Вы находитесь на странице: 1из 8

Proceedings of the Sixth International ASME Conference on Nanochannels, Microchannels and Minichannels

ICNMM2008
June 23-25, 2008, Darmstadt, Germany

ICNMM2008-62244
NUMERICAL STUDY OF MARANGONI THERMOCAPILLARY CONVECTION INFLUENCE
DURING BOILING HEAT TRANSFER IN MINICHANNELS
Anthony J. Robinson
Trinity College Dublin, Ireland
Mechanical and Manufacturing Department
arobins@tcd.ie

Cristina Radulescu
Trinity College Dublin, Ireland
Mechanical and Manufacturing Department
radulesc@tcd.ie

ABSTRACT
Marangoni thermocapillary convection and its contribution
to heat transfer during boiling has been the subject of some
debate in the open literature. Currently, for certain conditions,
such as microgravity boiling, is being shown that has a
significant contribution to heat transfer [1]. Typically, this
phenomenon is investigated for the idealized case of an isolated
and stationary bubble resting atop a heated solid which is
immersed in a semi-infinite quiescent fluid or within a twodimensional cavity. However, little information is available
with regard to Marangoni heat transfer in miniature confined
channels in the presence of a cross flow. As a result, this paper
presents a numerical study that investigates the influence of
steady thermal Marangoni convection on the fluid dynamics
and heat transfer around a bubble during laminar flow of water
in a minichannel with the view of developing a refined
understanding of boiling heat transfer for such a configuration.
This mixed convection problem is investigated for channel
Reynolds numbers in the range of 0 Re 500 and Marangoni
numbers in the range of 0 Ma 17114. The influence of the
thermocapillary flow is most pronounced for low Re and high
Ma numbers showing an average of 40% increase in heat
transfer. For low Ma and high Re inertial effects dominate and
the thermocapillary effect is not as noticeable. However, the
disruption of the fully developed flow does tend to enhance the
heat transfer at the expense of additional pressure drop.
INTRODUCTION
Based on the experimental results published in 1855 by
Thomson [2], C. G. M. Marangoni [3] later offered a viable
explanation of the effect of surface tension on drops of one
liquid spreading upon another [4]. Subsequent to this several
numerical and experimental studies [5, 6, 7] established that
thermocapillary Marangoni convection is in fact real physical
phenomenon at gas-liquid and liquid liquid interfaces
resulting from gradients in surface tension. The surface tension
gradients can be brought about by variations in the liquid
concentration or temperature. Once the existence of this

phenomenon was confirmed the main focus of the scientific


research work was to quantify the impact on the heat transfer
enhancement. Starting with the experimental results of
McGrew et. al. [8] and followed by the early analytical work of
Larkin [9], thermal Marangoni convection and its contribution
to heat transfer during boiling became the subject of some
debate in the open literature [10]. Recently, it has been
established that for certain conditions, such as microgravity
boiling, the thermocapillary induced flow has associated with it
a significant enhancement of heat transfer due to the liquid flow
in the vicinity of the interface [11, 12 and 13]. Despite the
research conclusions presented in literature there is still
insufficient information available with regard to Marangoni
flow contribution on the heat transfer in miniature confined
channels [14, 15] especially when it takes place in the presence
of a cross flow [16].
Boiling heat transfer in minichannels has become an
increasingly important topic due to its application in the
compact heat exchanger design such as those required for
electronics thermal management. Typically, for electronic
cooling applications involving two-phase flow, nucleate boiling
is the preferred regime of operation because the small increase
in wall superheat is accompanied by a disproportionately large
increase in the wall heat flux [17, 18]. Apart from the high rates
of heat transfer at relatively low volumetric flow rates the
isothermal nature two-phase convective boiling makes this a
very attractive technology in contrast with single-phase channel
cooling.
The objective of this paper is to provide qualitative
information regarding the fluid motion and the influence on
heat transfer around the centerline of a bubble placed on the
bottom heated wall of a rectangular section minichannel
(1x1x20 mm) in a cross flow configuration as illustrated in
Figure 1. This mixed convection problem is investigated for
channel Reynolds numbers in the range of 0 Re 500 by
increasing the inlet mass flow rate. For a fixed inlet temperature
the Marangoni number is varied in the range of 0 Ma 17114
by increasing the wall temperature of the channel. Furthermore

Copyright 2008 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 04/05/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

the influence of the bubble dimension on the flow pattern and


heat transfer is taken into consideration for the following
geometrical characteristics: Rb/H = 0.1 (B_1), Rb/H = 0.5 (B_2)
and Rb/H = 0.75 (B_3).

