Edel Bauer 2016

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 36

Accepted Manuscript

Title: Numerical and experimental investigation of the spray quenching


process with an euler-eulerian multi-fluid model

Author: W. Edelbauer, D. Zhang, R. Kopun, B. Stauder

PII: S1359-4311(16)30274-5
DOI: http://dx.doi.org/doi: 10.1016/j.applthermaleng.2016.02.131
Reference: ATE 7860

To appear in: Applied Thermal Engineering

Received date: 29-10-2015


Accepted date: 28-2-2016

Please cite this article as: W. Edelbauer, D. Zhang, R. Kopun, B. Stauder, Numerical and
experimental investigation of the spray quenching process with an euler-eulerian multi-fluid
model, Applied Thermal Engineering (2016), http://dx.doi.org/doi:
10.1016/j.applthermaleng.2016.02.131.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service
to our customers we are providing this early version of the manuscript. The manuscript will
undergo copyediting, typesetting, and review of the resulting proof before it is published in its
final form. Please note that during the production process errors may be discovered which could
affect the content, and all legal disclaimers that apply to the journal pertain.
Numerical and Experimental Investigation of the Spray Quenching Process with an

Euler-Eulerian Multi-Fluid Model

W. Edelbauer1, D. Zhang2, R. Kopun3, B. Stauder4


1
AVL List GmbH, Hans List Platz 1, AT-8020 Graz, phone +43 316 787 7258, fax +43 316 787 777,

e-mail: wilfried.edelbauer@avl.com *Corresponding author


2
AVL List GmbH, Hans List Platz 1, AT-8020 Graz, phone +43 316 787 4225, fax +43 316 787 777,

e-mail: dongsheng.zhang@avl.com
3
AVL-AST d.o.o., Trg Leona Stuklja 5, SI-2000 Maribor, phone +386 31 707 796,

e-mail: rok.kopun@avl.com
4
Nemak Linz AG, Zeppelinstrae 24, AT-4030 Linz, phone +43 664 / 516 2052,

e-mail: bernhard.Stauder@nemak.com

Highlights:

New enhanced Euler-Eulerian spray model for spray quenching applications

New powerful automatic time step sub-cycling procedure for the evaporation model

Simulation time speed-up due to new workflow for periodically frozen fluid flow

Good agreement between measured and simulated cooling curves of the test specimen

Abstract

An Eulerian spray model for simulations of spray quenching and cooling in water mist chambers has been

developed. Spray quenching allows targeted cooling of heated mechanical components to obtain increased material

strength properties by modifying the micro-structure evolution. For the numerical multi-domain simulation it was

necessary to implement a heat transfer model for the fluid-solid interface taking into account the increased heat

transfer due to phase change of impinging droplets. The droplet size distribution of the spray is represented by

discrete size classes for which separate transport equations are solved. The main challenge of the simulation was the

long process time of several minutes. To reduce simulation time, a new time step sub-cycling method for the

evaporation model and a method for periodically frozen fluid flow have been developed. The aim of this paper is to

present an enhanced Euler-Eulerian spray model capable for the simulation of the spray quenching process. For

1
Page 1 of 35
model validation, measurements of the temperature history at certain locations within a quenched test geometry have

been performed and compared with the numerically predicted results from the CFD simulations. The paper discusses

the numerical method, the experimental set-up, and demonstrates the feasibility of the workflow for industrial

applications.

Keywords: Eulerian spray, quenching, spray quenching, evaporation, time-step sub-cycling.

Nomenclature

Variables:

A Area (m2)

B M, B T Spaldings mass and heat transfer number (-)

cD Drag coefficient (-)

cP Specific heat capacity (J/kgK)

cT Turbulent dispersion force coefficient (-)

CFL CourantFriedrichsLewy number (-)

CTC Correction factor or the heat transfer coefficient (-)

d Diameter (m)

D Diffusion coefficient (m2/s)

fmax Maximum change rate of mass or enthalpy (-)

g Gravitational acceleration (m/s)

h Specific enthalpy (J/kg)

hTC Heat transfer coefficient (W/m2K)

H Interfacial heat exchange term (W/m)

i Loop index (-)

k Turbulence kinetic energy (m2/s2)

L Latent heat (J/kg)

m Mass (kg)

M Molecular weight (kg/kmol)

M Momentum inter-phase exchange term (N/m)

nd Number of droplets (-)

2
Page 2 of 35
nph Number of phases (-)

nsp Number of gas species (-)

nsub Number of sub-cycling time steps (-)

N Number density of droplets (1/m)

Nu Nusselt number (-)

p Pressure (Pa)

Pr Prandtl number (-)

q Heat flux (W/m)

qT Sub-cycling coefficient (-)

Q Heat transfer (J)

Re Reynolds number (-)

Sc Schmidt number (-)

Sh Sherwood number (-)

t Time (s)

T Temperature (K)

U Velocity component (m/s)

v2 Reynolds stress component normal to wall (m2/s2)

V Volume (m3)

Vs Droplet mass flux (kg/m2s)

x Space (m), molar fraction (-)

Y Species mass fraction (-)

Volume fraction (-)

Mass inter-phase exchange rate (kg/m3s)

Turbulence dissipation rate (m2/s3)

Energy source (W/kg)

Thermal conductivity (W/mK)

Dynamic viscosity (Ns/m)

Number density exchange rate (1/ms)

3
Page 3 of 35
Kinematic viscosity (m/s)

Density (kg/m3)

Shear stress (N/m2)

Ratio of wall normal velocity scale and turbulence kinetic energy (-)

Subscripts:

amb Ambient (here: condition in gaseous phase)

c Continuous phase (liquid phase)

C Collisions

d Dispersed phase (droplet phase)

drop Single droplet

E Evaporation

eff Effective

g Gaseous phase index

i, j Index notation

int Interface

k, l Phase indices

M Mass

P Primary break-up

Q Heat

s Species index

S Secondary break-up

sat Saturation

vap Vapor

w Wall

Superscripts:

t Turbulence

Abbreviations:

CFD Computational fluid dynamics

4
Page 4 of 35
DDM Discrete droplet method

HTC Heat transfer coefficient

FF, NFF Frozen flow, Non frozen flow

ODE Ordinary differential equation

1. Introduction

The ongoing demand on reduced weight and increased mechanical and thermal load of IC engine components leads

to sophisticated manufacturing processes of the components. Quenching is a well-established process for casted

aluminum alloy components like cylinder heads and engine blocks to increase the material strength properties by

modifying the micro-structure evolution. The residual stresses and the mechanical properties strongly depend on the

local time-temperature history obtained during the quenching process. One can distinguish between immersion

quenching and spray quenching. For the latter process galleries of nozzles perform the atomization of the cooling

fluid. The heat transfer from the solid to the fluid is determined by the impinging spray droplets and their

evaporation. In praxis, the optimization of these processes is expensive and time consuming due to many trial and

error experiments. The application of advanced numerical methods still plays a minor role in the development of the

production process.

The aim of this paper is to present a simulation approach for industrial spray quenching processes by applying an

Euler-Eulerian multi-fluid approach. Beside of modeling evaporation/condensation and droplet-wall heat transfer,

the main difficulty is to treat the time-scales of the involved processes. Typical spray quenching processes take place

within the time frame between 60 and 1200 seconds. The time required for single spray droplet to be heated,

evaporated and impinged is several orders of magnitude faster. Thus, simulation of the complete spray quenching

process would require unfeasible high number of time steps. By certain techniques presented in this paper, like sub-

cycling and temporarily frozen fluid flow field, the total number of time steps could be reduced and a significant

speed-up of the simulation time was achieved.

