You are on page 1of 20

International Journal of Multiphase Flow 116 (2019) 250–269

Contents lists available at ScienceDirect

International Journal of Multiphase Flow


journal homepage: www.elsevier.com/locate/ijmulflow

Single vapour bubble growth under flash boiling conditions using a


modified HLLC Riemann solver
D. Dietzel a, T. Hitz b, C.-D. Munz b, A. Kronenburg a,∗
a
Institut für Technische Verbrennung, Universität Stuttgart, Herdweg 51, 70174 Stuttgart
b
Institut für Aerodynamik und Gasdynamik, Universität Stuttgart, Pfaffenwaldring 21, 70569 Stuttgart

a r t i c l e i n f o a b s t r a c t

Article history: Common fuel-oxidizer combinations in orbital manoeuvring systems consist of toxic substances but will
Received 18 June 2018 be replaced in the future. Liquid oxygen (LOX) is one potential oxidizer but it rapidly superheats under
Revised 28 February 2019
the initial low-pressure conditions in the combustion chamber. Bubble nucleation and growth dominate
Accepted 12 April 2019
the efficient disintegration of the liquid jet, which is called flash boiling. An investigation of the small
Available online 17 April 2019
scale bubble dynamics will help to improve existing models for the break-up of the LOX jet and the mix-
Keywords: ing with the fuel. Direct numerical simulations (DNS) can be used to analyse the underlying processes if
Single vapour bubble the solver handles phase transition effects at extreme ambient conditions. In this paper we use a fully
Riemann solver compressible discontinuous Galerkin solver combined to a level-set equation and to a modified HLLC
Level-set Riemann solver. The main goal of our investigation is to assess closure conditions for the mass transfer
Multiphase flow models which accurately represent the physics of vapour bubble growth. We couple three models for the
Phase change
estimation of vaporisation mass fluxes to the interface Riemann solver. The Hertz-Knudsen relation and a
kinetic relation predict the volumetric expansion well but cannot represent the instantaneous mass flux.
A sub-grid scale heat flux model predicts the mass flux qualitatively better but the volumetric bubble ex-
pansion matches only at late time intervals. The dependency of the calibrated coefficients on the physical
conditions is similar for the different vaporisation models.
© 2019 Elsevier Ltd. All rights reserved.

1. Introduction 2015). However, any current modelling strategy does not resolve
bubble nucleation and growth of individual bubbles and will
Space probes, rockets and satellites require manoeuvring sys- require closures at the sub-grid scale (SGS). A quantification of
tems for e.g. attitude control, re-boost and re-entrance into the these SGS processes, i.e. bubble nucleation and bubble growth,
atmosphere. Small thrusters provide the necessary propulsion is necessary to improve the existing models. A possible strategy
using the recoil energy of the exhaust gases from an upstream would be the use of direct numerical simulations (DNS) to build a
combustion process (Tanaka et al., 2011). Such orbital maneuvering database that allows for a detailed investigation of the underlying
systems typically use hypergolic propellants such as monomethyl physics and possible closures. DNS of flash vaporisation resolving
hydrazine, but current interest focuses on their replacement by all the necessary processes are, however, extremely difficult to
non-toxic fuel-oxidizer combinations. One such combination for realize and two of the key aspects are listed here:
the new generation thrusters is gaseous methane and liquid
• A compressible flow field solver is a precondition for the ac-
oxygen (LOX) (Manfletti, 2014). The disintegration of the LOX jet
curate modelling of the early stages of flash vaporisation. The
controls the subsequent mixing with the fuel and determines the
early bubble growth is initiated by the local pressure field
prospects of successful ignition events. If the chamber pressure is
which changes as the bubbles grow. The vapour densities in-
below saturation conditions, the LOX becomes superheated and
side the bubbles are directly linked to their pressures and the
small vapour bubbles nucleate and grow. Consequently, the jet
fluid cannot be considered incompressible.
quickly expands and breaks up into small droplets. This process
• It is important to realize that the term DNS – as it is com-
is called flash boiling and can enhance the mixing process and
monly used –refers to a macroscopic approach where contin-
efficiency of the subsequent combustion (Lamanna et al., 2014;
uum assumptions remain valid, all (turbulent) scales including
interface boundaries are well resolved and do not require mod-

Corresponding author. elling. For flash vaporisation, however, the phase change ap-
E-mail address: kronenburg@itv.uni-stuttgart.de (A. Kronenburg). pears in unclosed form and modelling at molecular scale would

https://doi.org/10.1016/j.ijmultiphaseflow.2019.04.010
0301-9322/© 2019 Elsevier Ltd. All rights reserved.
D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269 251

be needed. The coupling of molecular dynamic simulations and investigation is to find a suitable calibration for different types of
DNS for fully 3D bubble dynamics is beyond current capabilities vaporisation models such that the expansion of vapour bubbles in
and some modelling of phase change is required. superheated jets is accurately predicted. For this purpose we use
the growth of single vapour bubbles as a reference to validate the
The current paper will primarily focus on the implementation model calibration. Using this simplified setup we determine the re-
of different closure strategies for the phase transfer rates in a com- quired coefficients, we demonstrate challenges in the compressible
pressible DNS code, their calibration and their influence on bubble DNS of vapour bubble growth and show how these challenges can
growth. A good number of compressible multiphase solvers have be overcome. In particular, we cover the following aspects:
been developed within the last decade with recent examples in-
troduced in (Saurel et al., 2008; Zein et al., 2010; Lauer et al., • We quantify the growth of single vapour bubbles for cryogenic
2012; Chang et al., 2013; Houim and Kuo, 2013; Saurel et al., 2016; oxygen for different temperature, pressure and time ranges.
Lee and Son, 2017; Fechter et al., 2017; 2018). Saurel and his co- • We justify simplifications in the initial conditions for vapour
workers (Saurel et al., 2008; 2016) as well as (Zein et al., 2010) bubble growth that considerably reduce the computational
solved separate equation systems for the two phases including one time. This aspect is particularly important for future simula-
volume fraction equation for each phase. The phase indicator dif- tions of multiple bubble growth.
fuses across several cells which are treated as homogeneous mix- • We use the Hertz–Knudsen relation (HKR), the kinetic relation
tures of liquid and vapour. This allows to treat two-phase flows (KR) from Fechter et al. (2018) and an SGS heat flux model
for larger length scales because interfaces of individual bubbles (SHFM) to estimate the vaporisation mass flux and analyse their
do not need to be resolved. However, a sharp separation of the calibration coefficients for the different temperature, pressure
two phase, as it is needed for a DNS of bubble dynamics at a and time ranges.
sub-grid scale, is difficult to achieve. Chang et al. (2013) solved
the fully compressible Navier-Stokes equations with a level-set The paper is organized as follows: In Section 2 the fundamental
approach (Sussman and Smereka, 1994) that keeps the inter- equations are presented. The numerical approach and the setup of
face between liquid and vapour sharp. However, Chang et al. the present test cases are described in Sections 3 and 4. We show
did not account for phase change. Lauer et al. (2012), Lee and our results in Section 5 and summarize our findings in Section 6.
Son (2017) and Fechter et al. (2017, 2018) combined the level-set
algorithm with compressible solvers that can account for phase
transition. Lauer et al. (2012) used interface Riemann solvers with- 2. Fundamentals
out phase change but source terms accounted for the mass trans-
fer from liquid to vapour. They estimated the mass fluxes with The fully compressible multiphase solver presented in
the Hertz–Knudsen relation and calibrated the model coefficients Fechter et al. (2018) serves as a basis for the work presented
for oscillating cavitation bubbles. However, the physics of oscil- here. Viscous forces are neglected since pressure force and phase
lating bubbles differs from the physics of vapour bubble growth. transition will dominate the expansion of the vapour bubbles. The
Moreover, Lauer et al. (2012) considered water and the coefficients thermal boundary layers at the bubble walls do not exceed 15%
may not be valid for other fluids. Lee and Son (2017) used a ghost of the bubble diameter for the relevant process conditions (e.g.
fluid method and incorporate phase transfer into the jump condi- Fig. 1(b)) and are typically even smaller for the majority of our
tions which determine the ghost states. They estimated the mass test cases (cf. Section 4 for details on the numerical configuration).
flux from the heat balance at the interface which requires a suit- An adequate resolution would require cell sizes of much less than
able resolution of the temperature boundary layer. The heat flux 1 μm, but such high resolution requirements are not feasible for
determines the mass flux at the interface and no model coeffi- simulations of multiple bubbles and their interactions. Thus, we
cient requires calibration. Lee and Son (2017) used this solver to do not resolve the heat diffusion but account for its influence by
investigate the growth and shrinkage of single compressible bub- the calibration of the vaporisation models. Under these conditions
bles. However, the current paper focuses on LOX jets under flash the Navier–Stokes equations can be simplified to the Euler equa-
boiling conditions where thermal boundary layers are thin and tions as introduced in Section 2.1. In Section 2.2, we present the
their resolution is not feasible if numerous bubbles are to be re- different models for the vaporisation mass flux and the interface
solved in the computational domain. Fechter et al. developed an capturing approach is outlined in Section 2.3.
exact (Fechter et al., 2017) and an approximative (Fechter et al.,
2018) Riemann solver for general two-phase Riemann problems.
The equation system consisted of Rankine–Hugoniot and interface 2.1. Conservation laws and jump conditions at the interface
jump conditions and they used a kinetic relation to estimate va-
porisation rates. In Fechter et al. (2018) they showed that the ap- The compressible Euler equations that govern the conservation
proximate Riemann solver, called HLLP Riemann solver, is faster of mass, momentum and energy read
than the exact solver and it gives a reasonable approximation of
the phase transition in shock tubes and for single droplets. This Ut + ∇ · F (U ) = 0 (1)
Riemann solver is a modification of HLLC-type Riemann solvers
(Hu et al., 2009) and will be the basis of our work. However, gen- where
eral calibration laws for the model parameter in the kinetic re-
lation do not exist but are required to accurately predict the ex- U = ( ρ , ρ u1 , ρ u2 , ρ u3 , E )T (2)
pansion of vapour bubbles under flash boiling conditions. None of
these approaches can directly be used for the simulation of multi- is the vector of the conserved variables. Here, ρ , u1 , u2 , u3 and E
ple bubble growth in superheated LOX jets. The model coefficients are the density, the three Cartesian velocity components and the
in Lauer et al. (2012) and Fechter et al. (2018) are not necessarily total energy, respectively. The latter is the sum of the internal en-
u2
valid for the physical conditions of our investigation and general ergy e and the kinetic energy, E = ρ (e + 2i ). Here, i indicates the
calibration laws do not exist. The resolution of the thermal bound- three Cartesian velocity components and we use Einsteins notation
ary layer as in Lee and Son (2017) is not feasible in cases with mul- to sum up. The subscript t in Eq. (1) indicates the partial derivative
tiple bubbles for our process conditions. Thus, the main goal of our in time. The flux vectors F(U) for the three Cartesian coordinate
252 D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269