Figure 1. Physical domain showing a hemispherical bubble near the


entrance of the minichannel

NOMENCLATURE
B_1
first bubble under consideration
B_2
second bubble under consideration
B_3
third bubble under consideration
Ca
capillary number
hMa
height of Ma recirculation cell
[m]
H
height of the channel
[m]
k
thermal diffusivity
[m2/s]
lMa
length of Ma recirculation cell
[m]
L
length of the minichannel
[m]
Ma
Marangoni number (thermocapillary)
n
unit normal vector (height of the boundary)
p
static pressure
[N/m2]
Pr
Prandtl number
Rb
bubble radius (RB_1 RB_2 RB_3 radius of
B_1, B_2, B_3 respectively)[m]
Re
Reynolds number
lVref Rb
[16]
R
surface tension Re number
l
Tm
avg bulk liquid temperature at inlet [C]
Twall
heated wall temperature
[C]
Vavg
liquid inlet average velocity
[m/s]
Vref
reference thermocapillary velocity
T | T
[16]
[m/s]
l
u
velocity vector
X, y
Cartesian coordinates
T
temperature difference Twall - Tm [C]

Greek Symbols

stagnation point angle

dynamic viscosity

liquid surface tension


T
surface tension gradient

density

kinematic viscosity
Laplace divergence operator

Subscripts
avg
b
g
l
m
Ma
ref
T

average
bubble
gas
liquid
bulk liquid
Marangoni
reference
tension

[Ns/m ]
[N/m]
[N/mK]
[kg/m3]
[m2/s]

MATHEMATICAL FORMULATION
Figure 1 shows the simplified schematic of the channel
through which water at a mean inlet temperature Tm flows with
the average velocity Vavg. The flow is assumed to be
hydrodynamically fully developed at the inlet with a parabolic
velocity profile. The heated bottom wall is maintained at a
constant temperature - Twall while the top wall is considered to
be insulated. The hemispherical gas bubble is situated on the
heated wall creating a cross flow configuration due to the bulk
liquid flow directed perpendicular to the bubble axis. This is
located near the entrance of the channel since this is the
expected region of the nucleate boiling flow regime [19] with
stratified or slug flow regimes being more likely downstream. It
was concluded that the bubble nuclei grow slowly to visible
size in the laminar inlet flow [19]. As a result the flow and heat
transfer problem has been simplified by considering steady
state conditions for three different dimensions of the bubble
starting with the incipient stage of growth and at the final stage
before bubble sliding is anticipated [14]. The bubble shape
deformation is neglected based on the assumption that the
capillary number Ca = Vavg / T is much less that unity.
Lastly, it is well known that the flow around the bubble
within a small channel is inherently a threedimensional
problem. However, at the mid-plane of the spherical bubble the
flow is approximately twodimensional. In this respect the
problem can be treated qualitatively as a twodimensional
phenomenon as consistent with the work of Bhunia and
Kamotani [16].
Governing Equations
The governing continuity, momentum and energy
equations for steady flow are [22, 23 and 24]:
The continuity equation is solved in the following form:
u=0
Eq. 1
Conservation of momentum is described by Eq. 2:

Eq. 2
p + 2 u

The energy equation without internal heat generation is reduced


at Eq. 3:

u u =

u T = k 2 T

Eq. 3

The governing equations were solved subject to the following


assumptions and boundary conditions: i) the bubble is
represented as a hemispherical gas-liquid interface, ii) the heat
flux zero at the bubble interface, iii) the top wall is insulated,
iv) the inlet flow is hydro-dynamically fully developed, v) the
bottom heated wall is at a constant temperature, vi)
gravitational affects are negligible. vii) temperature variations
in physical properties are not considered, viii) the liquid is
incompressible and ix) the shear stress in the liquid at the gasliquid interface of the bubble is balanced by the surface tension
such that the boundary condition at the bubble interface is a no
slip and a directly applied Marangoni stress given as [24, 27]:

u n = 0

u
=
T
n T

Eq. 4

Here, n is the unit vector normal to the bubble interface.

Copyright 2008 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 04/05/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

The governing equations were solved numerically with the


CFD package Fluent Version 6.3.26 [20] and the physical
domain and grid was created in the Gambit Version 2.2.30 [21].
Cartesian coordinates were used with a non-uniform grid of
14400 cells and 29420 faces. In order to resolve the flow and
temperature fields accurately, grid clustering near the bubble
was implemented. The accuracy of the resulting simulations
was been confirmed by reproducing the numerical results of
Bhunia and Kamotani [16].
RESULTS AND DISSCUSION
The results presented in this study have been carried out
for water which has a surface tension gradient of d/dT= 0.1477x10-3 N/mK. The problem is investigated for channel
Reynolds numbers in the range of 0 Re 500 by increasing the
inlet average velocity between 0.1m/sRe0.5m/s. The
Marangoni number is varied in the range of 50 Ma 17114 by
increasing the difference between the inlet temperature and the
wall temperature of the channel (T= Twall-Tm) between 1C
Tw30C. Consistent with the works of [25, 26 and 27] the
Marangoni number has been defined as:
Ma =

( / T )(Twall Tm ) Rb2
= R Pr

strong enough to cross over the line of symmetry of the bubble


towards the front region. For the highest Marangoni number
obtained for Rb/H = 0.1 (i.e. Ma=300) the high shear rate at the
front bubble interface interacts with the strong rear vortex to
form a recirculation cell near front region of the bubble. With
regard to the rear vortex itself increasing Ma has the effect of
increasing the strength of the vortex, as is evident from the
increasing concentration of the streamlines, as well as
increasing the vortex size as it is seen to penetrate deeper into
the bulk liquid and along the heated wall as well as moving
forward along the bubble surface. In order to visualize the
velocity magnitude on x coordinate, Figure 3 shows the case of
Ma = 100 together with an information of the geometrical
dimensions of the recirculation thermocapillary cell.

Eq. 5

To approximate different stages of bubble growth i.e.


nucleation to bubble sliding [14], the bubble size relative to the
channel height has been investigated for Rb/H=0.1 (B_1),
Rb/H=0.5 (B_2) and Rb/H=0.75 (B_3).
The primary objective of this study is to provide a
qualitative description of the effect of thermocapillary
convection on the flow field and heat transfer during bubble
growth in a miniature channel.
The Effect of Marangoni Convection on the Flow Field
Figure 2 presents the streamlines for steady flow around
the bubble for the case Rb/H = 0.1, Re= 100 and Ma = 0, 50,
100, and 300. To provide a baseline case for comparison,
steady flow around a bubble with no thermocapillary effect has
been simulated for this test case and each test case to follow.
This is equivalent to imposing a constant surface tension
(/T=0) such that Ma=0 even though there are temperature
gradients along the interface. Due to these, the surface tension
is highest near the top of the bubble and lowest near the heated
wall. This surface tension variation generates thermocapillary
flow along the bubble surface away from the hot wall towards
the bulk liquid. Figure 2 illustrates clearly that increasing the
driving potential for thermocapillary flow, which in this
situation is T= Twall-Tm, the influence of the surface tension
driven flow becomes stronger which is apparent from the
increased deformation of the streamlines as compared with the
baseline Ma=0 case.
Upstream (front) side of the bubble: in this region the
thermocapillary action accelerates the liquid flow along the
bubble surface. The shear driven flow at interface has the effect
of drawing the relatively colder bulk liquid downward towards
the front corner of the bubble as apparent from the deformation
of the near wall streamlines towards the hot front corner of the
bubble for the Ma 50 and 100 cases. Downstream side of the
bubble: in this region a sizable vortex is formed even at low
Marangoni numbers (Ma=50) when the recirculation cell is