One key feature of the spray quenching simulation is the heat transfer model between impinging droplets and the

solid surface. This process is considerably different to that of immersion quenching which was discussed in previous

works by Srinivasan et al. [1] and Kopun et al. [2]. Since the late 1960s, spray cooling was investigated extensively

by Eugene and Mizikar [3], Gaugler [4] and others. Pioneering works mainly have been conducted for individual

droplet cooling and/or very low liquid droplet mass flux to obtain a basic understanding. Table 4 gives an overview

5
Page 5 of 35
about several experimental spray cooling and quenching studies, and it summarizes the range of their characteristic

parameters. The following four major heat transfer mechanism occurs during spray cooling and have been reported

by several authors, i.e. Pais et al. [5], Yang et al. [6], Rini et al. [7], Pautsch and Shedd [8], Shedd and Pautsch [9]

and others:

1. Evaporation on the surface of the liquid film.

2. Forced convection arising from droplets impinging on the hot surface.

3. Enhancement of nucleation sites on the heated surface.

4. Presence of secondary nucleation sites on the surface of spray droplets.

Mudawar and Estes [10], Visaria and Mudawar [11], and Lin and Ponnappan [12] proposed empirical models for the

critical heat flux of spray cooling. The applicability of these models is limited to their validated conditions. All three

boiling regimes appear in the spray quenching process, and consequently, although spray quenching is widely used

in industry, in the open literatures only limited investigations on industrial scale spray quenching process could be

found, as there are Wendelstorf et al. [13], Al-Ahmad and Yao [14], and Mascarenhas and Mudawar [15]. The

common conclusion from these publications is that the measured heat transfer rate strongly depends on the

temperature difference between wall and fluid, the local droplet mass flux, the droplet velocity and their size. Yao

and his co-operators (Cox [16], Choi and Yao [17] and Deb and Yao [18]) investigated laboratory-produced, mono-

dispersed spray cooling and obtained correlations for the heat transfer for a wide range of droplet mass fluxes from

0.016 to 2.05 kg/m2s. Mudawar and his team (Klinzing et al. [19], Bernardin and Mudawar [20], Mascarenhas and

Mudawar [15]) published a series of papers on the spray cooling, but the application range is narrow and strictly

defined. Based on extensive experimental data, Yao and Cox [16] provided two correlations for the spray heat

transfer. The heat transfer rate is either a function of the spray Reynolds or the Weber number. Wendelstorf et al.

[13] performed a series of measurements using full cone nozzles, where the droplet mass flow rates range from 3 to

30 kg/m2s and the temperature difference between hot surface and spray droplets varies from 150 to 1100 K. The

authors present a reasonable boiling curve correlation as function of the full temperature difference and the droplet

mass flux. Due to its wide application range, this correlation is adopted in this paper.

The modeling of droplet evaporation and condensation is a further important part of the spray quenching model. It is

based on single droplet models. Especially for the Euler-Lagrangian Discrete Droplet Method (DDM), as suggested

by Dukowicz [21], several specific evaporation models can be found in the literature, such as Amsden et al. [22],

6
Page 6 of 35
Varnavas and Assanis [23], Abramzon and Sirignano [24], [25], and several others. The latter model, which is

developed for non-unity Lewis numbers, has been adapted for Euler-Eulerian multi-fluid approach and is applied in

the current study.

The paper is organized as follows. After the introduction in the first section, the second section discusses the

modeling approach, containing a summary of the entire Euler-Eulerian multi-fluid model, the turbulence,

evaporation and the heat transfer model for the droplet wall interaction. Then the sub-cycling methodology of the

evaporation model and the applied approach of the frozen fluid flow field are presented. The third section discusses

the performed experiment, the simulation set-up and the comparison between measured and numerically predicted

results. The presented models of this work have been implemented by the authors into the CFD code AVL-FIRE.

2. The Modeling Approach

2.1 Euler-Eulerian Multi-Fluid Model

The standard model for the simulation of gas flows moderately laden with droplets is the Euler-Lagrangian Discrete

Droplet Method (DDM). Beside of many benefits, one drawback of the DDM is that an adequate number of spray

parcels in every computational cell is required to provide a statistically reliable description of the droplet size

spectrum in the entire computational domain. This can lead to high computational effort due to the large number of

spray parcels required for covering the entire computational domain. Furthermore, MPI parallelization efficiency of

the DDM suffers from parcels crossing the domain interfaces. In order to avoid these difficulties, the Euler-Eulerian

multi-fluid approach has been applied for the spray simulations in the present study. The solved transport equations

are summarized in this section.

Based on the work of Drew and Passman [26], the complete set of conservation equations is solved for an arbitrary

number of interpenetrating phases. Each phase is characterized by its volume fraction . For the Euler-Eulerian

spray approach, one gaseous phase and an arbitrary number of phases for the droplet size classes are defined. The

mass, momentum and energy conservation equations for phase k with volume fraction k, are described by the

equations

(1)

(2)

7
Page 7 of 35
(3)
.

The terms kl, Mkl,i and Hkl represent the interfacial mass, momentum and heat transfer terms between phase k and

another phase l respectively. Uniform pressure, pk = p for all phases is assumed. The energy equation is solved for

the relative total enthalpy hk, which includes the thermal and the kinetic energies as described by

. Once the enthalpy equation is solved, the phase temperature Tk is calculated

from hk with the integral specific heat capacity cP,k. The term qk is the heat flux vector consisting of a conductive and

a turbulent part, and k is the specific enthalpy source.

The gaseous phase consists of an arbitrary defined number of gas species. For the spray quenching simulations the

gaseous phase is defined as composition of oxygen O2, nitrogen N2 and water vapor H2O. For each species s of the

gaseous phase k=g the transport equation for the vapor mass fraction Ys , described by

(4)

is solved. DY,eff is the effective diffusion coefficient. SY,s is the volumetric mass transfer rate of species s from phase

change, chemical reactions, etc.

2.2 Turbulence Model

In general the nature of the flow in the spray quenching process is turbulent. The applied turbulence model is based

on the eddy-viscosity concept. Different to the isotropic standard k- model, as proposed by Launder and Spalding

[27], the applied four equation k---f model, proposed by Hanjalic et al. [28] and enhanced by Basara et al. [29],

[30] is not isotropic. It damps the turbulence close to the wall depending on the wall normal velocity scale . This

damping is especially important for the spray impingement area.

The k---f model is particularly suitable for predicting near wall turbulence. The authors propose a version of the

eddy-viscosity model based on Durbin's elliptic relaxation concept [31]. The aim is to improve numerical stability of

the original - model by solving a transport equation for the velocity scale ratio instead of the wall

8
Page 8 of 35
normal velocity scale . This model has been adopted for multi-fluid applications, and it is fully available in the

applied CFD code. For every phase the k---f model solves transport equations for the turbulence kinetic energy k,

the turbulence dissipation rate and the velocity scale . Additionally, elliptic relaxation equations for f, which is a

parameter closely related to the pressure-strain redistribution term, are solved for each phase. According to the

closure assumptions for multiphase turbulence of Lee et al. [32], it is assumed that the turbulence level of the

dispersed phase is equal to the continuous phase, and the interaction between these phases is neglected. A summary

of the k-f model equations can be found in the publication of Greif et al. [33].

