Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

Journal of

Heat Transfer Technical Notes


This section contains shorter technical papers. These shorter papers will be subjected to the same review process as that for full papers.

Dissipation in Small Scale Gaseous are such that the flow and heat transfer characteristics are in the
slip-flow regime where the usual slip-flow boundary conditions
Flows

Nicolas G. Hadjiconstantinou
u gas兩 wall⫽ ␣
2⫺ ␴ v du
␴v

d ˜␩
冏 wall
(1)


Mechanical Engineering Department, Massachusetts
2 ␥ 2⫺ ␴ T ␭ dT
Institute of Technology, Cambridge, MA 02139 T gas兩 wall⫺T w ⫽ ␤ (2)
␥ ⫹1 ␴ T Pr d ˜␩ wall

are known to provide a good approximation to the velocity and


We discuss the effect of shear work at solid boundaries in small temperature fields, respectively. Here ␭ is the molecular mean free
scale gaseous flows where slip effects are present. The effect of path, ␴ v is the momentum accommodation coefficient, ˜␩ is the
shear work at the boundary on convective heat transfer is illus- coordinate normal to the wall, T w is the local wall temperature, ␴ T
trated through solution of the constant-wall-heat-flux problem in is the energy accommodation coefficient, Pr is the gas Prandtl
the slip-flow regime. We also present predictions for the dissipa- number, and ␥ is the ratio of specific heats. The coefficients ␣ and
tion in terms of the mean flow velocity in pressure-driven and ␤ introduce corrections to the original results of Maxwell ( ␣ ⫽ ␤
gravity-driven Poiseuille flows for arbitrary Knudsen numbers. ⫽1) that were obtained through an approximate method 关1兴. For
All results are verified using direct Monte Carlo solutions of the air, ␣ and ␤ are usually taken to be equal to unity 关2兴. Linearized
Boltzmann equation. 关DOI: 10.1115/1.1571088兴 solutions of the Boltzmann equation 关3,4兴 show that for hard
spheres ␣ ⬇ ␤ ⬇1.1. In what follows we will assume, without loss
Keywords: Heat Transfer, Microscale, Molecular Dynamics, of generality, that both accommodation coefficients are equal to
Monte Carlo, Nanoscale unity, or equivalently that their contribution has been absorbed in
␣ and ␤.
In this paper we discuss two aspects of dissipation in small Under the assumption of a long channel (LⰇH) and low speed
scale ideal-gas flows. We first discuss the effect of shear work at flow 共small pressure gradient or external field兲 we approximate
the boundary in slip flow and how this affects convective heat the flow as hydro-dynamically and thermally locally fully devel-
transfer in small scale channels. Shear work at the boundary has, oped; we also assume that the temperature changes and compress-
incorrectly, been neglected in recent studies of viscous heat dissi- ibility cause the flow to deviate negligibly from this state. Under
pation in slip-flow convective heat transfer. We show that this these assumptions we assume that locally the pressure gradient is
effect scales with the Brinkman number, and subsequently derive constant, u x ⫽u x (y), T⫽T(x,y) and ␶ xy ⫽ ␶ xy (y), where ␶ xy is
an expression for the fully developed slip-flow Nusselt number the xy component of the shear stress tensor. The velocity profile is
under constant-wall-heat-flux conditions in the presence of vis- then given by the slip-flow (Knⱗ0.1) Poiseuille profile

冋冉 冊 册
cous dissipation and compare this expression to direct Monte
Carlo solutions of the Boltzmann equation. The second aspect ub 1 y2
discussed is related to dissipation in transition-regime flows. Us-
u x⫽ ␣ Kn⫹ ⫺ 2 (3)
1 4 H
ing the solution of the Boltzmann equation for gravity and ␣ Kn⫹
6
pressure-driven flows in two-dimensional channels and energy
conservation arguments, we provide expressions for the energy where u b is the bulk 共average兲 velocity over the channel cross-
dissipation in these flows as a function of the flow speed for section and the Knudsen number is given by Kn⫽␭/H.
arbitrary Knudsen numbers. Our results are verified by direct The temperature equation in the presence of viscous heat gen-
Monte Carlo solutions of the Boltzmann equation. eration and flow work reduces to
For simplicity, we consider two-dimensional smooth channels
of length L in the axial 共x兲 direction, with perfectly accommodat- ⳵T dP ⳵qy ⳵ux
ing walls that are a distance H apart in the transverse 共y兲 direction. ␳ c pu x ⫺u x ⫽⫺ ⫹ ␶ xy (4)
The flow in the axial direction is sustained either by an imposed ⳵x dx ⳵y ⳵y
pressure gradient or external field g. The gas velocity field is
denoted uᠬ ⫽uᠬ (x,y)⫽(u x (x,y),u y (x,y),u z (x,y)), and T where c p is the specific heat at constant pressure, and q i is the
⫽T(x,y), P⫽ P(x,y) and ␳ ⫽ ␳ (x,y) denote the temperature, component of the heat flux vector in the i th direction. Our consti-
pressure and density fields respectively. tutive definitions follow those of 关5兴. Note that conservation laws
We first consider the case of pressure-driven flow that is heated will be given in their general form 共without continuum constitu-
or cooled by a constant wall-heat-flux. The channel dimensions tive models兲 applicable in all Knudsen regimes. Explicit use of the
continuum constitutive models will be limited to the derivation of
Contributed by the Heat Transfer Division for publication in the JOURNAL OF
the Nusselt number in the slip-flow regime. The integral form of
HEAT TRANSFER. Manuscript received by the Heat Transfer Division July 12, 2002; the above equation in the transverse direction under the assump-
revision received February 11, 2003. Associate Editor: P. S. Ayyaswamy. tion of negligible density variation in this direction is