Figure 2: Streamlines of steady flow for Rb/H = 0.1(B_1) at Re=100

Near the downstream end the surface velocity due to the


bulk liquid flow it is in an opposite direction with the
recirculation cell created by the thermocapillary flow. The point
on the bubble surface where the forward and reverse flow meet
is known in literature as the stagnation point [16].
As is evident in Figure 2, and which will be discussed later
in more detail, this separation point appears to move closer to
the front region of the bubble with increasing Ma. Figure 3
presents the influence of the bubble dimension (B_1 and B_3)
on the stagnation point position for the same Re = 100 and
temperature difference = 10C. It is noticed that this point is
situated closer to the front region of the bubble (at higher
separation angle ) for the smaller bubble dimension (B_1).

Copyright 2008 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 04/05/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Figure 3. Magnitude of x-velocity and stagnation point position for


B_1, B_3 at Re 100 and T=10C

Figure 4 illustrates the influence of inertial effects of the


channel flow by considering the identical configuration as in
Figure 2 but for the higher Reynolds number case of Re=300.

Figure 4: Streamlines of steady flow for Rb/H = 0.1(B_1) at Re=300

As one would expect the increase in the cross-flow


velocity has an important impact on the flow structure by
suppressing the influence of the Marangoni flow. This is clear
considering that for Re=300 it takes a Marangoni number of

Ma=300 to roughly reproduce the flow structure that a


Marangoni number of Ma=100 was able to produce for
Re=100.
With the view of future development of a dynamic bubble
growth model it seemed instructive to investigate the influence
of the bubble size on the flow and heat transfer within the
channel for this idealized case of steady two-dimensional flow.
Figures 5 and 6 show the simulated flow patterns around
bubbles with aspect ratios Rb/H=0.5 and 0.75 respectively for a
fixed Reynolds number of Re = 100.
Increasing the relative size of the bubble from Rb/H = 0.1
to Rb/H = 0.5 has a notable influence on the flow pattern as
evident from Figure 5. It must first be noted that the Ma
number increases disproportionately compared with the
increase in T and Rb/H since the length scale has been chosen
as Rb2/H for this study to incorporate the influence of
confinement on the flow and heat transfer. Compared with the
relatively unconfined case depicted in Figure 2 for Rb/H = 0.1,
the flow structure within the liquid for the more confined case
of Rb/H = 0.5 indicates that the thermocapillary induced
convection has a more profound influence on flow for a like
driving temperature differential and a significantly higher Ma.

Figure 5: Streamlines of steady flow for Rb/H = 0.5(B_2) at Re=100

The presence of the confining upper wall tends to form


elongated yet more concentrated recirculation zones at the
downstream end of the bubble compared with the more
unconfined case at identical T. For T=30C, Figure 5 shows

Copyright 2008 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 04/05/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

that confinement effects result in the formation of three


recirculation zones, in particular, an elongated vortex spanning
a considerable portion of the top region of the bubble interface.
Figure 6 presents the extreme situation of Rb/H = 0.75
where the bubble is at its maximum growth dimension before
sliding in the minichannel due to the inertial forces of the bulk
liquid flow would be expected [14]. It is noticed that the flow
pattern is altered notably compared with the previous two
cases. In this case the confining and adiabatic top wall tends to
restrict the rear vortex from encroaching on the frontal region
with increasing T. This pinching of the rear vortex is strong
enough that for the highest driving potential of T=30C no
secondary recirculation zones develop since the separation
point is far enough back on the bubble interface that it does not
significantly obstruct the frontal shear flow. This is also
illustrated in Figure 3 where it is shown that the stagnation
point on the bubble surface, where the forward and reverse
flows meet, has the tendency to occur at smaller separation
angles, , for the larger bubbles at identical T and Re.