2.3 Eulerian Spray Model

The Eulerian spray model describes the interfacial exchange term of the Euler-Eulerian multi-fluid model with

respect to the droplet physics. The frame of the model is briefly described in this section. Further details can be

found in the publications of von Berg et al. [34], [35], and Edelbauer et al. [36]. Model enhancements and

validation, preferably on high pressure Diesel injections, are described by Vujanovi et al. [37] and Petranovi et al.

[38], [39]. The Eulerian spray model solves transport equations for the gaseous phase and an arbitrary number of

liquid droplet phases, where the transport equations are discussed in section 2.1. Each liquid phase represents a

droplet size-class. The phase specification of the Eulerian spray model is shown in Table 5. The first phase is

gaseous and the others are liquid representing the droplet size distribution.

In the Eulerian spray model the interfacial mass exchange term kl in equation (1), gets its contribution from droplet

evaporation E, primary break-up P, secondary break-up S, and collisions C as described by the equation

(5)

The mass exchange between two phases for a process acts in both directions, so that one can write Process,kl =

-Process,lk. This means that a mass gain for phase k is a mass loss for phase l and vice versa, so that mass

conservation over all phases is ensured. For the intended spray quenching simulations, the droplet size distribution at

the inlet is pre-defined based on the nozzle specification sheet. Consequently, the spray inlet boundary condition is

not located at the nozzle orifice, it is located a certain distance downstream, where break-up and collision processes

are in most instances finalized, and the droplet size distribution can be assumed to be constant. Thus, collision,

primary and secondary break-up models are not enabled in the performed simulations. Evaporation takes place

between the liquid phases from 2 to nph and the gaseous phase 1. The interfacial mass exchange rate due to

evaporation between the droplet phase k and the gaseous phase 1 is determined by the equation

9
Page 9 of 35
(6)

where Nk denotes the number of droplets per unit volume, and determines the evaporation mass exchange rate

of a single droplet. Assuming spherical droplets of diameter dk, the liquid volume fraction can be expressed by

. Consequently the droplet diameter of phase k is calculated by equation

(7)

where for the volume fraction k and the droplet number density Nk, separate transport equations are solved. This

means that within each size class the droplet diameter varies in space and time. Additionally the size class borders

are continuously adapted. The transport equation for the droplet number density Nk yields

(8)

The source term on the right-hand side, kl, gets its contributions from all droplet producing or annihilating

interfacial mass transfer processes. Evaporation and thermal expansion Th due to temperature change lead to

changes of the droplet diameter, but do not influence the number of droplets. Thus, E,kl and Th,kl are zero. The

child droplets from primary and secondary break-up or the coalescence from collision would create the sources P,kl,

S,kl, C,kl. Since these processes are not taken into account, the right-hand side of equation (8) is zero.

The momentum exchange between the gaseous phase 1 and the droplet phases k from 2 to nph, Mk1, is determined by

drag forces, MD,k1, and turbulent dispersion forces, MT,k1, as described by the equation

(9)

The drag coefficient cD is based on a correlation from ORourke [41] and depends on the droplet Reynolds number

Re and the liquid volume fraction, as described by

(10)

The turbulent dispersion force is modeled following the approach of Lopez de Bertodano [42], with a constant or

modeled turbulent dispersion force coefficient cT.

2.4 Evaporation and Condensation Model

This section briefly describes the applied evaporation/condensation model for the Eulerian spray simulation, and

discusses the effects of the simulation time steps. The applied model of Abramzon and Sirignano [24], [25] is a

10
Page 10 of 35
single droplet model for non-unitary Lewis number. It has been adapted for the Euler-Eulerian framework to provide

the mass transfer term in equation (6). In this section index k=d denotes the droplet phase and index g the

gaseous phase. The single droplet mass transfer rate is described by the equation

(11)

where BM is the Spaldings mass transfer number, defined by the equation

(12)

The overbar in equation (11) indicates mixture properties at a certain reference condition in the gas layer around the

droplet. is the mixture gas density, is the diffusion coefficient, and is the modified Sherwood number, as

described by the equation

(13)

Yamb is the vapor mass fraction of the ambient gas, and Yd is the vapor mass fraction on the droplet surface, which

depends on the ambient pressure p, the saturation pressure psat and the fluid properties, as described by the equation

(14)

M is the molecular weight and x is the mole fraction of the gas species. The heat transferred into the droplet interior

is determined by the equation

(15)

where is the specific heat capacity of the vapor at reference conditions, and L is the latent heat of the fluid. BT

is the Spaldings heat transfer number. From equation (15) one can see that the heat transfer is directly proportional

to the mass transfer. By multiplying with the droplet number density Nk, one obtains the interfacial heat

transfer term Hkl of the enthalpy transport equation (3). Furthermore, infinite conductivity inside the droplet is

assumed leading to homogenous temperature distribution within the droplet. This model has been implemented by

the author into the CFD code FIRE, and it was applied and validated on high pressure Diesel injections, as shown

in previous publications (Edelbauer [43]; Petranovi et al. [38], Vujanovi et al. [37], and Edelbauer et al.[36]). The

application of this evaporation model for spray quenching is considerably different. In typical Diesel sprays, the

11
Page 11 of 35
droplet size varies between approximately zero and fifty micrometers during evaporation, and the injection period is

in the order of few milliseconds. Consequently, typical simulation time steps are in the order of few microseconds,

or smaller depending on the resolution of the computational mesh. These time steps are usually small enough to

cover the physical processes for evaporation/condensation as well as for droplet break-up. For typical spray

quenching processes the situation is different. There, the process usually takes place over several minutes. Injection

pressure is usually much lower and the droplet size varies between zero and a few hundred micrometer. While the

size of the droplets is in comparable order, the total process time is much longer. To achieve reasonable

computational times, simulation time steps between 10-3 and 10-2 seconds are desired. These large time steps may

lead to problems which are illustrated on two examples in Appendix A to get an understanding about the relevant

time scales of the evaporation/condensation process at typical spray quenching conditions.

The conclusion from these tests is that the evaporation effects under typical quenching conditions, are much faster

than the intended time step sizes for simulation of 10-3 to 10-2 seconds. The fast changes in temperature and diameter

cannot be captured by those time steps. Especially the sharp initial gradient of the liquid temperature can lead to

numerical problems and inaccurate solution, if the time step is not small enough within this period. To overcome

this problem, a time step sub-cycling procedure has been developed and implemented into the numerical solution

procedure, which is presented in the following section.

2.5 Time Step Sub-Cycling for Evaporation and Condensation

The idea of the time step sub-cycling procedure is not new. Amsden et al. [22] describes a methodology for the

DDM approach. The idea is to calculate the evaporation and condensation of every spray parcel with smaller time

steps, and to integrate the obtained mass and heat transfer rates between the liquid and the gaseous phase. The

choice of the sub time step is based on the assumption that during one CFD solver time step t the heat/mass

transfer to/from a computational parcel should not exceed some fraction of the energy/mass available for transfer.

This idea has been picked up for the Eulerian spray model to obtain integrated transfer rates which are then divided

by the CFD solver time step to provide the final sources for the volume fraction and enthalpy transport equations.