Fig. 1. Aspects of the sub-grid scale heat flux model.

directions read recent review on its application and calibration. The model equa-
⎛ ρ u1 ⎞ ⎛ ρ u2 ⎞ ⎛ ρu ⎞ tion reads
3 

⎜ ρ u1 + p ⎟ ⎜ ρ u1 u2 ⎟ ⎜ ρ u1 u3 ⎟
2
1 psat pvap
⎜ ⎟ F2 = ⎜ ρ u22 + p ⎟, F3 = ⎜ ρ u2 u3 ⎟, (3) m˙  = √ λvap − λcond
⎝ ρ u2 u1 ⎠,
F1 = (6)
⎝ ⎠ ⎝ 2 ⎠ 2π R Tliq Tvap
ρ u3 u1 ρ u3 u2 ρ u3 + p
u1 ( E + p ) u2 ( E + p ) u3 ( E + p ) where λvap and λcond are the vaporisation and condensation coeffi-
cients, respectively. They have to be scaled for a certain application
and p denotes pressure. An equation of state (EOS) gives the rela-
and we outline the calibration for our cases in Section 5.
tion between density, pressure and energy, i.e. p = f (ρ , e ), closing
the equation system (1) for a single phase. The open-source EOS The kinetic relation (KR)
library CoolProp (Bell et al., 2014) is coupled to the flow solver Fechter et al. (2018) used the KR in their original formulation
and any correlation is obtained from the functions provided in of the HLLP Riemann solver. They provided a detailed derivation
Bell et al. (2014). for the final model equation (Fechter et al., 2017),
We solve Eq. (1) separately for the individual bulk phases, i.e.
1
the liquid and the vapour phases. To couple these two bulk phases m˙  = T [[s − ssat ]]. (7)
we impose the jump conditions (Fechter et al., 2018) cent ref
This model ensures consistency with the second law of thermody-
[[ρ (un − Sint )]] = 0, namics, but entropy production at the interface is unknown and
[[ρ un (un − Sint ) + p]] = σ κ , (4) requires scaling by the coefficient cent (Fechter et al., 2017). For
[[ρ e(un − Sint ) + pun ]] + m˙  hlv = Sint σ κ . constant interface states, m˙  is inversely proportional to cent such
that the estimated mass flux decreases with an increasing model
Here, Sint , σ , κ and hlv are the interface velocity, the surface ten- coefficient.
sion coefficient, the curvature and the latent heat of vaporisation,
respectively. The mass flux due to vaporisation is m˙  and the oper- A sub-grid scale heat flux model (SHFM)
ator [[φ ]] denotes the difference between the vapour and the liq- The SHFM is based on the energy balance at the interface. Ne-
uid state, i.e [[φ ]] = φvap − φliq . The Euler equations do not resolve glecting the heat diffusion in the vapour phase we obtain the mass
the heat flux but the term m˙  hlv supplies the enthalpy jump across flux as the ratio of an approximated heat flux in the liquid phase
the interface. We model m˙  as will be presented in Section 2.2 and and the latent heat, viz.
evaluate hlv at Tref = 0.5(Tvap + Tliq ) from the EOS (Fechter et al., qliq − qvap qliq
2017; 2018). m˙  = ≈ . (8)
hlv hlv
The assumption of negligible diffusive heat fluxes in the vapour
2.2. Vaporisation model phase is justified as small temperature gradients in the bubble at
the interface can be expected. The heat diffusivity is higher in the
The interface velocity Sint and the mass flux m˙  in Eq. (4) are liquid and the heat flux in the vapour becomes important only for
coupled by very large temperature gradients. As the HLLP Riemann solver does
m˙  = ρliq (uliq − Sint ) = ρvap (uvap − Sint ) (5) not allow for diffusive effects, the thermal boundary layer will not
be resolved and we approximate the heat flux, qliq , by
and we use the following models to estimate m˙  and to close the
dT T∞ − Tsat
system of equations as given by the jump conditions (4). qliq = −kliq
dx x=xint
≈ −kliq 2
δ
(9)

The Hertz–Knudsen relation (HKR) assuming a parabolic temperature profile in the boundary layer
The HKR originates from Hertz (1882) and Knudsen (1950) and as illustrated in Fig. 1(a). Similar approaches to estimate the heat
is based on a mass balance at the liquid-vapour interface. This flux in an unresolved thermal boundary layer exist for droplets
balance law is derived from the kinetic gas theory comparing the (Sazhin, 2014) but have not intensively been discussed in the con-
molecule fluxes from and to the interface. Schrage (1953) compre- text of vapour bubble growth. The concept of the model within our
hensively discussed the HKR and (Persad and Ward, 2016) gave a compressible solver is the following:
D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269 253

• The saturation temperature is determined according to the multiplying the resulting system of equations by a test function φ ,
vapour pressure inside of the bubble. Starting from isothermal the governing equations in reference space read
conditions (see Section 4) the mass flux is initially set to zero.
• The bubble growth is triggered numerically (see Section 4) and JUt φ dξ + ∇ξ · Fˆ (U )φ dξ = 0, (13)
E E
with increasing size the pressure in the bubble will decrease.
• Consequently, the saturation temperature decreases and the with the Jacobi determinant of the transformation J. An integration
temperature gradient in Eq. (9) becomes non-zero. The mass by parts of the second term in Eq. (13) gives the weak formulation
flux increases proportional to the temperature gradient. of the conservation laws in the discontinuous Galerkin framework,
• If the mass flux increases continuously, the vapour pressure will

increase and the decreasing temperature gradient relaxes the ∂
JU φ dξ + (Fˆ · N )φ dS − Fˆ (U ) · (∇ξ φ )dξ = 0. (14)
mass flux. ∂t E ∂E E

The accuracy of the SHFM depends on the estimation of the The first and the second terms on the left hand side define the
boundary layer thickness in Eq. (9) which we determine as time derivative of the conserved quantity and the fluxes across the
surface of the DG cell, respectively. Each cell provides an inner cell
δ = Cδ R. (10)
resolution which is accounted for by a continuous in-cell ansatz
Fig. 1(b) shows an example for the ratio δ /R = Cδ as a function of function. As basis of the ansatz function Lagrange polynomials are
the bubble radius. In a first approach we keep Cδ constant within used, where the polynomial degree N determines the accuracy of
one simulation but calibrate the coefficient for each individual test integration given in the third term on the left side of Eq. (14). As
case. polynomials are local for every cell, discontinuities may arise at
the cell face and Riemann solvers are used to determine consistent
2.3. Level-set equation fluxes across the cell boundaries. We apply HLLC Riemann solvers
at single phase cell boundaries (Toro, 2009) and couple the DG ap-
We track the liquid-vapour interface using the level-set ap- proach to a third order Runge–Kutta scheme for time integration.
proach introduced by Sussman and Smereka (1994) where a signed
distance function (x, t) is transported as
3.2. Resolution at the interface

+ uLS ∇ · ( ) = 0. (11)
∂t At the phase interface of small bubbles a strong contact dis-
The level-set determines the distance to the interface and the zero- continuity appears for many physical quantities, most notably for
isocontour of defines the interface location. We obtain the ad- density and pressure. Fluxes across the interface cannot be deter-
vection velocity uLS of the interface from Eq. (5) and extend this mined in the same manner as for single phase cell interfaces since
velocity to several neighbouring cells. We solve a Hamilton–Jacobi the discontinuity would lead to strong instabilities in the solution
equation to reinitialize such that it remains a signed distance procedure. Moreover, phase change complicates the Riemann prob-
function (Sussman and Smereka, 1994). The normal vector of the lem at the interface where additional terms in the jump conditions
interface and its curvature are determined using the space deriva- appear. We determine the fluxes across the interface using the
tives of according to approximate Riemann solver from Fechter et al. (2018), which is
based on the ghost fluid method of Fedkiw et al. (1999). Fig. 2 de-
∇ ∇
 int =
n and κ = ∇ · . (12) picts a typical two-phase Riemann problem including rarefaction,
|∇ | |∇ | shock and contact waves as well as the phase interface as an ad-
ditional discontinuity. The different fluid states describe the initial
3. Numerical methods liquid state, rarefied liquid, vaporised fluid, compressed vapour and
the initial vapour state from the left to the right, respectively. An
3.1. Discontinuous Galerkin spectral element method exact solution of the Riemann problem requires an iterative solu-
tion of the Rankine–Hugoniot and jump conditions across the dif-
The Euler equations are solved numerically by a discontinuous ferent waves. The HLLP Riemann solver simplifies its solution using
Galerkin (DG) spectral element method. Here, we give a brief intro- the following assumptions:
duction but the reader is referred to Hindenlang et al. (2012) and
Kopriva (2009) for a more detailed description. Within the dis- • The outer wave speeds SL and SR are approximated by a lin-
continuous Galerkin approach the simulation domain is discretised earization of the Lax curves according to Fechter et al. (2013).
with a set of control volumes . Each cell is handled separately • The contact wave is omitted as proposed by
and in a first step, a control volume is mapped from physical space, Harten et al. (1983) for single phase flow.
x, to a reference space, ξ . The computational cells are spanned to a • The energy coupling across the Rankine–Hugoniot condition on
unit reference element E of the size [-1;1] in each direction. After the vapour side is dropped.