944 Õ Vol. 125, OCTOBER 2003 Copyright © 2003 by ASME Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/11/2013 Terms of Use: http://asme.org/terms


␳ c pu bH
dT b
dx
⫺u b H
dP
dx
⫽2q o ⫹ 冕 H/2

⫺H/2
␶ xy
⳵ux
⳵y
dy (5)

where q o is the thermal energy transferred from the wall to the


fluid, and T b , the bulk temperature, is defined by
兰 ⫺H/2
H/2
u x Tdy
T b⫽ (6)
u bH
In addition to the thermal energy transfer, there is also dissipa-
tion due to shear work between the wall and the slipping gas. It is
the sum of these two contributions that is responsible for the axial
temperature gradient 共and perceived as a constant wall-heat-flux兲
as shown by the integral form of the total 共mechanical plus ther-
mal兲 energy equation
dT b
␳ c pu b H⫽ 关 u x ␶ xy ⫺q y 兴 ⫺H/2
H/2
(7)
dx
Fig. 1 Variation of the fully developed Nusselt number Nut
Equations 共5兲 and 共7兲 are linked by the mechanical energy balance with Brinkman number for KnÄ0.07. The solid line is the pre-
diction of Eq. „13…, and the stars denote DSMC simulations.
⳵P ⳵ ␶ xy ⳵ P ⳵ 共 u x ␶ xy 兲 ⳵ux
0⫽⫺u x ⫹u x ⫽⫺u x ⫹ ⫺ ␶ xy (8)
⳵x ⳵y ⳵x ⳵y ⳵y
which integrates to


more detailed discussion 共albeit for the continuum case兲 of the
H/2 ⳵ux dP competition between cooling and viscous heat dissipation, which
关 ␶ xy u x 兴 ⫺H/2
H/2
⫽ ␶ xy dy⫹u b H (9)
⫺H/2 ⳵y dx ultimately leads to Nu→⬁ for fairly large negative Br, can be
found in 关6兴.
and shows that the viscous heat and flow work terms in Eq. 共5兲 We performed DSMC simulations 关7兴 to verify Eq. 共13兲. Our
are, in effect, representing the contribution of the shear work at simulations were performed on a hard sphere system since for
the wall 共Eq. 共7兲兲. theory verification purposes it is preferable to perform simulations
We proceed by nondimensionalizing Eq. 共4兲 using ␪ ⫽(T on a molecular system whose properties are well characterized. A
⫺T w )/(q o H/ ␬ ) and Br⬘ ⫽ ␮ u 2b /(q o H), where ␬ is the thermal constant wall-heat-flux was achieved by applying a linearly vary-
conductivity of the gas, and ␩ ⫽y/H. When acceleration effects ing wall temperature. Our simulations represent the best compro-
are negligible, for constant wall-heat-flux under fully developed mise between high speeds for low relative statistical error and low
conditions, ⳵ T/ ⳵ x⫽dT b /dx⫽const. We thus substitute for ⳵ T/ ⳵ x speeds for negligible compressibility effects. Because the effect of
from Eq. 共5兲 and for this slip-flow calculation use the Fourier the wall shear stress is of the order of 10 percent for 兩Br兩ⱗ0.1, we
conduction law to obtain tried to minimize all possible sources of error. As a result, we used

冉 冉 冊冊
2 the fourth order approximations for the transport coefficients 共ap-
ux

冉 冉 冊冊
⳵ proximately 2 percent different from the typically-used first order
ux us 2
ub ⳵ 2␪ approximations 关8兴兲 and also used the values ␣ ⫽1.11 and ␤
2 1⫹6Br⬘ 1⫺ ⫺Br⬘ ⫽ (10)
ub ub ⳵␩ ⳵␩2 ⫽1.13 recommended by 关3,4兴. We additionally corrected our re-
sults for the effects of finite cell sizes and timestep 关9,10,11兴 in
where u s is the slip velocity, u x 兩 ⫾H/2 , given by our numerical solution. Our resulting error estimate including sta-
us 6 ␣ Kn tistical fluctuations is approximately 4 percent. The ratio of the
⫽ (11) expected average thermal creep velocity 关12兴 to the bulk velocity,
u b 1⫹6 ␣ Kn u c /u b , was less than 0.1.
We solve for ␪ using a symmetry condition at ␩ ⫽0 and the slip- The agreement between the DSMC simulations and the theoret-
flow relation 共2兲, and from this solution proceed to calculate T w ical results is very good 共see Fig. 1兲. This verifies the contribution
⫺T b . The fully developed slip-flow Nusselt number based on the of shear work at the boundary but also shows that the slip-flow
thermal energy exchange is then given by prediction is accurate to within 4 percent at Kn⫽0.07. Also, de-
spite the relatively large pressure gradients used ( P in / P out
q o 2H ⬃O(1.5), P out⬇1bar, L⬇20 ␮ m) the assumption of negligible
Nu⫽
␬ 共 T w ⫺T b 兲 compressibility and fully developed flow and temperature fields