Amsden et al. (1989) derived the maximum time step for the Lagrangian spray parcels from considerations about the

maximum diffusive mass transfer by assuming unitary Schmidt number, as described by the equation

(16)

12
Page 12 of 35
The equation here already takes into account the gas volume fraction g which reduces the effective gas volume in

the computational cell. Factor fmax determines the maximum change of the gas mass in the cell, the chosen default

value of fmax is 0.5. Vcell is the cell volume, g is the gas density, Sh is the Sherwood number, g is the molecular

viscosity of the gas, and nd is the number of parcels, here droplets, in the computational cell. A similar criterion from

Amsden et al. [22] based on the maximum change of the gas energy in the cell leads to the maximum sub-cycling

time step described by the equation

(17)

g is the thermal conductivity and cp,g is the specific heat capacity of the gas. Nu is the Nusselt number. For adapting

these correlations in the Euler-Eulerian multi-fluid approach, the number of droplets nd per computational cell has to

be determined by the equation

(18)

Substituted it into the equations (16) and (17), the sub-cycling time steps yield

(19)

(20)

Nusselt and Sherwood number are calculated according to the well-known correlation of Ranz and Marshal [44]

with and by assumed analogy of heat and mass transfer. The

number of sub-cycling time steps nsub for the computational cell Vcell and the droplet phase d is then simply

determined by the equation

(21)

The test cases in Appendix A have shown that the initial change of the liquid temperature may become very large.

Thus, it is reasonable to use small sub time steps at the beginning and to increase the sub time steps later. Therefore,

the time steps are calculated from a geometric series as described by the equation

(22)

13
Page 13 of 35
For this approach the first sub-cycling time step tsub,1 and coefficient q have to be determined. It has been decided

that the last sub time step tsub,n is nsub times larger (but maximum ten times) than the first sub time step tsub,1, as

described by

(23)

Thus coefficient qT is calculated from the equation

(24)

and the first sub time step tsub,1 can be determined by the equation

(25)

All further sub time steps are then calculated by the equation

(26)

During the sub-cycling procedure the droplet diameter and temperature are continuously updated. The new droplet

diameter and temperature for the next sub time step i are obtained from simple mass and heat balance and calculated

by the equations

(27)

where md,i is the single droplet mass. The final exchange terms from evaporation for the volume fraction and

enthalpy transport equations between gaseous phase g and droplet phases d are then described by the equations

(28)

Optionally, it is possible to make estimations about the temperature and the vapor mass fraction in gaseous phase for

each sub-time step, as shown next. It can be shown from equation (15) that the rate of heat transferred into the gas

equals . Consequently, the gas temperature at the next sub-time step is described by the

equation

14
Page 14 of 35
(29)

The vapor mass fraction of the gas is directly influenced by the evaporated mass of all droplet phases in the

computational cell, and it yields

(30)

The updates of Tg,i+1 and Yamb,i+1 are only reasonable for CFL numbers less than unity. For CFL>1 the change of

temperature and vapor content in the computational cell is mainly influenced by the convective mass transport, and

consequently by the condition of the upwind cell. Thus, the gas property update should not be performed there.

The evaporation model is implemented as depicted in Figure 5, where the flow chart describes the steps performed

to determine the interfacial exchange terms of the volume and enthalpy transport equation due to evaporation. For

every cell a loop over all gas droplet interfaces, index iint, is performed. nsub according to equation (21) is calculated,

and the sub-cycling time step loop, index isub, is performed. After calculation of the time step size tsub, the mass and

heat transfer rates for the single droplet, and , are calculated. Then the intermediate values for droplet

diameter and temperature are calculated by equation (27), and optionally also the intermediate temperature and

vapor mass fraction of the gaseous phase in the computational cell is executed by equations (29) and (30). Here it

must be mentioned that, this update is a rough estimation, since convective effects are not taken into account.

Furthermore, the estimated values of Tg,i+1 and Yg,i+1 depend on the interface index iint and fits best at the last

interface and the last sub time step, since then the accumulation of transferred heat and mass in the current

computational cell is finalized.

2.6 Wall Heat Transfer Model for Spray Quenching

During spray quenching, the prevailing heat transfer proceeds between the hot solid surface and the sub-cooled

droplets. It is reported by many researchers that the heat transfer rate between the hot surface and the impinging

spray droplets strongly depends on the temperature difference T=Tw-Tdrop and on the local droplets mass flux Vs.

Based on extensive measurements with water as coolant, Wendelstorf et al. [13] proposes the following correlation

for the heat transfer coefficient:

(31)

This equation results from experiments with full cone nozzles (VKE 6/60, 6/90 and 8/60), where the nozzle diameter

15
Page 15 of 35
was 350 m, T was varied between 150 and 1100 K, and Vs was in the range between 3 and 30 kg/m2s. Figure 6

left shows the curve for three different mass fluxes of Vs1=3, Vs2=15 and Vs3=30 kg/m2s which cover the

experimental range and consequently the validity of this correlation. The corresponding heat flux is shown in

Figure 6 right. One can observe the critical heat flux occurs at T between 200 and 250 K, and the Leidenfrost

temperature limit is located at T greater than 500 K. Other researchers report that the heat transfer coefficient

additionally depends on the size of the impinging droplets. In the measurements performed by Wendelstorf et al.

[13] this effect was, owing to the high droplet mass fluxes, negligible.

It has to be noted that equation (31) is an integral heat transfer correlation. The measurements were performed on

laboratory samples (quenched surface is a circular disc with a diameter of 70 mm), and the temperature difference

T is defined as the difference between the hot wall and the coolant temperature in the nozzle. In the presented CFD

simulations, this correlation is applied locally for every discretized surface element. Droplet mass flux Vs and

temperature difference T are determined locally from the cells adjacent to the surface element. To overcome these

uncertainties between integral and local heat transfer correlation and the effect of different nozzle types, a correction

coefficient CTC, which is now a model parameter, has been introduced. Similar corrections are reported by Pola,

Gelfi and Vecchia [40].

On the hot wall, evaporation mass exchange E,lg occurs between the droplets or the liquid film and the surrounding

gas, represented by phases l and g. The energy for this additional phase change results directly from the solid wall

and is described by the equation

(32)

where Tsat is the saturation temperature, Tw is the heated wall temperature and Aw is the surface area on the wall

boundary.

2.7 Alternating Frozen Fluid Flow Field

Due to multi-fluid flow and long process time the numerical simulation of spray quenching/cooling is

computationally expensive. In a primary survey, it is found that the effort for solving the transport equations of the

fluid, i.e. volume fraction, momentum, turbulence as well as pressure correction equation, dominates compared to

the effort for solving solid and fluid energy equations. Observing the flow field shows that after the spray droplets

arrive at the hot surface, the flow field varies very slowly with time proceeding. This behavior is also observed in

16
Page 16 of 35
the experiment. Consequently, it is practically meaningful and reasonable to keep velocity, pressure, volume fraction

and turbulence fields periodically frozen. This means that after the droplets impinged on the hot surface, the

transport equations for momentum, volume fraction, turbulence and the pressure correction equation are solved

periodically while enthalpy equations for fluid and solid domains are solved continuously. After a short period of a

few time steps where all transport equations are solved, there follows a longer period where only enthalpy equations

are solved. This process is periodically repeated.

3. Computational Results

3.1 Experimental Set-up

For the current study, experimental investigations of the spray quenching process have been performed. A simple

cuboid alloy made of AlSi7MgCu0 was quenched as displayed in Figure 7. Inside the work piece eight

thermocouples of type K (Ni-CrNi) with a temperature range of T=1200 K were located. Figure 7b shows the

dimensions of the cuboid sample and the positions of the eight thermocouples, illustrated by red dots from T1 to T8.

The holes for the thermocouples are positioned in the middle of the solid piece in a depth of 15 mm and a height of

30 mm. An air forced oven was used to heat up the work piece to homogeneous temperature of 780 K. Then the

sample was carried to the test station and placed on a fire refractory plate with the thickness of 40 mm. The

temperature profiles during quenching were recorded with the multi-meter DEWE-PORT 2000 at the frequency of

2 Hz.

The experimental set-up of the spray quenching process of the rectangular block specimen is summarized in Table 6.