Fig. 2. Wave fan of a two-phase Riemann problem with phase change. Left: Exact solution of the Riemann problem. Right: Approximate solution of the Riemann problem
in the HLLP context. The contact wave is omitted reducing the complexity of the problem.
254 D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269

Fig. 3. 1D example for the DG scheme with FV sub-cells: Each DG cell provides an inner cell resolution. In the vicinity of the interface the nodes of DG polynomials are
treated as FV cells. The numerical fluxes in the bulk phase are determined using a single-phase HLLC Riemann solver. The numerical fluxes at the phase interface are
determined using the HLLP Riemann solver.

The mathematical tool Maplesoft® solves the remaining system (RPE)


of equations and provides explicit equations for the internal states 3 1
2σ 4μ

U∗L and U∗R . Fechter et al. (2018) discussed these simplifications in RR̈ + (R˙ )2 = p − p∞ − − R˙ (15)
2 ρliq sat R R
detail and provided the explicit equations for the internal states.
The numerical fluxes are determined as functions of the outer and for single bubble growth. Here, R, R˙ and R̈ are the bubble radius
the inner states according to Hu et al. (2009). and its first and second time derivative, respectively. The solution
As continuous polynomials are used within one DG cell, it is of Eq. (15) provides the temporal evolution of the bubble radius R
not possible to capture the interface discontinuity within the cell and its derived quantities (growth velocity, acceleration). Lee and
at polynomial nodes. To enhance the interface resolution, the poly- Herman Merte (1996) coupled the RPE to the temperature equation
nomial nodes are treated as finite volume (FV) cells in the vicinity  2 
of the interface by subdividing each DG cell into N + 1 FV sub-cells ∂T ∂T ∂ T 2 ∂T
+ ur =α + (16)
per spatial direction. These cells are decoupled from their polyno- ∂t ∂r ∂ r2 r ∂ r
mial ansatz function and discontinuities may arise between them.
For the present investigation a sharp interface approach is used. and derived suitable boundary conditions at the bubble wall from
Every computational cell is considered as either liquid or gaseous Eq. (8). We illustrate the assumptions commonly made for single
and there are no mixed cells. Thus, interface fluxes can be deter- vapour bubble growth in Fig. 4. The surrounding liquid is super-
mined directly across FV cell faces in the refined region and the heated, i.e. it is below its saturation pressure at a metastable state
interface is shifted from one cell face to the next whenever the A, and vapour bubbles nucleate isothermally under saturation pres-
sign of changes. This approach is needed to avoid a smearing sure (state B). The surface tension force balances the pressure dif-
of the interface over several computational cells. Fig. 3 depicts this ference between vapour and liquid for a bubble of critical radius
concept that has been adapted from Sonntag and Munz (2014) and (Brennen, 1995)
Fechter and Munz (2015). 2σ
Rcrit = . (17)
4. Physical conditions and numerical setup psat (T∞ ) − pliq
Tiny fluctuations overcome the mechanical equilibrium and the
The reader is reminded here of the objective of the work: the bubble grows initially at an extremely low rate. After the initial
development and validation of a solver that approximates vapour delay the bubble grows at high rates driven by the inertia of the
bubble growth under flash boiling conditions. The problem as such liquid. The heat required for the vaporisation of the liquid cools
involves three-dimensional multiple bubble dynamics and their in- down the interface, and the vapour pressure as well as the growth
teractions. The macroscopic simulation requires, however, mod- rate decrease. This process is indicated in Fig. 4 by the path from
elling the interface mass transfer. Hence, a suitable strategy is de- state B to state C on the vapour side. A small boundary layer devel-
veloped and validated here with the aid of the relatively simple ops from the bubble wall towards the liquid, the vapour pressure
growth dynamics of a single, spherical bubble in a superheated approaches the ambient pressure and heat conduction drives the
liquid. For such a case, analytical models and semi-analytical so- bubble growth. Eqs. (15) and (16) provide a relatively simple, one-
lutions exist (Plesset and Properetti, 1977; Prosperetti and Plesset, dimensional integrable solution (Lee and Herman Merte, 1996),
1978; Brennen, 1995; Lee and Herman Merte, 1996; Robinson and they account for the inertia and the heat diffusion driven stages of
Judd, 2004; Prosperetti, 2017) and serve as reference for the mod- bubble growth and can thus serve as a reference for the DG code
els used in the discontinuous Galerkin solver. In Section 4.1 we development.
introduce the most important aspects of single vapour bubble The Rayleigh-Plesset equation (Rayleigh, 1917; Plesset and Prop-
growth and the reference solution. We summarize the relevant eretti, 1977) and its coupling to the temperature equation by
physical conditions in Section 4.2, and outline the numerical setup Lee and Herman Merte (1996) describe the growth of isolated bub-
in Sections 4.3 and 4.4. bles using several assumptions that warrant some further discus-
sion:
4.1. Single vapour bubble growth
1. 1st assumption: The system consists of a spherical bubble in an
Plesset (1949) and Plesset and Properetti (1977) extended the infinite liquid reservoir. This assumption allows to employ ra-
early work of Rayleigh (1917) to the Rayleigh-Plesset equation dial symmetry for the solution of bubble expansion. In a truly
D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269 255

Fig. 4. Initial conditions for the present setup. State A indicates a metastable superheated liquid. State B indicates a vapour bubble which nucleates within the liquid. We
assume an isothermal nucleation with the vapour bubble being saturated. The resulting pressure gap is balanced by the surface tension.

flashing liquid multiple bubbles occur and the bubble density Table 1
Initial conditions for bubble growth parametric study. Densities are given in
is typically high. The bubbles may therefore interact with their
[kg/m3 ], pressures are given in [bar], critical radii in [μm] and temperatures in
neighbour and deviate from sphericity. The present study, how- [K]. The Jakob number Ja is dimensionless.
ever, considers isolated bubbles in a sufficiently extended liquid
Rp = 5 Rp = 10 Rp = 50 Rp = 100
environment only, and the fluid properties need to be selected
with respect to the ambient conditions (i.e. pressure and tem- T∞ = 90.0 pliq 0.199 0.099 0.02 0.01
perature) typical for flash boiling of cryogenic oxygen. Multiple psat = 0.994, ρ liq 1141.927 1141.905 1141.887 1141.885
ρvap = 4.39 Rcrit 0.332 0.295 0.271 0.268
bubble arrangements will be discussed in forthcoming work.
Ja 117.545 298.334 1965.017 4185.898
2. 2nd assumption: The vapour in the bubble is saturated and in
T∞ = 100.0 pliq 0.508 0.254 0.051 0.025
thermodynamic equilibrium with the liquid at the interface. If the psat = 2.54, ρ liq 1090.303 1090.230 1090.171 1090.164
growth rate of the bubble exceeds the rate at which vapour ρvap = 10.425 Rcrit 0.106 0.094 0.086 0.085
can be generated, the bubble pressure drops below the sat- Ja 62.883 158.303 1022.836 2162.338
uration pressure (Plesset and Zwick, 1954; Plesset and Prop- T∞ = 110.0 pliq 1.087 0.543 0.109 0.054
eretti, 1977). An estimation of the maximum rate of vapour psat = 5.434, ρ liq 1033.779 1033.566 1033.395 1033.374
generation is difficult because the accommodation coefficients ρvap = 21.281 Rcrit 0.039 0.034 0.031 0.031
Ja 39.698 99.442 632.984 1329.518
of the underlying kinetic model are in general not known
T∞ = 120.0 pliq 2.045 1.022 0.204 0.102
(Prosperetti, 2017). Shusser et al. (20 0 0) gave an example for
psat = 10.0223, ρ liq 969.208 968.612 968.133 968.073
an extension of the Rayleigh-Plesset equation incorporating a ρvap = 39.308 Rcrit 0.015 0.013 0.012 0.012
kinetic model for phase transfer rates, and nonequilibrium ef- Ja 28.780 72.016 453.795 948.112
fects are briefly addressed in Prosperetti (2017). Up to today,
uncertainties in the estimation of these nonequilibrium effects
at the interface persist, but it is reasonable to assume here that
processes at molecular scale are much faster than macroscopic context of flash boiling (Sher et al., 2008; Bar-Kohany and Levy,
processes (e.g. bubble expansion) as interface velocities are not 2016) and it is deemed reasonable to use the implementation by
too high (cf. Fig. 24). We therefore use the equilibrium assump- Lee and Herman Merte (1996) for the present investigation.
tion as it is common in similar studies.
3. 3rd assumption: the vapour pressure inside the bubble is con-
stant and the liquid is incompressible. The growth velocities of 4.2. Physical conditions
the isolated bubbles in this work do not exceed 20 m/s and
are therefore considerably smaller than the speed of sound in Experiments on flash boiling LOX jets are currently conducted
the vapour. It is valid to assume that the bubble pressure in- at the German Aerospace Centre (DLR) in Lampoldshausen and
staneously follows the pressure at the bubble wall (Plesset and first results of the jet break-up under such conditions have been
Zwick, 1954). The speed of sound of the liquid oxygen is larger presented in Lamanna et al. (2015). In these experiments, the
than 640 m/s for the present operating conditions confirming temperature of the cryogenic oxygen that is injected into a test
the assumption of an incompressible liquid phase. chamber ranges from T = 90 K to T = 120 K and the same range
is covered here. We define the superheat ratio between saturated
It is emphasized here that the approximation of Lee and Her-
vapour pressure and ambient pressure as (Lamanna et al., 2014)
man Merte (1996) is in general valid for both, the inertia and the
heat diffusion controlled regimes of bubble growth. Plesset and psat
Zwick (1954) derived an analytical model for the heat diffusion Rp = . (18)
p∞
controlled stage taking into account the evaporative cooling. This
model is, however, valid for thin boundary layers only. The proce- Table 1 lists the initial conditions of the different test cases that
dure of Lee and Merte resolves the temperature field and assump- span the entire temperature range and provide conditions from
tions regarding the boundary layer thickness are not needed. Up to very moderate to high superheat. We label the cases CX − Y with
today, the Rayleigh-Plesset equation is a common reference for the X and Y being the superheat ratio Rp and the initial temperature,
understanding and the quantification of the growth process in the respectively (e.g. C5-120 for R p = 5 and T0 = 120 K).
256 D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269