冉 冊冉 冉 冊冊
seems to be reasonable. Given that slip-flow does not describe the
140 u s 2 54 30 u s 12 us 2
correct state of the gas close to the wall 共due to the presence of the
⫺2Br 1⫺ ⫺ ⫹ Knudsen layer兲 but rather provides a recipe for obtaining the hy-
17 ub 17 17 u b 51 ub

冉 冊
⫽ (12) drodynamic fields far away from the wall 关1兴, the importance of
6 us 2 us 2
140 ␥ Kn
1⫺ ⫹ ⫹ ␤ wall effects in this phenomenon makes the good agreement all the
17 u b 51 u b 17 ␥ ⫹1 Pr more remarkable.
where Br⫽NuBr⬘ /2⫽ ␮ u 2b /( ␬ (T w ⫺T b )). The fully developed In view of the interest in ever decreasing device sizes, it is
interesting to explore dissipation in the transition regime (0.1
slip-flow Nusselt number based on the total energy exchange be-
ⱗKnⱗ10). It is well known 关2兴 that Poiseuille profiles in this
tween wall and gas is then
regime become flat 共Eqs. 共3兲 and 共11兲 are not valid兲 and large

Nut ⫽Nu⫹
共 ␶ xy u x 兲 兩 H/22H
␬ 共 T w ⫺T b 兲
⫽Nu⫺12Br
us
ub
1⫺
us
ub 冉 冊 (13)
amounts of slip are observed at the walls leading to significant
dissipation there, which, as we saw above, does not affect the
temperature field inside the channel. In the interest of simplicity,
Although the above expression is exact, it is only expected to hold we will discuss steady, fully developed, Poiseuille-type flows with
for small Brinkman numbers since high velocities will violate the constant-temperature walls. These flows are the extension to arbi-
assumption of negligible compressibility and fluid acceleration. A trary Knudsen numbers of the Poiseuille flow assumed in the con-

Journal of Heat Transfer OCTOBER 2003, Vol. 125 Õ 945

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/11/2013 Terms of Use: http://asme.org/terms


vective heat transfer problem solved above and are thus interest-
ing to study in their own right as well as in connection with
convective heat transfer in the transition regime 关7兴. Note that no
linear constitutive closures for the shear stress tensor and heat flux
vector are now assumed.
For these flows, the total energy equation simplifies to
⳵ ⳵T
共 ␶ u ⫺q y 兲 ⫽ ␳ c p u x ⫽0 (14)
⳵ y xy x ⳵x
which shows that there is no net energy exchange with the bound-
ary, and thus the thermal energy flux at the wall is balanced by the
shear work at the boundary. Using Eq. 共9兲 or the thermal energy
equation we obtain

H/2
关 q y 兴 ⫺H/2 ⫽u b
dP
dx
H⫹ 冕 H/2

⫺H/2
␶ xy
⳵ux
⳵y
dy (15)

Note that the fully developed temperature profile will, in general,


be non-uniform in the y direction with an average temperature that Fig. 2 Nondimensional flowrate Q̄ as a function of the Knud-
is different from the wall temperature. The temperature profile can sen number. The solid line denotes the numerical solution of
be calculated if closures for the shear stress tensor and heat flux the Boltzmann Eq. †12‡, the stars denote DSMC simulations of
gravity-driven flow and the dashed line denotes the slip-flow
vector are provided. prediction. Error estimates are given by the star size.
The bulk velocity in Poiseuille flow for arbitrary Knudsen num-
bers is given by 关12兴

Hu b ⫽⫺
1 dP
P dx
冑 RT 2
2
H Q̄ (16) cantly less than 0.1兲 so we expect that the solutions of the Boltz-
mann equation for pressure-driven flow to be valid for gravity-
driven flow. Figure 2 shows that upon replacing ⫺d P/dx by ␳ g,
where R⫽k b /m is the gas constant, k b is Boltzmann’s constant, m
Q̄ as determined for pressure-driven flow accurately describes
is the molecular mass and Q̄⫽Q̄(Kn) is a proportionality coeffi-
gravity-driven flow.
cient that has been determined semianalytically by solution of the
Substituting from above, we obtain
Boltzmann Eq. 关1兴 and found to be in good agreement with mo-
lecular simulations and experimental data 关1,2兴. In the transition
regime, Q̄(Kn) varies slowly about its minimum value (1.5 ␳ u 2b 冑2RT
ⱗQ̄(0.1⬍Kn⬍10)ⱗ3) occuring at Kn⬇1. q t⫽ (21)
Since the flow is steady and fully developed (d P/dx)H Q̄
⫽ 关 ␶ xy 兴 ⫺H/2
H/2
. Thus,
The energy dissipated inside the channel is
␳ u b 冑2RT
⫽ 关 ␶ xy u x 兴 ⫺H/2 ⫽⫺ ␳ u b 冑2RT
H/2 H/2
关 q y 兴 ⫺H/2 us (17)
Q̄ q t ⫹ 关 ␶ xy u x 兴 ⫺H/2
H/2
⫽ 共 u b ⫺u s 兲 (22)

and

冕 H/2

⫺H/2
␶ xy
⳵ux
⳵y
dy⫽
␳ u b 冑2RT

共 u b ⫺u s 兲 (18)
We performed direct Monte Carlo simulations to verify these
predictions. We show here results from the gravity-driven flow for