A twin-fluid atomization nozzle operated with water and air was applied for the current study. The reproducibility of

the spatial position was secured by fixing screws, as shown in Figure 7a.

3.2 Simulation Set-up

Figure 8 demonstrates the physical domain and the boundary conditions applied in the CFD simulations, and in

Table 7 the simulation set-up is summarized. The present computational domain consists of 800 000 cells in total,

where the solid domain has approximately 50 000 cells. The mean cell size of the solid domain is approximately

2 mm. The mesh dependency test, presented in Appendix B, has shown that this is an adequate cell size. The

computational domain has rather large dimensions in order to cover the whole working piece and to ensure enough

distance to the numerical boundary conditions. Atmospheric pressure is set at the static pressure boundaries on the

top, bottom and sides as depicted by the blue surfaces in Figure 8a. At the back side of the domain, a no-slip wall

17
Page 17 of 35
boundary condition with constant environment temperature of 20 C is applied (yellow surface). The fire refracting

plate, illustrated by the black surface, was not explicitly modelled as solid body, but an adiabatic wall boundary

condition with zero heat flux is applied. Since the experimental configuration is symmetric, only half of the

geometry is modeled in order to reduce the computational effort. Therefore symmetry boundary conditions,

illustrated by the green surface, are applied. For the spray, an inlet boundary condition with prescribed velocity and

volume fraction is applied. The normal distance between injector orifice and solid is 75 mm. Instead of resolving the

fine nozzle orifices of the six-hole atomizer, a ball-shaped inlet boundary condition is located 11.5 mm from the

nozzle orifice and 63.5 mm from the solid. The domain interface between fluid and solid domains, illustrated by the

red surface, is determined by a no-slip wall boundary condition. There, the wall temperature is calculated in order to

obtain balanced heat flux on both sides of the interface. The applied correction coefficient CTC is a model parameter

which reflects the differences between the empirical heat transfer correlation in equation (31) and the present

numerical model, as discussed in section 2.6. The maximum time step in the simulation is 0.01 seconds.

In the current study, two different quenching simulations have been investigated and compared with the

measurement data. The first simulation uses the time step sub-cycling methodology for the evaporation model as

described in section 2.5 but with the non-frozen fluid flow, whereas the second simulation uses both, sub-cycling

and temporary frozen fluid flow, as described in section 2.7. The initial droplet size diameters and the inlet volume

fractions are specified according to the data of the injection specification in order to approximate the droplet size

distribution from the manufacturer. The three droplet phases together with the gaseous phase lead to four phases in

total. The gaseous phase consists of three species, oxygen O2, nitrogen N2 and water vapor H2O for which separate

transport equations are solved. The three thermocouples, T1, T5, and T7 in Figure 8b, are used for comparison

between measured and numerically predicted temperatures results. Both simulations, non-frozen flow (NFF) and

frozen flow (FF), share the same monitoring points.

3.3 Simulation Results

The comparison of measured temperatures and numerically predicted results is presented next. It focuses on the

temperature and liquid phase distributions. The latter is represented by the so called total volume fraction which is

the sum of all liquid volume fraction values in the computational cell.

3.3.1. Non frozen fluid flow


Simulations were conducted on a Linux cluster with 20 CPU cores and the total simulation time was approximately

18
Page 18 of 35
94 hours. Figure 9 displays the comparison between measured and numerically predicted temperature histories at the

three monitoring locations, T1, T5, and T7. In general a good agreement was achieved, although the numerical

simulations slightly overestimate the cooling rates. It can be seen that at the monitoring points, which are closer to

the prolongation of the spray axis, positions T1 and T5, faster cooling than at the monitoring point T7 occurs. This is

due to the fact that at T1 and T5 more droplets are impinging on the hot surface and consequently, the amount of heat

consumed for evaporation is higher. It can be seen in Figure 6 right that with the initial solid temperature of 500 C

the cooling process starts below the Leidenfrost point. Consequently, the cooling curve passes the transition and

nucleate boiling regime. At the end time of 200 s, the maximum deviation between simulated and measurement

results is less than 40 K in all monitoring points.

The total liquid volume fraction distribution at different time instants is presented in Figure 10, where red color

represents water droplets and blue is gas. It can be seen that at 0.01 seconds the spray has not reached the solid

surface. This takes place at around 0.03 seconds. During this free flow period only droplet evaporation takes place.

When the droplets get in contact with the hot surface, the evaporation rate increases dramatically due to the heat

transfer from the solid to the liquid. One can observe that the spray shape is relatively similar over the whole

simulation time. But at later time, when the work piece is already cooled down, one can observe higher droplet

concentration. Between 150 and 200 seconds, the quenched surface at the spray axis prolongation has almost

reached environmental temperature, and therefore the evaporation rate is very small.

Figure 11 displays the surface temperature distribution and the total liquid volume fraction at different time instants

during the entire quenching cooling process. At 0.01 seconds the solid has its initial temperature of 773.15 K, while

after 10 seconds the cooling is clearly visible. At this time monitoring point T 1 shows a temperature of

approximately 700 K while T7 has around 750 K. It can be seen that between 50 and 150 seconds the temperature

field is very heterogeneous and the temperature difference within the work piece reaches T=150 K. The numerical

results of all monitoring point locations show that the surface temperature has dropped substantially at the areas

closer to the prolongation of the spray axis, whereas slower cooling is observed at the outer monitoring points where

less droplets impact the hot surface, i.e. T7 and T8. As expected, the local droplet mass flux has dominant influence

on the entire quenching cooling process.

The heat transfer coefficients and heat flux at three different time instants for one droplet phase are illustrated in

Figure 12. There, solid surface temperature, heat transfer coefficient and the heat flux are shown for the second

19
Page 19 of 35
droplet phase with 37 m class diameter (Table 7). It is important to mention that the total heat flux is the sum over

all droplet phases and the gas phase, and that is consequently considerably higher. The change of the heat transfer

owing to the transition between different heat transfer regimes, as discussed in Figure 6, is visible. For better

contrast in the figures, the color-bar has been adapted. While at the beginning at 0.04 s, where the spray has just

impinged the hot surface, the wall heat flux shows a minimum value, a few seconds later at 5 s the heat flux is

significantly higher, and at 100 s heat flux and heat transfer coefficient are reduced again. It can be concluded that

for this droplet phase the critical heat flux in the impinged area is passed at around 5 s. It can be also observed that

the change in the heat transfer coefficient at the three time instants is not as significant as in the evaluation in

Figure 6. This difference results from the fact that an integral heat transfer correlation is applied locally for every

discretized surface element in the CFD simulation, as discussed in section 2.6.

3.3.2. Frozen fluid flow


This simulation has been performed with the same set-up under the same conditions. The only difference was that

the fluid flow was periodically frozen for every 3 seconds as discussed in section 2.7. During the initial period of 0.5

seconds all equations were solved followed by a frozen period where only enthalpy equation was solved up to 3

seconds. Then all equations were solved for further 0.5 s followed by a further frozen period of 2.5 s. This procedure

was repeated up to the final time of 200 seconds. This measure leads to a total simulation time of 37 hours compared

to 94 hours for the previous case. A speed-up of 2.5 was achieved. The temperature histories of the simulations with

non-frozen and alternating frozen fluid flow of the three monitoring points, T1, T5, and T7, are shown in Figure 13.