Fig. 6. Reference solution for case C100-90 with (solid lines) and without (dashed-
Fig. 5. Dimensionless bubble radius over time for different combinations of T0 and lines) coupling of the temperature Eq. (16). Black lines show the dimensionless ra-
Rp , respectively. The dashed line defines the analytical solution in the inertia con- dius R+ and blue lines show the instantaneous mass flux. (For interpretation of the
trolled range. The dash-dotted line indicates the analytical solution in the heat dif- references to colour in this figure legend, the reader is referred to the web version
fusion controlled range. of this article.)

Note that single bubble growth is typically characterized by the


Jakob number Ja because flash vaporisation is not limited to single bubble growth,
ρliq cp,liq (T∞ − Tsat ( p∞ )) nucleation rates are high and bubbles will rapidly interact and
Ja = . (19)
ρsat,vap ( p∞ )hlv merge. Only case C5-120 becomes heat diffusion controlled within
this time interval and all other cases are either inertia controlled
It measures the degree of superheat as the ratio between heat
or lie within the intermediate range. Increasing Rp or Ja shifts the
stored in the liquid and the latent heat needed for vaporisation
dimensionless solution towards the inertia controlled range for the
(Prosperetti, 2017). We base our parameter study on the pressure
same final time while an increase in T0 has the opposite effect.
ratio Rp instead because it is one important key parameter in flash
Lee and Herman Merte (1996) observed the same trend comparing
boiling (e.g. Lamanna et al. (2014)). Table 1 lists the corresponding
different experimental investigations. This analysis shows that the
Ja which we also consider in the evaluation of our results.
majority of our test cases is inertia controlled justifying the use
We have discussed in Section 4.1 that bubble growth passes
of the Euler equations. At the same time, we cannot calibrate the
different stages, where different physical mechanisms determine
vaporisation models towards a mass flux which is solely derived
the growth rates. If time and bubble size are normalized by
from the solution of Eq. (15) without coupling of the temperature
Prosperetti (2017)
equation as we explain with the aid of Fig. 6.
A2 t AR Fig. 6 demonstrates the effect of heat diffusion and evapora-
t+ = , R+ = 2 , (20)
B2 B tive cooling on the reference solution. The black lines depict the
  bubble radius over time for case C100-90 with (solid lines) and
2 hlv ρvap (T∞ −Tsat )
with A = 3 ρliq Tsat and B = Ja 12
π αliq , then the asymp- without (dashed lines) diffusive terms, where the initial bubble
totes of the inertia and heat conduction driven stages can be growth is nearly identical. The blue lines show the correspond-
approximated by linear and square root dependencies of R+ on ing vaporisation mass flux and we observe considerable differences
t + , respectively (Prosperetti, 2017). The normalization given by between the coupled and the uncoupled solutions even at early
Eq. (20) was originally introduced by Mikic et al. (1970) and covers stages, where the bubble radii still feature very similar values. The
the inertia and heat diffusion driven stages. It uses several assump- mass flux is determined as m˙  = ρvap R˙ . Without heat diffusion the
tions and empirical correlations which may introduce appreciable mass flux increases and remains at a high level. At the same time,
uncertainties. However, Eq. (20) compares suprisingly well with the pressure in the bubble and the vapour density remain con-
multiple experimental and numerical studies (Prosperetti, 2017; stant. If we consider evaporative cooling (and thereby heat diffu-
Avdeev and Zudin, 2005). In the context of the current study, these sion), the mass flux decreases after an early peak, the pressure in
correlations provide first estimates of the relevant bubble growth the bubble decreases and the vapour density decreases as well. In
stages and the asymptotic limits for flash boiling of cryogenic oxy- the inertia controlled range, the growth rate of the bubble (R˙ ), is
gen. Other analytical solutions coupling the different stages of bub- not affected by the different mass fluxes because the vapour den-
ble growth are presented, among others, by Theofanous and Pa- sity changes for the different approximations. Therefore, the volu-
tel (1976), Prosperetti and Plesset (1978) as well as Avdeev and metric flow rate V˙ = A · m˙  /ρvap (with A being the bubble surface)
Zudin (2005), but are not further discussed here. is similar for the two approaches in the inertia controlled regime.
The The t + − R+ relationship according to Mikic et al. (1970) is We conclude that evaporative cooling affects the bubble pressure
demonstrated in Fig. 5. Here, the reference solutions for sin- and the mass flux throughout the entire growth process, but the
gle vapour bubble growth have been normalized as introduced growth rate at later stages only. In the DG solver, the bubble pres-
by Eq. (20) and the dashed and dash-dotted lines indicate the sure will also drop despite the omission of heat conduction and
asymptotic limits. It is apparent that inertia driven growth holds subsequent temperature decrease at the interface since the bub-
for times smaller than t + = 10−2 only, while diffusion controlled ble is fully resolved and surface effects become small as the radius
growth requires growth for time scales longer than t + = 50. The increases. The mass flux of the DG solver should therefore not be
intermediate range is defined from 10−2 < t + < 50. In Fig. 5 we calibrated towards the constant mass flux derived from the RPE
show the cases C5-90, C5-120, C100-90 and C100-120, which cover solution without heat diffusion, as this strongly overpredicted the
a wide spectrum of Ja including the highest and the lowest degree growth rate. We conclude that evaporative cooling and heat diffu-
of superheat within our parameter range. We restrict the maxi- sion are required to predict all physical quantities accurately, even
mum physical simulation time of our investigation to t = 100 μs for the inertia driven stage. In Section 2 we have discussed the
D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269 257

Fig. 7. Results obtained from the RPE coupled to the temperature equation for different temperatures (colours) at R p = 5.

challenges in resolving the temperature boundary layers, in par- 4.3. Numerical setup
ticular regarding our forthcoming work with multiple bubbles. To
keep the computational effort reasonable, we continue using the The spherically symmetric reference solution allows for the use
Euler equations and we calibrate our vaporisation models such that of identical symmetry conditions for the DG solver, and the source
the growth rates of the bubbles are matched for particular time in- term (Toro, 2009)
tervals. We assess the validity of this approach in Section 5.1. 

2
ρu
Fig. 6also indicates that the relevant length scales of bubble
S= ρ u2 (23)
growth easily cover three orders of magnitude. For different phys- r
u (E + p )
ical conditions these ranges may even be larger for the same time
interval. This is shown in Fig. 7(a) where we depict the physical is added on the right hand side of Eq. (1) reverting the 3D sys-
bubble radii at R p = 5 for the different temperatures. The bubble tem to a 1D configuration with spherical symmetry. Fig. 8 depicts
grows by four orders of magnitude for T0 = 120 K which is impos- the geometrical setup where we set the bubble in the left part of
sible to resolve in full 3D computations. As the volumetric jet ex- the domain. The left boundary at r = 0 indicates the bubble centre
pansion and jet break-up are key quantities of interest, it is fair and we apply a symmetry condition here. The superheated liquid
to stipulate that the early stages of growth are not relevant. In fills the right side with an outflow boundary condition at r = L. We
Fig. 7(b) we depict the total vaporised mass (solid lines) and the keep pressure wave reflections small by using a weak outflow con-
mass flux (dashed lines) as functions of the bubble radius R. The dition and a large domain size L. In the weak outflow condition the
mass flux peaks at low R but the total vaporised mass accumulates Riemann problem is solved for the initial state of the liquid and the
during the final stages, due to its R2 -dependence. Thus, internal state at the boundary. The default domain size satisfies the
condition
• it is valid to assume that early bubble growth can be neglected tend cliq
L= . (24)
and that 2
• the initial bubble radius for the DG simulation can be increased ensuring that a pressure wave can travel only one time forth and
to R0 = 0.1Rend . The computations would then cover 99.9% of back through the domain. A stretched computational grid is de-
the volume expansion of a single bubble growing from Rcrit to fined by
Rend .
1 − C rref
C = f · NDG , r ∗ = and rnode = L · r ∗ (25)
We introduce a normalized radius 1 −C
determining the radial position rnode of a polynomial node. Here,
R f, NDG and rref are an adjustable stretching coefficient, the num-
R∗ = (21)
Rcrit ber of discontinuous Galerkin elements and the node coordinate in
an equidistant reference domain ranging from 0 to 1, respectively.
and use the initial radii
The degree of the polynomial is kept constant for all simulations
R0 with N = 3. The default stretching factor is f = 1.03 and the num-
R∗0 = = {1, 5, 10, 25, 50} (22) ber of DG elements NDG is set such that a bubble is initially re-
Rcrit
solved with at least ten discretisation points. The numerical test in
for each combination of T∞ and Rp . The reference solution defines the present section are accomplished using the Hertz–Knudsen re-
our final simulation time for each combination of R∗0 , T∞ and Rp lation. The model coefficients are calibrated against the reference
with tend = f (Rend = 10R0 ). When we initialize the DG simulation solution from Section 4.2 to λvap = 0.038 and λcond = 0.0376, re-
with R∗0 = 1 we increase the start radius by one FV sub-cell (i.e. spectively.
one discretisation point) to trigger the growth of the bubble. Note The preceding section discussed the condition for the initial ra-
that we always start from the critical radius in our reference solu- dius according to Eq. (22). The increase in initial radius reduces
tion. In summary, our test matrix is defined by three independent the computational cost compared to R0 = Rcrit for two reasons:
variables. Two of these parameters, T0 and Rp , define the physical firstly, for a fixed number of cells across the bubble radius the to-
conditions and the third parameter, R∗0 , defines the range of the tal number of computational cells reduces. Secondly, the required
length and time scales scales that are covered during the simula- time step increases with the cell width and therefore with R0 if,
tion time. again, the number of cells across the bubble radius is constant. For
258 D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269

Fig. 8. Geometrical setup for single bubble growth.