The above equation suggests that as the Knudsen number in-


creases and the velocity profile becomes flatter, most of the energy
is dissipated at the walls, as expected.
For a steady and fully developed gravity-driven flow, we have

⳵ ⳵qy
共 ␶ u 兲 ⫹ ␳ gu x ⫽ (19)
⳵ y xy x ⳵y

The total energy transfer to the wall, q t , is equal to the work done
by the gravity force

q t ⫽ 关 q y ⫺ ␶ xy u x 兴 ⫺H/2
H/2
⫽ ␳ gu b H (20)

Here we use the fact that the bulk velocity in gravity-driven


flow can be determined by utilizing the similarity between gravity
and pressure-driven flows. Malek Mansour et al. 关13兴 have shown
that although from a Boltzmann equation perspective gravity and Fig. 3 Nondimensional total heat exchange q t Õ„ ␳ u 2b 冑2 RT … as
pressure-driven flows are different, the difference in the solutions a function of the Knudsen number. The solid line denotes the
scales with the square of the characteristic gravity parameter ⑀ theoretical prediction Eq. „21…, the stars denote DSMC simula-
⬅gmH/(2k b T) which is expected to be small in practical appli- tions of gravity-driven flow and the dashed line denotes the
cations. This parameter in our simulations was very small 共signifi- slip-flow prediction. Error estimates are given by the star size.

946 Õ Vol. 125, OCTOBER 2003 Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/11/2013 Terms of Use: http://asme.org/terms


which a net energy exchange between wall and gas occurs and can J. L. Zakin
thus be measured. Figure 3 shows that Eq. 共21兲 is in good agree- Ohio State University, Department of Chemical
ment with simulation results.
Engineering, Columbus, OH 43210, U.S.A.