The difference between the two simulations is very small demonstrating the feasibility of this methodology. The

frozen fluid flow simulation shows a slight delay in the cooling curve.

4. Summary and Conclusions

This work presents an Euler-Eulerian spray model for spray quenching applications. The model was originally

developed for high-injection Diesel sprays, and has been enhanced for spray quenching simulations. Therefore, a

spray droplet-solid heat transfer model, taking into account the phase change of droplets impinging on a hot surface,

has been implemented into a CFD code. Two domains, fluid and solid, are coupled through a heat exchanging

interface. While in the fluid domain the complete set of transport equations is solved for one gaseous and an

arbitrary number of liquid droplet phases, in the solid domain only the enthalpy transport equation is solved. The

droplet size distribution is represented by the diameter classes of the liquid droplet phases. Numerical investigations

20
Page 20 of 35
of the evaporation and condensation on single droplets have been performed to demonstrate that droplet temperature

and diameter changes take place on short time scales. These time scales are considerably smaller than the desired

CFD time steps required for covering the quenching process. In order to reduce the number of time steps, a sub-

cycling procedure has been developed for the evaporation and condensation model. Additionally, a methodology for

periodically frozen fluid flow field for further reduction of the computational effort is presented. Experiments of the

spray quenching process have been performed on a prismatic specimen. The solid temperature histories have been

measured on certain locations within the work piece. There is a good overall agreement between the measured and

numerically predicted temperature curves. Theses curves are the base for stress simulations which will be performed

in the next step, in the same way as it was performed by Kopun et al. [45] and Li et al. [46]. Nevertheless, the

achieved accuracy of the cooling curves allows qualitative prediction of the stresses, for quantitative prediction

further increase of the simulation accuracy is necessary.

The current work has shown that the Euler-Eulerian spray approach is suitable for industrial spray quenching

applications. Through measures like time step sub-cycling and periodically frozen fluid flow the computational

effort could be considerably reduced. In the current work a speed-up of 2.5 between simulations without and with

periodically frozen fluid flow was achieved. As next step, further simulations with increased number of droplet size

classes, in order to have a better representation of the droplet size distribution, will be performed. The application of

this method on more complex work pieces, e.g. cylinder heads, is part of further investigation.

Acknowledgements

Parts of this research work were performed within the FFG funded project QUENCH-IT, project number 828697.

Other parts are funded by the European Union, European Social Fund and SPIRIT Slovenia, Slovenian Public

Agency for Entrepreneurship, Innovation, Development, Investment and Tourism.

References

[1] Srinivasan, V., Moon, K. M., Greif, D., Wang. D.M., Kim M., 2010. Numerical simulation of immersion

quench cooling process using an Eulerian multi-fluid approach. Appl. Therm. Eng. 30 (5), 499-509.

http://dx.doi.org/10.1016/j.applthermaleng.2009.10.012.

[2] Kopun, R., Skerget, L., Hribersek, M., Zhang, D., Stauder, B., Greif, D., 2014, Numerical simulation of

immersion quenching process for cast aluminium part at different pool temperatures. Appl. Therm. Eng. 65

(1-2), 74-84, http://dx.doi.org/10.1016/j.applthermaleng.2013.12.058.

21
Page 21 of 35
[3] Eugene, A., Mizikar, A., 1970. Spray cooling investigation for continuous casting of billets and blooms.

Iron Steel Eng. 47 (6), 53-60.

[4] Gaugler, R. E., 1966. An experimental study of spray cooling of high temperature surfaces. PhD thesis,

Department of Mechanical Engineering, Carnegie Institute of Technology, Pittsburg, PV, USA.

[5] Pais M. R., Chow L. C. and Mahefkey E. T., 1992. Surface roughness and its effects on the heat transfer

mechanism in spray cooling, J. Heat Transfer 114, 211-219.

[6] Yang J., Chow L. C. and Pais M. R., 1996. Nucleate Boiling Heat Transfer in Spray Cooling. J. Heat

Transfer 118, 668-671.

[7] Rini D. P., Chen R. H. and Chow L. C., 2002. Bubble Behavior and Nucleate Boiling Heat Transfer in

Saturated FC-72 Spray Cooling. J. Heat Transfer 124, 63-72.

[8] Pautsch A. G. and Shedd T. A., 2005. Spray impingement cooling with single- and multiple-nozzle arrays.

Part I: Heat transfer data using FC-72. Int. J. Heat Mass Transfer 48, 31673175.

[9] Shedd T. A. and Pautsch A. G., 2005. Spray impingement cooling with single- and multiple-nozzle arrays.

Part II: Visualization and empirical models. Int. J. Heat Mass Transfer 48, 31763184.

[10] Mudawar I. and Estes K. A., 1996. Optimizing and Predicting CHF in Spray Cooling of a Square Surface.

J. Heat Transfer 118, 672-679.

[11] Visaria M. and Mudawar I., 2008. Theoretical and experimental study of the effects of spray inclination on

two-phase spray cooling and critical heat flux. Int. J. Heat Mass Transfer 51, 23982410.

[12] Lin L. C. and Ponnappan R., 2004. Critical Heat Flux of Multi-Nozzle Spray Cooling. J. Heat Transfer 126,

482-485.

[13] Wendelstorf, J., Spitzer, K. H., Wendelstorf R., 2008. Spray water cooling heat transfer at high temperature

and liquid mass fluxes. Int. J. Heat Mass Transfer 51, 4902-4910.

http://dx.doi.org/10.1016/j.ijheatmasstransfer.2008.01.032 .

[14] Al-Ahmadi, H. M., Yao, S. C., 2008. Spray cooling of high temperature metals using high mass flux

industrial nozzles. Exp. Heat Transfer 21, 38-54.

[15] Mascarenhas, N., Mudawar, I., 2012. Methodology for predicting spray quenching of thick-walled metal

alloy tubes. Int. J. Heat Mass Transfer 55, 2953-2964.

http://dx.doi.org/10.1016/j.ijheatmasstransfer.2012.02.018.

22
Page 22 of 35
[16] Yao, S. C., Cox, T. L., 2002. A general heat transfer correlation for impacting water sprays on high-

temperature surface. Exp. Heat Transfer 15, 207-219.

[17] Choi, K. J., Yao, S. C., 1987. Heat transfer mechanisms of horizontally impacting sprays. Int. J. Heat Mass

Transfer 30, 1291-1296.

[18] Deb, S., Yao, S. C., 1989. Analysis on film boiling heat transfer of impacting sprays. Int. J. Heat Mass

Transfer 32, 2099-2112. http://dx.doi.org/10.1016/0017-9310(89)90117-8.

[19] Klinzing, W. P., Rozzi, J. C., Mudawar, I., 1992. Film and transition boiling correlations for quenching of

hot surfaces with water sprays. J. Heat. Treat. 9(2), 91-103.

[20] Bernardin, J. D., Mudawar, I., 1997. Film boiling heat transfer of droplet streams and sprays. Int. J. Heat

Mass Transfer 40, 2579-2593. http://dx.doi.org/10.1016/S0017-9310(96)00297-9.

[21] Dukowicz, J. K., 1980. A particle fluid numerical model for liquid sprays. J. Comp. Phys, 35, 229-253.

[22] Amsden, A. A., ORourke, P. J., Butler, T. D., 1989. KIVA-II: A Computer Program for Chemically

Reactive Flows with Sprays. Los Alamos National Laboratory, LA-11560-MS.

[23] Varnavas, C., Assanis, D. N., 1996. A high temperature and high pressure evaporation model for the

KIVA-3 code. SAE 960629.