Table 2 Table 3
Relative computational time for different R∗0 in percent. Parameters for domain size sensitivity
study. The default length L is given by
R∗0 = 5 R∗0 = 10 R∗0 = 25 R∗0 = 50 Eq. (24).
tCPU /tCPU,Rcrit 15% 7% 2% 1%
L in [mm] f [–] DG cells

L/2 0.165 1.029 100


L 0.33 1.03 120
a given final simulation time, the total number of time steps re- 2L 0.66 1.031 140
duces with increasing time step size. The total computational time
is proportional to the number of time steps and the number of
cells. The computational time relative to the computational time Table 4
Parameters for cell size sensi-
with R0 = Rcrit reads
tivity study.
tCPU tRcrit NDG
= . (26) Nbub f [–] DG cells
tCPU,Rcrit t NDG,Rcrit 5 1.062 60
Table 2 lists the results for R∗0 = 5 . . . 50. They do not strongly de- 10 1.03 120
20 1.015 240
pend on T0 nor on Rp . The meshes for the present analysis were
constructed such that the number of cells across the bubble radius
and the cell width approaching the right end of the domain (cf.
Fig. 8) are constant for R0 = Rcrit and R0 = R∗0 · Rcrit , respectively. at the final time and an increase in p on the liquid side is observed
The computational time reduces to at least 15% compared to the for L/2 compared to L and 2L. The differences arise due to weak
cases with R0 = Rcrit . At R∗0 = 50 the computational time reduces shock wave reflections at the outflow boundary condition. The de-
to 1% emphasising the benefit of using Eq. (22). viation in the pressure field does not affect the bubble growth but
using the domain size L in Eq. (24) guarantees mesh independence
4.4. Mesh resolution and time step restriction throughout the investigation, and L is used for the remainder of
the paper. Fig. 10 depicts the same quantities for a varying reso-
A grid dependency study has been conducted to ensure inde- lution of the initial bubble with Nbub = [5, 10, 20]. Table 4 lists the
pendence of results on the major numerical parameters. These in- corresponding mesh parameters. The three cases are in very good
clude the domain size, L, the bubble resolution, i.e. the grid size, agreement indicating mesh independent results for the given crite-
and the time step that is here expressed with the aid of the ria. Fig. 10 also illustrates results when using a reduced time step
CFL number. The influence of these parameters is studied for C5- (here, the CFL number is set to 0.45). Again, no significant change
120 with R∗0 = 50. First, the domain size is varied to assess the in the growth rates and in the resulting pressure field is observed.
influence of wave reflections from the boundary: L is given by When comparing the different numerical setups, the bubble radii
Eq. (24) and increased and decreased by a factor of 2 for the do- deviate by a maximum of 2.2%, 1.4% and 0.9% for cases with a vari-
main size sensitivity study. The specific parameters are given in able cell width, time step size and domain length, respectively. It
Table 3. Fig. 9(a) depicts the bubble radius over time which is not is noted here (1) that - similar to R - the numerical parameters do
sensitive to the domain size. Diagram 9(b) shows the pressure field not unduly affect the evolution of m˙  and V and (2) that very

Fig. 9. Results of C5-120 with different domain sizes L for Nbub = 10.
D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269 259

Fig. 10. Results of C5-120 with different bubble resolutions Nbub and with CF L = 0.45 for L = L2 .

similar mesh sensitivities have been observed when using the ki-
netic relation (KR) and the sub-grid scale heat flux model (SHFM)
as a vaporisation model. Mesh independence of all results reported
below can be assumed.

5. Results and discussion

5.1. Hertz–Knudsen relation

A direct calibration of the HKR is difficult because two model


coefficients appear in Eq. (6). Setting λcond = λvap reduces the ac-
curacy of the model but simplifies the estimation of the remain-
ing model coefficient. Lauer et al. (2012) calibrated this single co-
efficient according to a reference solution for the simulation of
Fig. 11. Normalized bubble radius R∗ over time.
cavitation bubbles. We use a similar strategy but suggest λcond =
0.99λvap to stabilize the simulations. At the initial time steps,
when the solution deviates only slightly from mechanical and ther-
modynamic equilibrium, vaporisation is favoured and early bubble
growth is stabilized. Here, we are not interested in bubble collapse
nor bubble oscillations. At later time steps, the proposed reduction
in the condensation rate does not change the results. The condition
λcond = 0.99λvap holds throughout the paper but we refer only to
λvap in the subsequent discussion. In the following, we compare
the DG results to the reference solution and we start our discus-
sion for case C5-120 with R∗0 = 50.

C5-120, R∗0 = 50
The base case of our analysis is C5-120 with R∗0 = 50. The black
line in Fig. 11 depicts the normalized bubble radius R∗ over time
obtained from the DG solver while the red line shows the corre- Fig. 12. Vaporisation mass flux m˙  of case C5-120 with R∗0 = 50.
sponding reference solution. The DG solution agrees well with the
reference solution for the relevant time and length scales. The de-
viations are large only for t/tend < 0.2 because the initial radius for
the DG solution had been increased. The corresponding vaporisa-
tion coefficient λvap = 0.038 scales the mass flux towards m˙  ≈ 65
kg/(m2 s), see Fig. 12. The vaporisation mass flux converges to a
constant value because the liquid remains superheated in the HLLP
approximation. Initially, the vapour in the bubble is saturated and
Eq. (6) gives a zero mass flux. We are not in mechanical equi-
librium, the bubble grows and the pressure in the bubble drops.
With decreasing pressure the mass flux increases and the pres-
sure drop is slowed down. After a short time the bubble pres-
sure converges towards the ambient pressure and a steady state
is reached for the bubble pressure and the mass flux. Fig. 13 de-
picts the pressure field at different time steps. We observe that the
bubble pressure has approximately reached the ambient level af-
ter t/tend = 0.08. After that time, the bubble pressure and the mass Fig. 13. Pressure field of C5-120 and R∗0 = 50 at different time steps.
260 D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269

Fig. 14. Integrated quantities of C5-120 and R∗0 = 50.

25% if we start from the critical radius R∗0 = 1. The generated vol-
ume matches well the reference solution only close to tend . In con-
trast to that, we observe a very good agreement for both mtot and
V starting from R∗0 = 10. While we predict mtot differently well
for the combinations of R∗0 , Rp and T0 we obtain very good matches
in V for all cases with minor restrictions for R∗0 = 1.

Variation in the physical conditions


We conduct the next simulations with R∗0 = 50 and vary the su-
perheat ratio Rp and the temperature T0 . Fig. 17 shows the volu-
metric expansion for the cases C100-120 and C5-100, which agree
well with the reference solution. We obtain similar results for the
various combinations of Rp , T0 and R∗0 . Regarding the simulation of
single vapour bubble growth we conclude that
Fig. 15. Normalized bubble radius R∗ over time for different length scales. • the simplifying assumptions discussed in Section 4.3 are valid
in most of our cases,
flux remain unchanged. The qualitative behaviour of m˙  in the ref- • we can predict the bubble growth in a wide parameter range
erence solution cannot be matched but the integrated mass flux using the HLLP Riemann solver combined to the HKR,
depicted in Fig. 14(a) deviates by only 10%. We have calibrated the • the instantaneous mass flux cannot be predicted with the HKR
vaporisation coefficients such that we match the final bubble ra- within the HLLP Riemann solver.
dius. The total vaporised mass deviates from the reference solution
because the vapour densities are not exactly the same as in the 5.2. Kinetic relation and sub-grid scale heat flux
reference solution and the same volume contains a different mass.
Our main focus is the volume expansion of the bubbles. Fig. 14(b) In this section, we compare the two alternative vaporisation
shows the total change in the bubble volume where the deviations models to the HKR. It is first noted that all findings discussed in
are smaller. Thus, the HLLP Riemann solver coupled to the HKR is Figs. 15–17 equally hold for the KR and SHFM closures but are
a suitable approach to predict the volumetric expansion of single not repeated here for brevity of presentation. Fig. 18 compares
vapour bubbles. Note that we can improve the results for total va- the vaporisation mass fluxes, the bubble radii and the volumetric
porised mass adopting the calibration to the target quantity mtot . expansion. The red and black dashed lines show results for the
However, we have found that the calibration towards mtot is more KR and the SHFM, respectively. The peak mass flux (Fig 18(a)) is
difficult to obtain, particularly at low R∗0 . Given the good match in slightly higher for the KR when compared with the HKR (black
V we calibrate the vaporisation models towards R in the remain- solid line), but – as the evaporative cooling in the HLLP approach
ing investigation. The calibration of λvap is specific for the chosen is not resolved – it rapidly asymptotes towards a constant value.
time scale. In the next step, we analyse the validity of our ap- The asymptotes of the HKR and KR deviate by approximately 5%.
proach for the same physical conditions, i.e. constant T0 and Rp , Particularly at early times m˙  fluctuates but the volumetric ex-
at different time intervals implicitly defined by R∗0 (cf. Section 4.2). pansion is nearly identical to the HKR approximation (Fig. 18(a)).
Similarly, trends observed for different initial radii R∗0 and for
Variation of the initial radius R∗0 different physical conditions follow trends described for the HKR
We decrease R∗0 towards the critical radius R∗0 = 1 keeping the closure. The black dashed lines in Fig. 18 represent the SHFM.
time interval of each simulation determined by R∗end = 10R∗0 . We Here, the mass flux agrees qualitatively better with the reference
depict the bubble radii as a function of time in Fig. 15 where each solution than the HKR and the KR solutions. We observe a peak
individual calibration provides a good match to the reference so- during the initial stages and a continuous decrease for larger time
lution for the final simulation time. The growth at the relevant intervals. This results from the dependency of the vaporisation
length and time scales matches the reference solution well as seen model on the continuously growing boundary layer thickness as
for R∗0 = 50 in Fig. 11. Fig. 16 depicts the total vaporised mass and it is defined by Eq. (10). However, we underpredict the reference
the total volume change as a function of time for C5-120 with mass flux for the entire time interval. In contrast, the volumetric
R∗0 = 1 and R∗0 = 10. We observe a deviation in mtot of more than expansion is overpredicted because the increased radius R∗0 = 50
D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269 261