Acknowledgments
The thermal pattern on a heated wall was studied for the flow of
The author wishes to thank Pirouz Kavehpour, Lowell Baker water and drag-reducing surfactant solutions in a channel. The
and Professor John Lienhard V for helpful comments and discus- wall of the channel was made of a thin foil, which was heated by
sions. This work was supported by the Singapore-MIT alliance. direct current. The temperature of the foil, which reflects the local
flow velocities, was measured by an infrared technique with high
spatial and temperature resolution. The microstructure of the sur-
References factant solution was studied by direct imaging cryogenic tempera-
关1兴 Cercignani, C., 1988, The Boltzmann Equation and Its Applications, Springer- ture transmission electron microscopy (Cryo-TEM). The most
Verlag, NY. prevalent structures observed are thread-like micelles, which have
关2兴 Beskok, A., and Karniadakis, G. E., 1999, ‘‘A Model for Flows in Channels been suggested to cause the modification of the thermal
and Ducts at Micro and Nano Scales,’’ Microscale Thermophys. Eng., 3, p. 43.
关3兴 Ohwada, T., Sone, Y., and Aoki, K., 1989, ‘‘Numerical Analysis of the Shear patterns. 关DOI: 10.1115/1.1609482兴
and Thermal Creep Flows of a Rarefied Gas Over a Plane Wall on the Basis of
the Linearized Boltzmann Equation for Hard-Sphere Molecules,’’ Phys. Fluids Keywords: Heat Transfer, Microstructures, Non-Newtonian, Poly-
A, 1, p. 1588.
关4兴 Sone, Y., Ohwada, T., and Aoki, K., 1989, ‘‘Temperature Jump and Knudsen mers, Rheological, Thermal, Turbulence, Visualization, Vortex
Layer in a Rarefied Gas Over a Plane Wall: Numerical Analysis of the Linear-
ized Boltzmann Equation for Hard-sphere Molecules,’’ Phys. Fluids A, 1, p.
363.
关5兴 Vincenti, W. G., and Kruger, C. H., 1965, Introduction to Physical Gas Dy- 1 Introduction
namics, Krieger, FL.
关6兴 Ou, J. W., and Cheng, K. C., 1973, ‘‘Viscous Dissipation Effects on Thermal Reduction of friction losses in turbulent flows caused by the
Entrance Region Heat Transfer in Pipes With Uniform Wall Heat Flux,’’ Appl. presence of polymer additives was first reported about fifty years
Sci. Res., 28, p. 289. ago. Virk 关1兴 examined friction factor and velocity profile results
关7兴 Hadjiconstantinou, N. G., and Simek, O., 2002, ‘‘Constant-Wall-Temperature
Nusselt Number in Micro and Nano Channels,’’ J. Heat Transfer, 124,
for a large number of high polymer solutions, mostly but not all,
p. 356. in water. He noted that at relatively low concentrations many so-
关8兴 Chapman, S., and Cowling, T. G., 1970, The Mathematical Theory of Non- lutions reached lower limiting values in their friction factor/
Uniform Gases, Cambridge University Press, Cambridge. Reynolds number data. Virk 关1兴 proposed an equation for the lim-
关9兴 Alexander, F., Garcia, A., and Alder, B., 1998, ‘‘Cell Size Dependence of
Transport Coefficients in Stochastic Particle Algorithms,’’ Phys. Fluids, 10, p.
iting maximum drag reduction asymptote 共MDRA兲. Polymer
1540, 共Erratum: Phys. Fluids, 12, p. 731共2000兲兲. effectiveness depends on the presence of high molecular weight
关10兴 Hadjiconstantinou, N. G., 2000, ‘‘Analysis of Discretization in the Direct species, which limits their applications because of the susceptibil-
Simulation Monte Carlo,’’ Phys. Fluids, 12, p. 2634. ity of a high molecular weight components to degradation in high
关11兴 Garcia, A., and Wagner, W., 2000, ‘‘Time Step Truncation Error in Direct
Simulation Monte Carlo,’’ Phys. Fluids, 12, p. 2621.
shear flows.
关12兴 Ohwada, T., Sone, Y., and Aoki, K., 1989, ‘‘Numerical Analysis of the Poi- In the past decade considerable interest has developed in ‘‘re-
seuille and Thermal Transpiration Flows Between Parallel Plates on the Basis pairable’’ drag reducing additives, such as surfactants for use in
of the Boltzmann Equation for Hard-Sphere Molecules,’’ Phys. Fluids A, 1, technological processes. Though surfactant drag reducing addi-
p. 2042.
关13兴 Malek Mansour, M., Baras, F., and Garcia, A. L., 1997, ‘‘On the Validity of
tives require higher concentrations than high polymers, their long
Hydrodynamics in Plane Poiseuille Flows,’’ Physica A, 240, p. 255. life and greater percent of drag reduction make them very attrac-
tive for recirculation flows. Chara et al. 关2兴 showed that the
MDRA proposed by Virk 关1兴 is not valid for the surfactant
Habon G. 共hexadecyldimethyl hydroxyethyl ammonium-2-
hydroxy-3-naphtoate兲 solutions.
The Effect of a Cationic Surfactant on Investigation of the thermal development region for surfactant
Turbulent Flow Patterns solutions was conducted by Gasljevic and Matthys 关3兴. Turbulent
transport mechanism in a drag reducing flow with surfactant ad-
ditive was investigated by Kawaguchi et al. 关4兴. They found that
G. Hetsroni only large eddies are dominant near the wall. Warholic et al. 关5兴
e-mail: hetsroni@tx.technion.ac.il presented measurements of turbulence properties of a solution of
tris-hydroxyethyl-ammonium acetate 共Ethoquad T13-50兲 and so-
Fellow ASME dium salicylate 共NaSal兲 flowing in a 5.08 cm⫻61 cm rectangular
channel. The profiles of average velocity obtained by Warholic
A. Mosyak et al. 关5兴 differ from those presented by Chara et al. 关2兴 for sur-
factant solutions. If turbulence production depends on the proper-
Department of Mechanical Engineering, Technion-Israel ties of the surfactant solution, one should expect different systems
Institute of Technology, Haifa 32000, Israel to produce different velocity and temperature fields. Thus, it might
be useful to connect the behavior of high drag-reducing surfac-
tants with their microstructure.
One of the most outstanding characteristics of the near wall
Y. Talmon turbulence structure in a boundary layer of a channel flow is the
Department of Chemical Engineering, Technion-Israel presence of coherent structures. Coherent structures are respon-
Institute of Technology, Haifa 32000, Israel sible for most of the turbulence production, dissipation and trans-
port phenomena. Donohue et al. 关6兴, and Achia and Thompson 关7兴
considering the near-wall coherent structures in so-called low
A. Bernheim-Groswasser
Contributed by the Heat Transfer Division for publication in the JOURNAL OF
Department of Chemical Engineering, HEAT TRANSFER. Manuscript received by the Heat Transfer Division August 30,
Ben Gurion University, Beer-Sheba 84105, Israel 2002; revision received June 18, 2003. Associate Editor: G. P. Peterson.

Journal of Heat Transfer Copyright © 2003 by ASME OCTOBER 2003, Vol. 125 Õ 947

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/11/2013 Terms of Use: http://asme.org/terms


drag-reducing solutions, in which the drag reduction did not ex-
ceed 30%. Hetsroni et al. 关8兴 studied this problem in high drag-
reducing solution 共drag reduction varied in the range of 46 – 85%兲.
It was shown that the dimensionless streak spacing depends on the
Reynolds number, based on wall-shear velocity of the solution.
On the other hand, the relation of steak spacing to the onset of
turbulent drag reduction was not clarified. The first objective of
the present study was to check whether or not the onset of turbu-
lence drag reduction might be connected to spacing of near-wall
turbulence structures. It was achieved by comparison of the ex-
perimental results obtained in the present study with data reported
in 关6,7兴.
Considerable research was also carried out to investigate micro-
structures of drag-reducing cationic surfactant systems 关9兴. The
second aim of the present study was to gain more insight into the
near-wall turbulence structure in high drag-reducing flow by ana-
lyzing solution micro-structure. Fig. 2 Test section „1-top of the channel, 2-stainless steel
strip, 3-window, 4-bottom of the channel, 5-IR camera…

