flows (Ph.D. thesis), Otto-von-Guericke-Universitat, Fakultät furMathematik, Magdeburg, 2006.

[39] S. Ganesan, L. Tobiska, Modelling and simulation of moving contact lineproblems with wetting effects, Comput. Visual. Sci. 12 (2009) 329–336.

S. Ganesan et al. / International Journal of Heat and Mass Transfer 78 (2014) 670–687 687

[40] M. Behr, F. Abraham, Free-surface flow simulations in the presence of inclinedwalls, Comput. Methods App. Mech. Eng 191 (47–48) (2002).

[41] J. Eggers, H.A. Stone, Characteristic lengths at moving contact lines for aperfectly wetting fluid: the influence of speed on the dynamic contact angle, J.Fluid Mech. 505 (2004) 309–321.

[42] L.M. Hocking, A moving fluid interface. Part 2: The removal of the forcesingularity by a slip flow, J. Fluid Mech. 79 (2) (1977) 209–229.

[43] C. Huh, L.E. Scriven, Hydrodynamic model of steady movement of a solid/liquid/fluid contact line, J. Colloid Interface Sci. 35 (1971) 85–101.

[44] R.G. Cox, The dynamics of the spreading of liquids on a solid surface. Part 1.Viscous flow, J. Fluid Mech. 168 (1986) 169–194.

[45] G. Karniadakis, A. Beskok, N. Aluru, Microflows and Nanoflows, Springer, 2005.[46] E.B.V. Dussan, The moving contact line: the slip boundary condition, J. Fluid

Mech. 77 (4) (1976) 665–684.[47] E. Lauga, P. Brenner, H.A. Stone, Microfluidics: The no-slip boundary condition,

in: J. Foss, C. Tropea, A. Yarin (Eds.), Handbook of Experimental FluidDynamics, Springer, New-York, 2005.

[48] E. Bänsch, Numerical methods for the instationary Navier–Stokes equationswith a free capillary surface, Habilitationsschrift, Albert-Ludwigs-Universität,Freiburg i. Br., 1998.

[49] K. Deckelnick, G. Dziuk, C. Elliott, Computation of geometric partial differentialequations and mean curvature flow, Acta Numer. 14 (2005) 139–-232.

[50] G. Dziuk, C. Elliott, Finite element methods for surface pdes, Acta Numer. 22(2013) 289–396.

[51] S. Turek, Efficient solvers for incompressible flow problems, Lecture Notes inComputational Science and Engineering, vol. 6, Springer-Verlag, Berlin, 1999,http://dx.doi.org/10.1007/978-3-642-58393-3. <http://dx.doi.org/10.1007/978-3-642-58393-3> (an algorithmic and computational approach, With 1CD-ROM (‘‘Virtual Album’’: UNIX/LINUX, Windows, POWERMAC; ‘‘FEATFLOW1.1’’: UNIX/LINUX)).

[52] J.W. Barrett, H. Garcke, R. Nürnberg, Eliminating spurious velocities with astable approximation of incompressible two-phase flow, Tech. Rep. 12,University Regensburg, 2013.

[53] E. Bänsch, Finite element discretization of the Navier–Stokes equations with afree capillary surface, Numer. Math. 88 (2) (2001) 203–235.

[54] V. Girault, P.-A. Raviart, Finite Element Methods for Navier–Stokes equations,Springer Verlag, Berlin, Heidelberg, New York, 1986.

[55] G. Matthies, Finite element methods for free boundary value problems withcapillary surfaces (Ph.D. thesis), Otto-von-Guericke-Universität, Fakultät fürMathematik, Magdeburg, 2002.

[56] M.G. Larson, Analysis of adaptive finite element methods (Ph.D. thesis),Department of Applied Mathematics, Chalmers University, Göteberg, Sweden,1996.

[57] J. Shen, C. Graber, J. Liburdy, D. Pence, V. Narayanan, Simultaneous dropletimpingement dynamics and heat transfer on nano-structured surfaces, Exp.Therm. Fluid Sci. 34 (4) (2010) 496–503.

[58] S. Sikalo, M. Marengo, C. Tropea, E. Ganic, Analysis of impact of droplets onhorizontal surfaces, Exp. Therm. Fluid Sci. 25 (7) (2002) 503–510.