Fig. 16. Total vaporised mass and volume change of C5-120 for different R∗0 .

Fig. 17. Volumetric expansion for R∗0 = 50 under different physical conditions.

Fig. 18. Comparison of HKR, KR and SHFM for case C5-120 and R∗0 = 50.
262 D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269

Fig. 19. Vaporisation coefficients λvap as a function of different characteristic quantities of our investigation.

results in a larger bubble surface which allows for a faster volume Fig. 19 depicts the vaporisation coefficient λvap as function of
expansion of the bubble (Fig. 18(a)). At the end of the simulation R∗0 , Rp , T0 , Ja, R+ or t+ (Fig. 19(a)–(f), respectively). For clarity of
time all vaporisation models predict the volumetric expansion presentation, the diagrams show the results from T = 90 K and T
well. A better estimation of the boundary layer thickness δ can = 120 K only, but trends for the intermediate values reveal similar
improve the prediction of the volumetric expansion of the SHFM trends and are therefore omitted from the figures.
for the early stages. This could be achieved by a dynamic deter-
mination of Cδ but further research would be required to find • Dependence on R∗0 (Fig. 19(a)): For fixed physical conditions, λvap
suitable models and is beyond the scope of the current work. scales inversely proportional with R∗0 . For the reference case,
the mass flux decreases monotonically with time/radius after
5.3. Prediction of model coefficients its early peak (cf. Fig. 12). The combination of the DNS solver
with the HKR closure yields – after a short transient stage at
In the previous sections we have demonstrated that our simu- the beginning of a simulation – a constant mass flux. Depend-
lation strategy is valid for a wide range of parameters. The tables ing on the growth stages that are covered between R∗0 and R∗end ,
in Appendix A list the corresponding calibration coefficients. The λvap scales this constant mass flux such that the mean value of
main disadvantage of this approach is the need to iteratively cali- the reference solution is approached. At small R∗0 the depen-
brate the model coefficients for each combination of the initial pa- dency can be reversed as the growth rate does not only de-
rameters T0 , Rp and R∗0 . A truly predictive simulation requires the pend on the (highly transient) mass flux but also on the intertia
modelling of the calibration coefficients and in this section, we dis- forces.
cuss the relation of the model coefficients to the characteristic pa- • Dependence on Rp (Fig. 19(b)): For small superheat ratios, no
rameters of our study. obvious relation between Rp and λvap is found. The solid and
D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269 263

Fig. 20. Comparison of estimated and calibrated λvap .

dash-dotted lines in Fig. 19(b) represent cases with R∗0 = 5 and


R∗0 = 50, respectively. While λvap decreases with increasing Rp
at low R∗0 , it increases with increasing Rp at high R∗0 . The ref-
erence growth rate increases with increasing Rp at later time
intervals. Therefore, λvap also increases with increasing Rp in
the range of large R∗0 . At smaller R∗0 the reference growth rate
also increases with increasing Rp but the differences are less
pronounced. The inertia forces strongly affect the early growth
stage, and λvap can therefore not follow the trend observed for
higher R∗0 .
• Dependence on T0 (Fig. 19(c)): The calibrated coefficients λvap
decrease with decreasing T0 if R∗0 and Rp are kept constant. As
not only temperature changes but also the saturation pressure
and - for a constant Rp - the ambient pressure are functions
of temperature, and all these quantities affect the HKR (see Fig. 21. Vaporisation mass fluxes normalized by the vapour density and the factor
Eq. (6)), the correlation between λvap and T0 is difficult to as- A in Eq. (20) as a function of the dimensionless time t + . After an initial transient
sess. The driving pressure difference in Eq. (6) and therefore the the curves for all cases approximately converge to a single curve.
resulting mass flux typically decrease for smaller temperatures.
Here, we observe that the calibration of λvap also reduces the
is smaller than one. The opposite applies for large R∗0 and Fig. 20(b)
mass flux for lower temperatures.
confirms this theory. The red line in Fig. 20(a) is a least-squares fit
• Dependencies on Ja, R+ and t + (Fig. 19(a)–(f)): The coefficient
of the data points, which can be used to approximate λvap,pre based
λvap does not reveal any simple dependencies on either of the
on the initial estimation λvap,RPE with
characteristic numbers Ja, R+ and t + . Trends are observed if Rp
and R∗0 are fixed (Fig. 19(a)) or if Rp and T0 are fixed (Fig. 19(f)) λvap,pre = λvap,RPE · 1.27 · (1 − exp(−R∗0 · 0.086 ))0.61 . (28)
but in general the characteristic number can spread over sev-
eral orders of magnitude for similar vaporisation coefficients. The largest deviation between λvap,cal and λvap,pre occurs for R p =
100, T0 = 120 K and R∗0 = 5. The predicted coefficient and the re-
It is thus important to note here that there is no apparent cor- sulting volume expansion differ by approximately 45%. Fig. 20(c)
relation of λvap with any of the characteristic quantities that would summarizes these deviations for all test conditions. The error tends
allow for an a priori estimate of the model parameter and alter- to decrease for increasing initial radii and Fig. 21 gives one pos-
native procedures for modelling need to be sought. Similar find- sible reason for this trend. Here, the mass fluxes are normalized
ings are made for the KR and SHFM closures and are presented in with the coefficient A of Eq. (20) and with the density of the sat-
Appendix B. urated vapour in the bubble. The black lines show solutions from
A possible alternative modelling procedure for the calibration the RPE which converge after an initial highly transient stage into a
coefficients shall now be based on the reference solution. Since the single curve. The slope of this curve is approximately proportional
results follow similar trends for the different vaporisation mod- to (t + )−0.5 and remains roughly constant within the time interval
els we restrict this analysis to the HKR. With λcond = 0.99λvap , considered. The modelling of the early transients is rather compli-
Eq. (6) can be rearranged to yield the vaporisation coefficient as cated with the present method but the volumetric expansion (that
√ is our target quantity) is small for these stages and the accuracy
m˙  2π R
λvap,RPE = psat . (27) improves for the relaxed stages.
√ − 0.99 √pvap
Tliq Tvap
5.4. Sensitivity analysis for the model coefficients and DNS results
We obtain the mass flux m˙  , the vapour pressure pvap and the
vapour temperature Tvap from our reference solution at R = Rend . The preceding sections discussed the dependency of the vapor-
The saturation pressure psat is determined as a function of the isation model coefficients on the physical initial conditions and the
liquid temperature Tliq = T∞ . Fig. 20(a) depicts the ratio between initial bubble radii. This section discusses the sensitivity of the DNS
the calibrated coefficients λvap,cal (Appendix A) and the coefficients results to changes in the model coefficients for fixed initial condi-
λvap,RPE from Eq. (27). The estimates λvap,RPE differ from the cali- tions.
brated coefficients, but the deviations are similar for constant R∗0 . Fig. 22 depicts the bubble radius as a function of time for dif-
For small R∗0 the mass flux in the RPE is larger than the mass flux ferent physical parameters and vaporisation models. The red and
in the DG solution. As a result λvap,RPE is larger and λvap,cal /λvap,RPE black lines represent the calibrated results and results with a 10%
264 D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269

Fig. 22. Sensitivity of the bubble radius to a 10% change in the model coefficient. All cases use R∗0 = 50.