2 Experiment
temperature field was fully developed, i.e., the temperature distri-
2.1 Experimental Facility. The experiments were con- bution on the heated wall did not change in the streamwise direc-
ducted in a rectangular channel. The two-dimensional channel tion. The test section is shown in Fig. 2.
flow offers several advantages for studies of near-wall coherent
structures, as flow visualization is then relatively easy. 2.2 Measurement Techniques. To measure the temperature
The channel flow system is shown in Fig. 1. The 7.2 m long, field on the heated wall a Thermal Imaging Radiometer was used
0.2 m wide, and 0.02 m deep rectangular channel comprised, with a typical horizontal and vertical resolution of 256 pixels per
twelve Plexiglas sections of length 0.6 m each. The sections were line, spatial resolution 0.05 mm, response time 0.04 s, sensitivity
carefully joined to ensure a hydraulic smooth surface throughout. 0.1 K. Since the heating strip was very thin 共0.05 mm兲, the tem-
The temperature measurements were carried out in the test sec- perature difference between its surfaces did not exceed 0.1 K, 关8兴.
tion. The heating strips 0.6⫻0.2 m each were installed inside the A computer program made it possible to store the information and
channel from the front end of the development section to a dis- to compute the statistics of the thermal field.
tance of 0.6 m beyond the test section. These strips were made of The water temperature was measured by a precision mercury
0.05 mm thick stainless steel and arranged so that the boundary thermometer with an accuracy 0.1 K. The mean flow velocity was
layer could be heated along different distances from the inlet to measured with an accuracy ⫾1%, the electric power was deter-
the test section. The latter was provided with two 0.2⫻0.16 m mined with an accuracy ⫾0.5%. The pressure drop was measured
windows to which the strips were bonded with contact adhesive by pressure transducers with an accuracy ⫾1.5%. The shear ve-
and coated on the air side with black mat paint about 0.02 mm locity was calculated with an accuracy ⫾4%.
thick. DC current up to 300A was applied to the heating strips,
and measurements were taken at different lengths of the heated 2.3 Properties of the Surfactant Solution. The surfactant
stretch. It was found that at the location of the test section, the used was Habon G—a cationic surfactant of molecular weight
500. The content of delivered active material is 53.5% active sur-
factant, 10.2% isopropanol and 36.3% water. Concentrations re-
ported are the concentrations of active surfactant. The head group
of the surfactant is hexadecyldimethyl hydroxyethyl ammonium
and the counter-ion is 3-hydroxy-2-naphthoate. The surfactant
molecules form large thread-like micelles which are effective in
causing drag reduction. In the present study we used an active
surfactant concentration of 530 ppm 共parts per million by weight
in filtered, deionized water兲. Our data, based upon pressure drop
and heat transfer measurements, did not show degradation of the
solution during the experiments.
The kinematic viscosity was determined by a Cannon-Fenske
capillary viscometer. The kinematic viscosity is ␯ ⫽1.35
•10⫺6 m2 s⫺1 at t⫽20°C and 1.15•10⫺6 m2 s⫺1 at t⫽40°C. In
the experiments performed, the temperature of the surfactant so-
lution was 22⫾0.5°C, and the heated wall temperature ranged
from 32 to 40°C. Thus, the Reynolds numbers for the 530 ppm
Habon G solution were calculated based on a kinematic viscosity
␯ ⫽1.25•10⫺6 m2 s⫺1 .
2.4 Microstructure Information. We examined the Habon
G solution by direct imaging cryogenic temperature transmission
electron microscopy 共Cryo-TEM兲 关9兴. Transmission electron mi-
croscopy is a direct imaging technique that affords microstructural
information that does not require modeling for interpretation 共con-
Fig. 1 Loop of rectangular channel „1-tank, 2-pump, 3-control trary to indirect methods such as light and x-ray scattering兲. Cryo-
valve, 4-flow meter, 5-strainghtener, 6-development section, TEM does not involve any staining or drying of the studied
7-test section, 8-IR camera, 9-outlet section, 10-heat ex- samples, thus it provides reliable images of the original micro-
changer… structures in the studied systems.

948 Õ Vol. 125, OCTOBER 2003 Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/11/2013 Terms of Use: http://asme.org/terms


Fig. 4 Thermal pattern on the heated wall: „a… Flow of 530 ppm
Habon G solution; and „b… Flow of water.