[24] Abramzon, B., Sirignano, W. A., 1989. Droplet vaporization model for spray combustion calculations. Int.

J. Heat Mass Transfer 32, 1605-1618. http://dx.doi.org/10.1016/0017-9310(89)90043-4.

[25] Abramzon, B., Sirignano, W. A., 1987. Approximate Theory of a Single Droplet Vaporization in a

Convective Field: Effects of Variable Properties, Stefan Flow and Transient Liquid Heating. Proc. in the 2nd

ASME-JSME Thermal Engineering Joint Conference, 11-18, Honululu, Hawai.

[26] Drew, D. A., Passman, S. L., 1998. Theory of Multicomponent Fluids, Springer, New York.

[27] Launder, B. E., Spalding, D. B., 1974. The numerical computation of turbulent flows. Comput. Meth. Appl.

Mech. Eng. 3, 269-289. http://dx.doi.org/10.1016/0045-7825(74)90029-2.

[28] Hanjalic, K., Popovac, M., Hadziabodic, M., 2004. A robust near-wall elliptic-relaxation eddy-viscosity

turbulence model for CFD. Int. J. Heat Fluid Flow 25, 1047-1051.

http://dx.doi.org/10.1016/j.ijheatfluidflow.2004.07.005.

[29] Basara, B., 2006. An Eddy Viscosity Transport Model Based on Elliptic Relaxation Approach. AIAA

Journal 44 (7), 1686-1690.

23
Page 23 of 35
[30] Basara, B., Krajnovic, S., Girimaji, S., Pavlovic Z., 2011. Near-Wall Formulation of the Partially Averaged

Navier-Stokes Turbulence Model. AIAA Journal 49 (12), 2627-2636.

[31] Durbin, P. A., 1991. Near-wall turbulence closure modelling without damping functions. Theor. Comput.

Fluid Dyn. 3, 113.

[32] Lee, S. J., Lahey Jr., R. T., Jones Jr., O. C., 1989. The Prediction of Two-Phase Turbulence and Phase

Distribution Phenomena Using a k-e Model. Japanese J Multiphase Flow 3 (4), 335-368.

[33] Greif, D., Sampl, P., Edelbauer, W., 2014. Cavitating injector flow simulations considering longitudinal

and lateral needle displacement. , International Journal of Automotive Engineering 5, 85-90.

[34] Von Berg, E., Edelbauer, W., Alajbegovic, A., Tatschl, R., Volmajer, M., Kegl, B., Ganippa, L. C., 2005.

Coupled Simulations of Nozzle Flow, Primary Fuel Jet Breakup, and Spray Formation. J. Eng. Gas

Turbines Power 127 (4), 897-907.

[35] Von Berg, E., Alajbegovic, A., Tatschl, R., Krueger, C., Michels, U., 2001. Multiphase Modelling of Diesel

Sprays with the Eulerian/Eulerian Approach. Proc. in 17th European Conference on Liquid Atomisation and

Spray Systems ILASS01, Zrich, Switzerland, September.

[36] Edelbauer, W., Suzzi, D., Sampl, P., Tatschl, R., Krueger, C., Weigand, B., 2006. New Concept for On-

Line Coupling of 3D Eulerian and Lagrangian Spray Approaches in Engine Simulations. Proc. in 10th Int.

Conf. on Liquid Atomisation and Spray Systems, ICLASS06, Kyoto, Japan, September.

[37] Vujanovi, M., Petranovi, Z., Edelbauer, W., Baleta, J., Dui, N., 2015. Numerical modelling of diesel

spray using the Eulerian multiphase approach, Energy Conversion and Management 104, 160-169.

http://dx.doi.org/10.1016/j.enconman.2015.03.040.

[38] Petranovi, Z., Edelbauer, W., Vujanovi, M. , Dui, N., 2013. Validation of Eulerian-Eulerian Approach

for Diesel Sprays, and 3D Coupling with Lagrangian Spray Approach. Proc. in 25th European Conference

on Liquid Atomization and Spray Systems, ILASS13, Chania, Greece, September.

[39] Petranovi, Z., Edelbauer, W, Vujanovi, M., Dui, N., 2014. The ORourke droplet collision model for the

Euler-Eulerian framework, Proc. in 26th European Conference on Liquid Atomization and Spray Systems,

ILASS14, Bremen, Germany, September.

[40] Pola A., Gelfi M. and Vecchia G. M La (2013). Simulation and validation of spray quenching applied to

heavy forgings. J. Materials Processing Technology, Vol, 213(12): 2247-2253

24
Page 24 of 35
[41] ORourke, P. J., 1981. Collective Dop Effects on Vaporizing Liquid Sprays. Ph. D. Thesis, Los Alamos

Lab. Report LA-9069 T.

[42] Lopez de Bertodano, M.A., 1998. Two fluid model for two-phase turbulent jets, Nucl. Eng. Des. 179,

65-74.

[43] Edelbauer, W., 2014. Coupling of 3D Eulerian and Lagrangian Spray Approaches in Industrial Combustion

Engine Simulations. Journal of Energy and Power Engineering 8, 190-200.

[44] Ranz, W. E., Marshall Jr., W. R., 1952. Evaporation from drops Part II. Chem. Eng. Prog. 48 (4),

172-180.

[45] Kopun, R., Greif, D., Kovai, Z., Tatsch, R., 2013, Numerical Investigation of Immersion Quenching

Process for Cast Aluminum Parts Using an Eulerian Multi-Fluid Approach, Proc. in 27th ASM Heat

Treating Society Conference, Indianapolis, Indiana, USA, September.

[46] Li, Z., Lynn Ferguson, B., Greif, D., Kovacic, Z., Urbars, S., Kopun, R., 2015. Coupling CFD and oil

quench hardening analysis of gear components, Proc. in 28th ASM Heat Treating Society Conference,

Detroit, Michigan, USA, October.

25
Page 25 of 35
Appendix A: Evaporation on Single Droplet under Typical Conditions

The aim of the two test cases presented in this section is to demonstrate the magnitude of the relevant time scales of

the evaporation/condensation process at typical spray quenching conditions. Therefore, the evaporation model

described in section 2.4 has been implemented as single droplet model into an explicit ordinary differential equation

(ODE) solver. There the ambient gas conditions, Yamb and Tg, are kept constant, and the evolution of droplet

diameter and temperature are calculated.

The conditions of test case 1 are summarized in Table A.1. The initial droplet diameter is 10 m and the initial

temperature is 20 C. The ambient gas temperature is 300 C and there is no ambient vapor. These conditions are

typical for droplets entering the hot gas layer next to heated solid surface during the spray quenching process.

Figure A.1 shows the time evolutions of droplet temperature and diameter, and the Spaldings heat and mass transfer

numbers, BT and BM. Positive values of BT and BM indicate evaporation, negative condensation. One can observe that

there is a very fast change of the droplet temperature from 293.15 to approximately 326 K. Then the liquid

temperature as well as BT and BM stay constant when the so-called wet bulb temperature is reached. In this period

heat and mass transfer are in balance, meaning that the energy transferred from the gas to the droplet is fully

consumed by the mass transfer and the latent heat. Consequently the first and second term on the right-hand side of

equation (15) are in balance, and is zero. During this time the d 2-law, meaning that the time for evaporation is

proportional to the square of the droplet diameter, is valid. After 0.004 s the droplet is completely evaporated

indicating that the time for evaporation is in the same order as the desired time step size for simulation.

The initial conditions of the second test case are summarized in Table A.2 and the results are shown in Figure A.2.