change in the model coefficient, respectively. In all cases, an in-


crease or decrease in the coefficient yields faster or slower bubble
growth. The maximum difference in the bubble radii between the
calibrated and the disturbed solutions is typically smaller than the
disturbance (usually about 2–5%, only in a specific case equal to
the disturbance) as the density of the fresh vapour also changes.
An increase in the mass flux as a result of a change in the coef-
ficient yields an increased density and the volume expansion re-
duces. The inverse effect is observed if the mass flux decreases.
The results show that a calibration of the vaporisation models is
needed for an accurate predictions of the bubble expansion but the
sensitivity to the exact value is moderate as a dampening rather
than an enhancement of a disturbance from the optimum value is
observed.
Fig. 23. Normalized bubble radius R∗ over time for C5-120 with R∗0 = 50. The DG
simulation with the 1D solver (black line) and the 3D solver (blue crosses) match
5.5. 3D simulation of single bubble growth
the reference solution (red line). (For interpretation of the references to colour in
this figure legend, the reader is referred to the web version of this article.)
As a final test case, we demonstrate the capabilities of the DG
solver in a truly three-dimensional system. Here we use case C5-
120 with an initial radius of R∗0 = 50. The HKR is taken for the clo- have discussed this effect in Sections 4.2 and 5.1. The right plane
sure of the interface mass transfer rate and the results from the 3D of the subfigures in Fig. 24 shows the characteristic velocity field
computation are compared to the 1D computation and to the refer- for single vapour bubble growth: the velocity peaks at the inter-
ence solution. Fig. 23 depicts the dimensionless bubble radius over face in the liquid phase and decreases towards the outside due
time. The three-dimensional case agrees very well with the one- to the spherical symmetry of the problem. The density fields in
dimensional simulation and with the reference solution. We also the bottom planes demonstrates that the sharpness of the inter-
show a three-dimensional representation of the bubble growth in face is guaranteed. Last but not least, mass and energy conser-
Fig. 24. The four subfigures from the top left to the bottom right vation needs to be ensured due to the non-conservative charac-
depict different simulation times ranging results from t = 0.0 μs to ter of the level-set approach. Fig. 25 illustrates the relative change
t = 1.0 μs. The left plane of each subfigure illustrates the pressure of the total mass and the total energy throughout the simula-
field. We observe a pressure wave propagating away from the in- tion. Both increase by less than 1% demonstrating that the present
terface as the bubble grows. Note, that the color bar scales with three-dimensional solver reasonably conserves mass and energy.
a maximum pressure of p = 3 bar which indicates that the global The results show that the DG solver developed here provides a
pressure level rapidly relaxes towards the ambient pressure. We useful tool that can be used in future studies for the computation
D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269 265

Fig. 24. Three-dimensional representation of the bubble growth.

isation mass flux. The first model is the Hertz–Knudsen relation


(HKR) which is based on the difference between saturation pres-
sure and vapour pressure. In the second approach we apply the
kinetic relation (KR) according to the original formulation of the
HLLP Riemann solver. It estimates the mass flux depending on the
deviation of the local entropy to the saturation entropy. The last
approach determines the vaporisation mass flux based on a sub-
grid scale heat flux and is denoted as sub-grid heat flux model
(SHFM). Here we have assumed a parabolic temperature profile in
the liquid within the boundary layer and have derived the tem-
perature gradient from the first derivative at the interface. These
models are calibrated to match the bubble radius of a reference
solution for wide pressure, temperature and time ranges as the ul-
timate objective of the DG code is its use to predict spray break-up
under flash boiling conditions. The operating conditions are char-
Fig. 25. Mass and energy conservation. acteristic for superheated cryogenic oxygen issued in combustion
chambers near vacuum.
We have observed similar results using the HKR and the KR.
of multi-bubble dynamics and their interactions leading to spray The instantaneous character of the reference mass flux is not cap-
break-up under flashing conditions. tured but the volumetric expansion of bubbles and the total va-
porised mass are well predicted. We have found uncertainties only
6. Summary and outlook at the smallest time scales. Typical length and time scales vary a
lot within our parametric study and the calibrated model coeffi-
The goal of the present investigation was the simulation of sin- cients change accordingly. For an increasing temperature the cal-
gle vapour bubble growth with a fully compressible two-phase ibration promotes vaporisation but a general dependency on the
solver. We have used the HLLP Riemann solver coupled to a discon- time scales or pressure ranges is not obvious. The SHFM qualita-
tinuous Galerkin method and a third order Runge–Kutta scheme to tively matches the reference solution for the mass flux well. How-
discretise the Euler equations in space and time, respectively. The ever, the bubble radius and the volumetric expansion match to the
solver conserves a sharp interface between the liquid and vapour. reference case only in the last part of a simulation. A more sophis-
We have compared three different models to estimate the vapor- ticated representation of the model coefficient may improve the
266 D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269

results for the volumetric expansion and gives potential for an ex- Acknowledgements
plicit model without a-priori calibration.
Generally, we have shown that the HLLP Riemann solver can be The authors acknowledge the financial support by the Deutsche
used for the numerical simulation of single vapour bubble growth. Forschungsgemeinschaft (DFG) as part of the Collaborative Re-
A coupling to different vaporisation models and a suitable calibra- search Center SFB TRR 75 “Droplet dynamics under extreme ambi-
tion of model coefficients can be based on simple a-priori compu- ent conditions” held jointly by the University of Stuttgart and the
tations using a reference case and the DG solver will allow for a University of Technology Darmstadt.
prediction of the volumetric bubble expansion. The solver is stable
in the most cases we have investigated even with high pressure Appendix A. Coefficients of the different vaporisation models
ratios between vapour and liquid. This model will be used in fu-
ture investigation for the simulation of multiple bubble growth in This section lists the calibrated coefficients for the HKR, the KR
superheated liquid jets. and the SHFM for each combination of T0 , Rp and R∗0 . If no value is
given, no stable solution was obtained.

Table A.1
λvap for R p = 5 and R p = 10.
R∗0 C5-90 C5-100 C5-110 C5-120 C10-90 C10-100 C10-110 C10-120

1 0.050 0.071 0.093 0.110 0.045 0.064 0.084 0.093


5 0.056 0.071 0.084 0.093 0.052 0.065 0.080 0.083
10 0.056 0.064 0.075 0.083 0.055 0.065 0.076 0.082
25 0.042 0.045 0.050 0.057 0.047 0.052 0.058 0.063
50 0.030 0.031 0.033 0.038 0.035 0.037 0.042 0.046

Table A.2
λvap for R p = 50 and R p = 100.
R∗0 C50-90 C50-100 C50-110 C50-120 C100-90 C100-100 C100-110 C100-120

1 – 0.058 0.075 0.082 – – 0.073 0.081


5 0.053 0.064 0.072 0.074 0.052 0.058 0.071 0.073
10 0.054 0.065 0.074 0.076 0.054 0.059 0.073 0.075
25 0.050 0.055 0.059 0.064 0.051 0.054 0.060 0.064
50 0.039 0.041 0.045 0.049 0.039 0.041 0.045 0.049

Table A.3
cent for R p = 5 and R p = 10.

R∗0 C5-90 C5-100 C5-110 C5-120 C10-90 C10-100 C10-110 C10-120

1 3793.1 907.02 278.16 119.15 4278.8 1012.0 333.72 139.34


5 4447.5 1114.60 424.96 193.12 4844.3 1340.4 476.09 231.92
10 4529.4 1250.74 495.66 224.82 4796.2 1375.0 503.63 256.33
25 5611.6 1886.19 747.02 350.78 5806.8 1881.2 724.47 340.54
50 8185.8 2746.51 1207.39 553.95 8152.1 2554.7 1085.46 507.25

Table A.4
cent for R p = 50 and R p = 100.

R∗0 C50-90 C50-100 C50-110 C50-120 C100-90 C100-100 C100-110 C100-120

1 – – 374.95 – – – 386.42 –
5 5261.9 1480.5 555.79 296.51 5369.6 1659.3 562.68 303.33
10 5193.0 1467.8 576.02 288.19 5159.7 1639.1 585.23 294.85
25 5683.7 1825.8 721.44 368.89 5627.7 1883.7 768.68 372.18
50 7667.4 2603.3 1032.56 524.86 7684.8 2612.0 1036.04 531.14

Table A.5
Cδ for R p = 5 and R p = 10.

R∗0 C5-90 C5-100 C5-110 C5-120 C10-90 C10-100 C10-110 C10-120

1 0.993 1.074 1.288 1.826 1.367 1.495 1.774 2.620


5 0.175 0.221 0.305 0.437 0.218 0.281 0.400 0.607
10 0.088 0.121 0.174 0.248 0.103 0.141 0.208 0.307
25 0.050 0.074 0.102 0.139 0.052 0.081 0.113 0.162
50 0.037 0.055 0.076 0.103 0.039 0.057 0.080 0.113
D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269 267

Table A.6
Cδ for R p = 50 and R p = 100.

R∗0 C50-90 C50-100 C50-110 C50-120 C100-90 C100-100 C100-110 C100-120

1 1.750 1.942 2.367 3.524 1.801 2.336 2.454 3.640


5 0.289 0.374 0.514 0.810 0.299 0.419 0.525 0.845
10 0.130 0.178 0.246 0.389 0.130 0.190 0.255 0.397
25 0.062 0.089 0.127 0.190 0.062 0.092 0.128 0.193
50 0.041 0.061 0.087 0.129 0.042 0.062 0.088 0.132

Appendix B. Dependency of cent and Cδ on initial conditions decreases. As discussed in Section 5.3 this results from the tran-
and characteristic numbers sient decrease of the reference mass flux with time/radius.
• Dependency on Rp (Fig. B.1(b)): The model coefficient depends
Fig. B.1 depicts the dependencies of the coefficient cent in the only weakly on the superheat ratio Rp . Again, the trends and
KR on R∗0 , Rp , T0 , Ja, R+ and t + . Eq. (7) shows that the mass flux conclusions are similar to those of the HKR in Section 5.3 con-
scales inversely proportional to cent and similar trends as observed sidering the inverse proportionality of the mass flux on cent .
for the HKR would be expected for 1/cent . • Dependency on T0 (Fig. B.1(c)): We observe a decrease of the
coefficient with increasing temperature if R∗0 and Rp are fixed.
• Dependency of cent on R∗0 (Fig. B.1(a)): With increasing R∗0 , the
As observed for the HKR, the model coefficient scales such that
model coefficient increases such that the asymptotic mass flux

Fig. B.1. Vaporisation coefficients cent as a function of different characteristic quantities of our investigation.
268 D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269

Fig. B.2. Vaporisation coefficients Cδ as a function of different characteristic quantities of our investigation.