effect on less global parameters such as the stagnation angle,


bubble surface temperature distribution and resulting wall heat
transfer in the vicinity of the bubble.
Figure 7 illustrates the thermal profile for Rb/H = 0.1,
Re=100 and Ma=0, 50 and 300 case for it is noticed the higher
values of the temperature profile liquid boundary layer formed
close to the heated wall. For Ma=0 the presence of the bubble
has a small influence on the thermal field, acting as a simple
obstruction to the flow. However, for the same wall
temperature but with Marangoni convection, a non-symmetric
jet of warm fluid is forced into the bulk of the flow. The size of
the warm jet is consistent with the size of the vortices observed
in Figure 2 and the penetration depth of the warm jet increases
with Ma in the same way as the size of the recirculation regions
increased in Figure 2. Near the front edge of the bubble it is
clearly evident that the deformation of the streamlines observed
in Figure 2 has associated with it the drawing in of the cooler
bulk liquid which will have important implications with regard
to the wall heat transfer.

Figure 7: Temperature profile for Rb/H = 0.1 at Re= 100 and Ma=0,
50 and 300.
Figure 6. Streamlines for steady flow for Rb/H = 0.75(B_3) at Re=100

The Effect of Marangoni Convection on the Thermal Field


and Wall Heat Transfer
The flow and thermal fields are directly coupled via the nonlinear convection term in Eq. 3 and more indirectly through the
Marangoni stress boundary condition given in Eq. 4. As a
result, the interaction of the flow and thermal fields must be
understood in relation to one another in order to elucidate the

Figure 8 (a, b) presents the wall heat flux distribution for


Rb/H = 0.1, Re=100 and 300 with Ma=30, 50, 100 and 200. The
corresponding heat flux distribution for the situation of no
Marangoni flow is also plotted for each temperature
differential. It is clear that the thermal and flow fields resulting
from thermocapillary convection have a direct impact on the
heat transfer in the vicinity of the bubble as is evident from the
peaks in the wall heat flux that appear around the bubble.

Copyright 2008 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 04/05/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

from a combined convection region to a dominantly


thermocapillary driven recirculation region where the average
liquid temperature is generally much higher than the bulk
liquid. Increasing the Marangoni number has two notable
effects on the bubble surface temperature profile. First the
stagnation point moves forward with initial increases in Ma.
However, increasing Ma beyond Ma=200 has a minimal effect
on the position of the separation point. Secondly, increasing Ma
tends to flatten the temperature profile along the bubble
interface with large gradients isolated to the front and rear of
the bubble.

(a) B_1 at Re=100 and Ma = 30, 50, 100, 200

(b) Heat flux [kW/m2] for B_1 at Re 300 and Ma = 30, 50, 100, 200
Figure 8: Wall heat flux for B_1 at Re = 100 and 300.

Upstream (front) side of the bubble: in this region the


increase in the heat flux is stronger due to the combined effect
of forced and thermocapillary convection with the
thermocapillary component drawing the cooler bulk liquid
towards the heated wall and thinning the thermal boundary
layer as depicted in Figures 2, 4 and 7. For Re=100 a two to
three fold increase in the local heat flux is evident. As expected,
the enhancement decreases somewhat with increasing Re,
however it is still substantial. Downstream of the bubble the
increase in the wall heat transfer is only evident for the higher
Ma because it is primarily the warm liquid that was extracted
from the frontal region that is being recirculated in this area.
Here the local heat flux increases by a factor of nearly 1.2 for
Re=100 and Ma=200 and tends to improve with increasing Re
with a 1.4 times improvement for Re=300 and Ma=200.
The bubble dimensionless interface temperature
distributions for Rb/H = 0.1, Re=100 and Ma=0, 30, 50, 100,
200 and 300 are plotted in Figure 9. Considering the Ma=0 and
Ma=50 case which both correspond with T=5C it is clear that
the Marangoni convection tends to diminish the thermal
gradients thus reducing the magnitude of T/ over the majority of
the bubble surface. For Ma>0 the general shape of the
temperature profiles are similar. In the front region of the
bubble the gradients are steep due to the combined influences
of the forced and thermocapillary convection as cold bulk fluid
is drawn towards the surface and accelerated along it. The
surface temperature decreases to a minimum at the stagnation
point. Behind the stagnation point the temperature gradients
along the bubble surface are less steep as the flow transitions