Fig. 3 Microstructure of the 530 ppm Habon G solution: „a… 3 Results


threadlike micelles; and „b… vesicles.
3.1 Cryo-TEM Images. The most prevalent structures we
observed are thread-like micelles, shown at high magnification in
Fig. 3共a兲. Micrographs like this one show quite clearly inner struc-
tural details of the micelles such as branching domains 共arrow兲
and overlap of micelles 共arrowhead兲. It should be noted that
We prepared vitrified specimens for Cryo-TEM in a controlled threadlike micelles have been suggested as the one, possibly the
environment vitrification chamber 共CEVS兲 to ensure preservation most important, microstructural feature that modifies flow patterns
of the original concentration, temperature, and thus the micro- in a flowing fluid and reduces drag 关10兴.
structure of the examined sample. As with all other types of ma- Another feature seen in the examined solutions are vesicles,
terial systems, TEM specimens of liquid systems have to be quite denoted by ‘‘V’’ in Fig. 3共b兲. These are balloon-like structures
thin. The penetration power of even high energy electrons is rather made of a double-layer membrane. Such structures are found in
limited. To avoid inelastic electron scattering that leads to image many biological and synthetic amphiphiles 关11兴. In this micro-
deterioration, and to take full advantage of phase-contrast in direct graph we see vesicles coexisting with threadlike micelles 共ar-
imaging, one needs to limit specimen thickness to about 0.2 ␮m. rows兲. Much to our surprise we have detected junctions between
The surfactant solution specimens were prepared for Cryo- vesicles and threadlike micelles. In fact, three such junctions, one
TEM imaging by applying a small drop of the studied solution denoted by an arrowhead are seen in the lower left part of the field
onto a perforated carbon film supported on an electron microscope of view of Fig. 3共b兲. The three micelles connecting the three
grid, blotting it to form a thin 共0.2 ␮m兲 film. By avoiding crystal- vesicles are connected by a three-fold junction seen just above the
lization of water the microstructure is not disturbed and the im- arrowhead. Zheng et al. 关12兴 hypothesize that the straining actions
ages obtained reflect true microstructures in the original solutions. of flow disrupt vesicles and thus induce structural instability of the
The solutions were quenched from 25°C, and 100% relative hu- fragments that leads to their reconstruction into networks of
midity. Specimens were examined in a Philips CM120 micro- branching threadlike micelles.
scope, operated at 120 kV, using an Oxford CT-3500 cryo-holder 3.2 Thermal Streaks at the Wall. One important aspect of
maintained below ⫺178°C. Micrographs were collected digitally turbulent flow is believed to be streamwise vortex structure with
关10兴 by a Gatan 791 MultiScan CCD camera with the DigitalMi- its accompanying low-speed streaks. The temperature distribution
crograph software package, using low-dose protocols to minimize on the heated wall can be considered as a trace of the flow struc-
electron-beam radiation-damage. ture there, i.e., the turbulent structures in the boundary layer cause

Journal of Heat Transfer OCTOBER 2003, Vol. 125 Õ 949

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/11/2013 Terms of Use: http://asme.org/terms


mensionless streak spacing, at given value of wall-shear, depends
on onset of wall-shear, velocity of the solution. The results are
consistent with the hypothesis that one of the prerequisites for the
phenomenon of drag reduction is sufficiently enhanced length and
time scale of the velocity field. Different drag-reducing solutions
produce velocity field with different length and time scales.
The existence of an extremely low value of onset velocity is
associated with microstructure of the solution. From the study of
microstructure one can conclude that the surfactant used in the
present study contains threadlike micelles with junctions between
them. These data provide the first experimental demonstration that
well developed network, including branched micelles, is con-
nected with streak formation.