vaporisation is enhanced for an increase in temperature. How- and T0 , respectively. This is particularly true for small R∗0 (solid
ever, the entropy as the key quantity in Eq. (7) also changes lines). Both, for increasing T0 and increasing Rp , the temper-
with T0 and a direct correlation is not apparent. ature difference T = [Tliq − Tsat ( p∞ )] increases. According to
Fig. 13, the pressure in the bubble rapidly approaches the am-
Fig. B.2 shows the resulting correlations for the SHFM. In gen- bient pressure such that T = [Tliq − Tsat ( p∞ )] reasonably ap-
eral, an increase in Cδ yields a decrease in the mass flux. However, proximates the temperature difference in Eqs. (9). Thus, an in-
the SHFM yields a similar transient behaviour as the reference so- crease in T0 and Rp would result in an increase of m˙  . At
lution and the correlations of Cδ in Fig. B.2 deviate from those ob- the same time, the bubble radius and accordingly the bound-
tained for the HKR and the KR, respectively. This is a consequence ary layer estimation decrease with increasing T0 and Rp which,
of the R-dependency of the mass flux given by Eqs. (8)–(10). again, allows for higher mass fluxes. The increase of Cδ avoids
nonphysically high mass fluxes.
• Dependency of Cδ on R∗0 (Fig. B.2(a)):Eqs. (9) and (10) show that
the mass flux decreases with increasing R∗0 . The high coeffi- Both, the KR and the SHFM do not reveal any clear functional
cients observed in Fig. B.2(a) at small R∗0 avoid an overpredic- dependence of the coefficients on the characteristic parameters R+ ,
tion of the mass flux when the radii and approximated bound- t + and Ja.
ary layer thicknesses are rather small.
References
• Dependencies on Rp and T0 (Fig. B.2(b) and B.2(c)): The trends
Cδ = f(R p ) and Cδ = f(T0 ) are explained by the same physical Avdeev, A., Zudin, Y., 2005. Inertial-thermal governed vapor bubble growth in highly
effects. Fig. B.2(b) and B.2(c) show that Cδ increases with Rp superheated liquid. Heat Mass Tranf. 41, 855–863.
D. Dietzel, T. Hitz and C.-D. Munz et al. / International Journal of Multiphase Flow 116 (2019) 250–269 269

Bar-Kohany, T., Levy, M., 2016. State-of-the-art review of flash-boiling atomization. Lee, J., Son, G., 2017. A sharp-interface level-set method for compressible bubble
Atom. Sprays 26 (12), 1259–1305. growth with phase change. Int. Commun. Heat Mass Transf. 86, 1–11.
Bell, I., Wronski, J., Quoilin, S., Lemort, V., 2014. Pure and pseudo-pure fluid ther- Manfletti, C., 2014. Laser ignition of an experimental cryogenic reaction and control
mophysical property evaluation and the open-source thermophysical property thruster: pre-ignition conditions. J. Propul. Power 30 (4), 925–933.
library coolprop. Ind. Eng. Chem. Res. 53, 2498–2508. Mikic, B., Rohsenow, W., Griffith, P., 1970. On bubble growth rates. Int. J. Heat Mass
Brennen, C.E., 1995. Cavitation and Bubble Dynamics. Oxford University Press. Transf. 13 (4), 657–666.
Chang, C.-H., Deng, X., Theofanous, T.G., 2013. Direct numerical simulation of inter- Persad, A.H., Ward, C.A., 2016. Expressions for the evaporation and condensation
facial instabilities: a consistent, conservative, all-speed, sharp interface method. coefficients in the Hertz–Knudsen relation. Chem. Rev. 116, 7727–7767.
J. Comput. Phys. 242, 946–990. Plesset, M., 1949. The dynamics of cavitation bubbles. J. Appl. Mech. 16, 277–282.
Fechter, S., Jaegle, F., Schleper, V., 2013. Exact and approximate Riemann solvers at Plesset, M., Zwick, S., 1954. The growth of vapor bubbles in superheated liquids. J.
phase boundaries. Comput. Fluids 75, 112–126. Appl. Phys. 25 (4), 493–500.
Fechter, S., Munz, C., Rohde, C., Zeiler, C., 2017. A sharp interface method for com- Plesset, M.S., Properetti, A., 1977. Bubble dynamics and cavitation. Annu. Rev. Fluid
pressible liquid-vapor flow with phase transition and surface tension. J. Comput. Mech. 9, 145–185.
Phys. 336, 347–374. Prosperetti, A., 2017. Vapor bubbles. Annu. Rev. Fluid Mech. 49, 221–248.
Fechter, S., Munz, C., Rohde, C., Zeiler, C., 2018. Approximative Riemann solver for Prosperetti, A., Plesset, M.S., 1978. Vapour-bubble growth in a superheted liquid. J.
compressible liquid vapor flow with phase transition and surface tension. Com- Fluid Mech. 85, 349–368.
put. Fluids 169, 169–185. Rayleigh, L., 1917. On the pressure developed in a liquid during the collapse of a
Fechter, S., Munz, C.D., 2015. A discontinuous Galerkin-based sharp-interface spherical cavity. Philos. Mag. J. Sci. 34, 94–98.
method to simulate three-dimensional compressible two-phase flow. Int. J. Nu- Robinson, A., Judd, R., 2004. The dynamics of sperical bubble growth. Int. J. Heat
mer. Methods Fluids 78, 413–435. Mass Tranf. 47, 5101–5113.
Fedkiw, R., Aslam, T., Merriman, B., Osher, S., 1999. A non-oscillatory Eulerian ap- Saurel, R., Boivin, P., Metayer, O.L., 2016. A general formulation for cavitating, boiling
proach to interfaces in multidimensional flows (the ghost fluid method). J. Com- and evaporating flows. Comput. Fluids 128, 53–64.
put. Phys. 152, 457–492. Saurel, R., Petitpas, F., Abgrall, R., 2008. Modelling phase transition in metastable
Harten, A., Lax, P.D., van Leer, B., 1983. On upstream differencing and Godunov-type liquids: application to cavitating and flashing flows. J. Fluid Mech. 607, 313–350.
schemes for hyperbolic conservation laws. Soc. Ind. Appl. Math. 25 (1), 35–61. Sazhin, S., 2014. Droplets and Sprays. Springer.
Hertz, H., 1882. Ueber die verdunstung der flüssigkeiten, insbesondere quecksilber. Schrage, R.W., 1953. A Theoretical Study of Interphase Mass Transfer. Columbia Uni-
Ann. Phys. Chem. 253, 177–193. versity Press, New York.
Hindenlang, F., Gassner, G.J., Altmann, C., Beck, A., Staudenmaier, M., Munz, C.D., Sher, E., Bar-Kohany, T., Rashkovan, A., 2008. Flah-boiling atomization. Prog. Energy
2012. Explicit discontinuous Galerkin methods for unsteady problems. Comput. Combust. Sci. 34, 417–439.
Fluids 61, 86–93. Shusser, M., Ytrehus, T., Weihs, D., 20 0 0. Kinetic theory analysis of explosive boiling
Houim, R., Kuo, K., 2013. A ghost fluid method for compressible reacting flows with of liquid droplet. Fluid Dyn. Res. 27, 353–365.
phase change. J. Comput. Phys. 85, 865–900. Sonntag, M., Munz, C.D., 2014. Shock capturing for discontinuous Galerkin methods
Hu, X., Adams, N., Iaccarino, G., 2009. On the HLLC Riemann solver for interface using finite volume subcells. In: Fuhrmann, J., Ohlberger, M., Rohde, C. (Eds.),
interaction in compressible multi-fluid flow. J. Comput. Phys. 228, 6572–6589. Finite Volumes for Complex Applications VII-Elliptic, Parabolic and Hyperbolic
Knudsen, M., 1950. Kinetic Theory of Gases. London Methuen. Problems. Springer International Publishing, Cham, pp. 945–953.
Kopriva, D.A., 2009. Implementing Spectral Methods for Partial Differential Equa- Sussman, M., Smereka, P., Osher, S., 1994. A level set approach for computing solu-
tions. Springer Science + Business Media B.V.. tions to incompressible two-phase flow. J. Comput. Phys. 114, 146–159.
Lamanna, G., Kamoun, H., Weigand, B., Manfletti, C., Rees, A., Sender, J., Os- Tanaka, N., Furkawa, K., Suemori, S., Matsuo, T., Nishida, M., Yasutake, A., 2011. The
chwald, M., Steelant, J., 2015. Flashing behavior of rocket engine propellants. greening of spacecraft reaction control systems. Mitsubishi Heavy Ind. Tech. Rev.
Atom. Sprays 25 (10), 837–856. 48 (4), 44–50.
Lamanna, G., Kamoun, H., Weigand, B., Steelant, J., 2014. Towards a unified treat- Theofanous, T., Patel, P., 1976. Universal relations for bubble growth. Int. J. Heat
ment of fully flashing sprays. Int. J. Multiph. Flow 58, 168–184. Mass Tranf. 19 (4), 425–429.
Lauer, E., Hu, X., Hickel, S., Adams, N., 2012. Numerical modelling and investigation Toro, E.F., 2009. Riemann Solvers and Numerical Methods for Fluid Dynamics.
of symmetric and asymmetric cavitation bubble dynamics. Comput. Fluids 69, Springer-Verlag, Berlin Heidelberg.
1–19. Zein, A., Hantke, M., Warnecke, G., 2010. Modeling phase transition for compressible
Lee, H.S., Herman Merte, J., 1996. Spherical vapor bubble growth in uniformly su- two-phase flows applied to metastable liquids. J. Comput. Phys. 229, 2964–2998.
perheated liquids. Int. J. Heat Mass Transf. 39, 2427–2447.

You might also like