Figures 9: Bubble interface temperature B_1 for Re=100 and Ma = 0,


30, 50, 100, 200 and 300

Figure 10 shows the effect of the top confining wall on the


thermal field around the bubble for which Rb/H = 0.5 for the
same T values in Figure 7.

Figure 10: Temperature distribution for B_2 at Ma =0, 1260, 7600.

Copyright 2008 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 04/05/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Compared with the unconfined small bubble, the confinement


effects tend to keep the warm recirculation zone near the rear of
the bubble with strong effects on the front region where the
shear flow causes entrainment of the cold bulk fluid towards
the wall which thins the thermal boundary layer in this region
more so than for the Rb/H = 0.1 case.
Figure 11 illustrates this point more clearly where the heat
flux profiles for Rb/H = 0.1, 0.5 and 0.75 are plotted for
Re=100 and T=10C. It is clear that the peak heat flux at the
front of the bubble increases notably with increasing Rb/H. The
peak local heat transfer enhancement at the rear of tends to
increase with increasing Rb/H but to a much lesser extent.

REFERENCES
[1] J. Straub, 1994. The Role of Surface Tension for TwoPhase Heat and Mass Transfer in the Absence of Gravity,
Experimental Thermal and Fluid Science, No. 9, pp. 253-273.
[2] J. Thomson, 1855. Philos Mag, 10:330.
[3] C.G.M. Marangoni, 1871. Ann Phys Chem (Poggendorff),
143:337.
[4] C.G.M. Marangoni, 1865. Tipografia Fusi, Pavia.
[5] H. Bernard, 1901. Ann Chem Phy, 23:26.
[6] H. Bernard, 1900. Ann Chem Phys, 23:62
[7] J.R.A. Pearson, 1958. J. Fluid Mech, 4:489.
[8] J.L. McGrew, F.L. Bamford and T.R. Rehm, 1966.
Marangoni Flow: An Additional Mechanism in Boiling
Heat Transfer, Science, New Series, Vol. 153, No. 3740,
pp. 1106 1107.
[9] B. K. Larkin, 1970. Thermocapillary flow around a
hemispherical bubble, AIChEJ 16, pp. 101-107.
[10] K. Sefiane and C.A. Ward, 2007. Recent Advances on
thermocapillary flows and interfacial conditions during the
evaporation of liquids, Advances in Colloid and Interface
Science 134 125, pp. 201 223.
[11] Y.S. Koo and D.B.R. Kenning, 1972. Thermocapillary
flow near a hemispherical bubble on a heated wall, Journal
Fluid Mechanics, Vol. 53, Part 4, pp. 715 735.

Figure 11: Heat flux profiles for Rb/H = 0.1, 0.5 and 0.75 at Re=100
and T=10C.

CONCLUSIONS
This paper presents a twodimensional numerical model
that investigates the influence of steady thermal Marangoni
convection on the fluid dynamics and heat transfer around a
bubble during laminar flow of water in a rectangular
minichannel. This mixed convection problem is investigated for
channel Reynolds numbers in the range of 0 Re 500 and
Marangoni numbers in the range of 0 Ma 17114. The
thermocapillary effect has a significant impact on heat transfer
for this configuration with an average increase of 35% in the
heat flux figures at the downstream of the bubble while the
combination between thermocapillary and forced convection
mechanisms results in an average of 60% increase at the front
of the bubble. Further extension of the present work is expected
based on a three-dimensional model capable to quantify
precisely the impact of this flow on heat transfer enhancement.
ACKNOWLEDGEMENTS
We gratefully acknowledge the support from Science
Foundation Ireland that sponsored this research through grant
number ENMF 249. We acknowledge the help provided with
the CFD software by our colleagues Seamus OShaughnessy
and Geoff Bradley.