There the initial droplet diameter is again 10 m, initial droplet temperature is 45 C, and the ambient gas

temperature has only 80 C but there is a considerable amount of vapor in the ambient gas. With these conditions

condensation takes place at the initial phase until the droplet temperature reaches equilibrium at approximately

1.510-4 s. In this initial phase BT and BM are negative. Then again balance between heat and mass transfer at the wet

bulb temperature is reached, and the droplet starts to evaporate without further temperature change indicated by

constant positive Spaldings heat and mass transfer numbers. Since the temperature difference between gas and

droplet of test case 2 is much smaller than that of case 1, the evaporation is slower.

26
Page 26 of 35
Initial droplet diameter dd 10 m
Initial droplet temperature Td 45 C (293.15 K)
Ambient gas temperature Tg 300 C (573.15 K)
Ambient vapor mass fraction Yamb 0.0
Table A.1: Single droplet evaporation for test case 1

Initial droplet diameter dd 10 m


Initial droplet temperature Td 45 C (318.15 K)
Ambient gas temperature Tg 80 C (353.15 K)
Ambient vapor mass fraction Yamb 0.137
Table A.2: Single droplet condensation for test case 2

Figure A.1: Droplet radius and temperature evolution (left) and Spalding heat and mass transfer number (right) of a

single droplet with initial temperature of 20 C and constant ambient conditions of 300 C and Yamb=0.0.

27
Page 27 of 35
Figure A.2: Droplet radius and temperature evolution (left) and Spalding heat and mass transfer number (right) of a

single droplet with initial temperature of 45 C and constant ambient conditions of 80 C and Yamb=0.137.

Appendix B: Mesh Dependency Test

A grid size dependency test on a quasi-two-dimensional sector mesh has been performed to check the required

resolution. The simulation set-up is similar to that described in section 3.2. Three different meshes, the coarse

Mesh 1, the finer Mesh 2 and the finest Mesh 3, as depicted in Figure B.3, are applied for this test. The solid part,

marked with red color in Figure B.3, is cooled by the spray injected from the right hand side. The mesh resolution of

the solid part and the mean cell size, determined by the square root of the element surface area in the solid cross-

section, are listed in Table B.3. The height of the boundary layer on the fluid side is adopted according to the mesh

resolution of the solid part.

28
Page 28 of 35
Figure B.3: Computational meshes for the grid size dependency test: coarse Mesh 1 (left), fine Mesh 2 (middle),
finest Mesh 3 (right).

Mesh 1 Mesh 2 Mesh 3


Resolution of solid part 14 x 7 29 x 14 31 x 20
Mean cell size (mm) 5.5 2.7 2.2
Table B.3: Specification of mesh dependency test

The cooling curves of the three different mesh configurations are shown in Figure B.4. The monitoring point for this

comparison is located in the center of the solid part. While the case with Mesh 1 shows faster cooling, the cases with

Mesh 2 and Mesh 3 show very similar cooling behavior. This means that the resolution of these meshes is sufficient.

For the simulations presented in section 3, a cell size of 2 mm has been chosen.

Figure B.4: Comparison of the cooling curves at a certain monitoring point for three meshes with different
resolution.

29
Page 29 of 35
Figure 5: Flowchart of the evaporation procedure.

Figure 6: Heat transfer coefficient (left) and heat flux (right) for three droplet mass fluxes of V s1=3, Vs2=15 and
Vs3=30 kg/m2s.

30
Page 30 of 35
Figure 7: Rectangular solid used in the experiment with the position of the thermocouples (a) and the test station for
spray quenching measurements (b).

Figure 8: Computational domain set up with boundary conditions (a) and positions of the monitoring points (b).

31
Page 31 of 35
Figure 9: Comparison of experimental (Exp) and numerically predicted solid part temperature for non-frozen fluid
flow (CFD_nFF) at monitoring points (a) T1, (b) T5, and (c) T7.

Figure 10: Total liquid volume fraction distribution at different time steps: (a) 0.1 s, (b) 10 s, (c) 50 s, (d) 100 s, (e)
150 s, (f) 200 s in cuts along the spray axis.

32
Page 32 of 35
Figure 11: Solid temperature and total liquid volume fraction distributions at different time steps: (a) 0.1 s, (b) 10 s,
(c) 50 s, (d) 100 s, (e) 150 s, (f) 200 s.

33
Page 33 of 35
Figure 12: Temperature, heat transfer coefficient and heat flux at different time steps for the second droplet phase
with 37 m class diameter: (a) 0.04 s, (b) 5 s, (c) 100 s.

Figure 13: Comparison of part temperature histories between non-frozen (CFD_nFF) and frozen fluid flow
(CFD_FF) at monitoring points (a) T1, (b) T5, and (c) T7.

34
Page 34 of 35
Table 4: Overview about some experimental studies on spray cooling and their characteristic parameters.

Researcher Droplet flux Droplet size (mm) Cooling area


Al Ahemed and Yao, 2008 [14]* 1.5 6.6 kg/(m2s) 6.252 cm2
0.0147 0.0204 m3/(m2s) 0.036 0.045 (FC-87)
Lin and Ponnappan, 2004 [12] 0.0144 0.0201 m3/(m2s) 0.041 0.052(FC-72) 1 cm 2 cm
0.0246 0.0363 m3/(m2s) 0.062 0.080 (Methanol)
Mudawar series [10][11][15]** 0.0006 0.00996 m3/(m2 s) 0.405 1.351
Rini, Chen and Chow, 2002 [7] 0.0166 0.0563 m3/(m2 s) 0.102 0.113 (FC-72) 1 cm 1cm
Shedd and Pautsch, 2005 [9] 0.000533 0.02147 m3/ (m2 s) 81.5 cm 1.5 cm
Wendelstorf et al., 2007 [13] 3 30 kg/(m2 s) 0.3 0.4 52 cm2
Yao and Cox, 2002 [16]*** 0.016 2.05 kg/(m2 s) 0.13 25
*Al- Ahemed and Yao(2008) used full cone industry spray HHX-12 FullJet.

**Mudawar published a series of papers on spray cooling and or quenching, cooling area varied in each paper.

***Yao and Cox (2002) provided the correlation by compiling the available experimental data.

If not mentioned, water droplets were used in the abovementioned work.

Table 5: Phase specification of the Eulerian spray model

Phase 1 2,..,nph
Content Gas Droplets
Table 6: Experimental set-up

Specimen dimensions (length / thickness / height) 500 / 30 / 60 mm


Normal distance injector - specimen 75 mm
Type of twin fluid atomizer (water air) Hennlich 136.208.35.12
Water pressure / volume flow rate of atomizer 0.3 MPa / 0.19 l/min
Air pressure / volume flow rate of atomizer 0.4 MPa / 147 l/min
Number of nozzle holes / total spray angle 6 / 60 deg
Initial solid temperature 500 C (773.15 K)
Cooling time 250 s
Table 7: Simulation set-up

Dimensions of computational domain 300 / 300 / 500 mm


(width/length/height)
Species of the gaseous phase H2O, O2, N2
Initial gas temperature / pressure 20 C / 0.1 MPa
Number of droplet size classes 3
Injection velocity at inlet (gas and liquid) 2.66 m/s
Droplet phase diameter at inlet boundary 13 37 74 m
Liquid volume fraction at inlet boundary 2.7810-3 6.1510-3 1.7910-3
Droplet temperature at inlet 20 C
Correction factor CTC 9.0
Maximum time step 0.01 s
Simulation time 200 s

35
Page 35 of 35

You might also like