Acknowledgments
This research was supported by the Fund for the Promotion of
Research at the Technion. A. Mosyak is supported by The Minis-
try of Absorption of Israel. This research was also supported by
Fig. 5 Relationship between dimensionless streak spacing the Israeli Ministry of Science.
and wall shear velocity „䊉-present study; 䊏-Donohue et al.
„1972…; and 䉲-Achia and Thompson „1977……
Nomenclature
Dh ⫽ channel hydraulic diameter
the temperature distribution on the wall, including the thermal L ⫽ channel length
streaks. The streak spacing results were obtained by IR visualiza- u* ⫽ ( ␶ / ␳ ) 0.5, shear velocity
tion and from spatial correlation of the temperature. The images of *
u on ⫽ onset shear velocity
thermal streak structures are shown in Figs. 4共a兲 and 4共b兲. Figure ␭ ⫽ thermal streaks spacing
4共a兲 shows a typical example of frames reproduced from a video ␭⫹ ⫽ ␭u * / ␯ , dimensionless thermal streaks spacing
motion picture of surfactant drag reducing flow. Figure 4共b兲 is a ␳ ⫽ density
representative image of the thermal streak structure, at the same ␷ ⫽ kinematic viscosity
shear velocity, for water flow. These pictures are plan view, the ⌬p ⫽ pressure drop
flow moves from the bottom to the top of the figures. The thermal ␶ ⫽ ⌬pD h /4L, shear stress
streak spacing was calculated by two-point correlation’s 关8兴. The
dimensionless thermal streak spacing ␭ ⫹ ⫽␭u * / ␯ 共␭ is the streak
spacing, u * shear velocity, calculated from the pressure drop, References
u * ⫽ 冑␶ / ␳ , ␶ shear stress, ␳ density, ␯ kinematic viscosity兲 ob- 关1兴 Virk, P. S., 1975, ‘‘Drag Reduction Fundamentals,’’ AIChE J., 21, pp. 625–
656.
tained from the temperature field on the heated wall in water flow 关2兴 Chara, Z., Zakin, J. L., Severa, M., and Myska, J., 1993, ‘‘Turbulence Mea-
was ␭ ⫹ ⫽100⫾10. The result agrees well with data reported by surements of Drag Reducing Surfactant Systems,’’ Exp. Fluids, 16, pp. 36 – 41.
Iritani et al. 关13兴. In the drag reducing flows ␭ ⫹ increase when the 关3兴 Gasljevic, K., and Matthys, E. F., 1997, ‘‘Experimental Investigation of Ther-
mal and Hydrodynamic Development Regions for Drag Reducing Surfactant
shear velocity increases. Solutions,’’ J. Heat Transfer , 119, pp. 80– 88.
Figure 5 represents streak spacing measurements for different 关4兴 Kawaguchi, Y., Tawaraya, Y., Yabe, A., Hishida, K., and Maeda, M., 1996,
drag reducing solutions. The data of Donohue et al. 关6兴 and Achia ‘‘Turbulent Transport Mechanism in Drag Reducing Flow With Surfactant Ad-
and Thompson 关7兴 for polymer solutions are also shown for com- ditive Investigated by Two Component LDV,’’ in Eighth International Sympo-
sium on Applications of Laser Techniques to Fluid Mechanics, 2, July 8 –11,
parison. Points E, F, G, at which the lines ␭ ⫹ ⫽ f (u * ) cross the Lisbon, Portugal, pp. 29.4.1–29.4.7.
line ␭ ⫹ ⫽100 indicate a threshold value of wall shear for the 关5兴 Warholic, M. D., Schmidt, G. M., and Hanratty, T. J., 1999, ‘‘The Influence of
particular drag reducing solution. The increase in the value of ␭ ⫹ a Drag-Reducing Surfactant on a Turbulent Velocity Field,’’ J. Fluid Mech.,
388, pp. 1–20.
associated with drag reduction takes place at values of u * higher 关6兴 Donohue, G. L., Tiederman, W. G., and Reischman, W. G., 1972, ‘‘Flow Vi-
than the onset wall shear. Thus, the drag reducing effect starts sualization of the Near-Wall Region in Drag-Reducing Flow,’’ J. Fluid Mech.,
with wall shear stresses larger than a threshold value, which de- 56, pp. 559–575.
关7兴 Achia, B. U., and Thompson, D. W., 1977, ‘‘Structure of the Turbulent Bound-
pends on the nature of the additive and its concentration. When ary in Drag-Reducing Pipe Flow,’’ J. Fluid Mech., 81, pp. 439– 464.
surfactant flow is compared with flows of polymer solution, one 关8兴 Hetsroni, G., Zakin, J. L., and Mosyak, A., 1997, ‘‘Low-Speed Streaks in
can see that the 530 ppm Habon G solution has a significantly Drag-Reduced Turbulent Flow,’’ Phys. Fluids, 9, pp. 2397–2404.
lower value of onset shear velocity. The correlation between ␭ ⫹ 关9兴 Talmon, Y., 1999, ‘‘Cryogenic Temperature Transmission Electron Microscopy
in the Study of Surfactant Systems,’’ Modern Characterization Methods of
and the percentage of drag reduction was given by Oldaker and Surfactants Systems, B. P. Binks, Editor, Marcel Dekker, NY, pp 147–178,
Tiederman 关14兴: ␭ ⫹ ⫽1.9DR⫹99.7. The fact that we refer here to Chap. 6.
parallel shift of the ␭ ⫹ profile, means that at the same value of 关10兴 Lu, B., Zheng, Y., Scriven, L. E., Davis, H. T., Talmon, Y., and Zakin, J. L.,
1998, ‘‘Effect of Variations Counterion-to-Surfactant Ratio on Rheology and
friction velocity, the percentage of drag reduction increases with Microstructures of Drag Reducing Cationic Surfactant Systems,’’ Rheol. Acta,
the decrease of the onset velocity. Correspondingly we consider 37, pp. 528 –548.
the value of ␯ /u on* as the length scale and the value of ␯ /u * 2
on as
关11兴 Evans, F. E., and Wennerström, H., 1999, The Colloidal Domain, 2nd ed.,
the time scale. Different drag-reducing solutions produce velocity VCH, New York.
关12兴 Zheng, Y., Lin, Z., Zakin, J. L., Talmon, Y., Davis, H. T., and Scriven, L. E.,
field with different length and time scale. 2000, ‘‘Cryo-TEM Imaging the Flow-Induced Transition From Vesikles to
Threadlike Micelles,’’ J. Phys. Chem. B, 104共22兲, pp. 5263–5271.
关13兴 Iritani, Y., Kasagi, N., and Hirata, N., 1983, ‘‘Heat Transfer Mechanism and
Conclusion Associated Turbulence Structure in the Near Wall Region of a Turbulent
The shear velocity for onset of drag reduction for Habon G Boundary Layer,’’ in Fourth Symposium of Turbulent Shear Flows, 12–14
September 1983, Karslruhe, F.R. Germany.
surfactant solutions is significantly lower than those for polymer 关14兴 Oldaker, O. K., and Tiederman, W. G., 1977, ‘‘Spatial Structure of the Viscous
solutions. Changes in the shape of thermal streaks and an increase Sublayer in Drag-Reducing Channel Flow,’’ Phys. Fluids, 20共10兲, pp. 133–
in streak spacing are the main features in drag reduced flow. Di- 144.

950 Õ Vol. 125, OCTOBER 2003 Transactions of the ASME

Downloaded From: http://heattransfer.asmedigitalcollection.asme.org/ on 11/11/2013 Terms of Use: http://asme.org/terms

You might also like