[12]C. Reynard, R. Santini and L. Tadrist, 2001.


Experimental study on the gravity influence on the
periodic thermocapillary convection around a bubble,
Experiments in Fluids, No. 31, pp. 440 446.
[13] P.H. Hadland, R. Balasubramaniam, G. Wozniak and R.S.
Subramanian, 1999. Thermocapillary migration of
bubbles and drops at moderate to large Marangoni number
and moderate Reynolds number in reduced gravity,
Experiments in Fluids, No. 26, pp. 240 248.
[14] A. Mukherjee, S.G. Kandlikar, 2006. Numerical study of
the effect of surface tension on vapour bubble growth
during flow boiling in microchannels, Proceedings of the
ICNMM2006, Fourth International Conference on
Nanochannels, Microchannels and Minichannels, Paper
No. ICNMM 2006 96050, Limerick, Ireland.
[15] P.S. Clokner and G.F. Naterer, 2007. Interfacial
thermocapillary pressure of an accelerated droplet in
microchannels: Part I. Fluid flow formulation,
International Journal of Heat and Mass Transfer, Volume
50, Issues 25-26, pp. 5269-5282.
[16] A. Bhunia and Y. Kamotani, 2001. Flow around a bubble
on a heated wall in a cross flowing liquid under

Copyright 2008 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 04/05/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

microgravity condition, International Journal of Heat and


Mass Transfer, no. 44, 2001, 3895 3905.
[17] J.S. Bintoro, A. Akbarzadeh, M. Mochizuki, 2005. A
closed loop electronics cooling by implementing single
phase impinging jet and mini channels heat exchanger,
Applied Thermal Engineering 25, pp. 2740.
[18] S. Narumanchi, A. Troshko, D. Bharathan, V. Hassani,
2007. Numerical simulations of nucleate boiling in
impinging jets: Applications in power electronics cooling,
International Journal of Heat and Mass Transfer 51, 112.
[19] D.S. Wen, Youyou Yan, D.B.R. Kenning, 2004.
Saturated flow boiling of water in a narrow channel: time
averaged heat transfer coefficients and correlations,
Applied Thermal Engineering 24, pp. 1207 1233.
[20] Fluent software, http://www.fluent.com

[21] Gambit, http://www.fluent.com/software/gambit/index.htm

[22] D.J. Tritton, 1998. Physical Fluid Dynamics, Oxford


Science Publication, Clarendon Press, ISBN 0-19-8544936, pp. 162-172.
[23] R.L. Panton, 1993. Incompressible flow, Wiley
Interscience Publication, John Wiley & Sons, ISBN 0-47189765, pp. 102-106.
[24] FLUENT User Guide, http://www.fluent.com

[25] P. Arlabosse, L. Tadrist, H. Tadrist and J. Pantaloni, 2000.


Experimental analysis of the heat transfer induced by
thermocapillaryconvection around a bubble, Transactions
of the ASME, Journal of Heat Transfer 122, 66 73.
[26] S. Petrovic, A.J. Robinson, L.R. Judd, 2004. Marangoni
heat transfer in subcooled nucleate pool boiling,
International Journal of Heat and Mass Transfer 47 (2004)
51155128.
[27] S. OShaughnessy and A.J. Robinson, 2007. Numerical
Investigation of Marangoni Convection Caused by the
Presence of a Bubble on a Uniformly Heated Surface,
Proceedings of ITP2007 Interdisciplinary Transport
Phenomena V: Fluid, Thermal, Biological, Materials and
Space Sciences, October 14-19, 2007, Bansko, Bulgaria.

Copyright 2008 by ASME

Downloaded From: http://proceedings.asmedigitalcollection.asme.org/ on 04/05/2016 Terms of Use: http://www.asme.org/about-asme/terms-of-use

Вам также может понравиться