You are on page 1of 12

PHYSICS OF FLUIDS 23, 082003 (2011)

An integral perturbation model of flow and momentum transport in rotating


microchannels with smooth or microstructured wall surfaces
Vince D. Romanina) and Van P. Careyb)
Department of Mechanical Engineering, University of California, Berkeley 94720-1740, USA

(Received 18 October 2010; accepted 7 July 2011; published online 17 August 2011)
This paper summarizes the development of an integral perturbation solution of the equations
governing flow momentum transport and energy conversion in microchannels between disks of
multiple-disk drag turbines such as Tesla turbines. Analysis of this type of flow problem is a key
element in optimal design of Tesla drag-type turbines for geothermal or solar alternative energy
technologies. In multiple-disk turbines, high speed flow enters tangentially at the outer radius of
cylindrical microchannels formed by closely spaced parallel disks, spiraling through the channel to
an exhaust at a small radius, or at the center of the disk. Previous investigations have generally
developed models based on simplifying idealizations of the flow in these circumstances. Here,
beginning with the momentum and continuity equations for incompressible and steady flow in
cylindrical coordinates, an integral solution scheme is developed that leads to a dimensionless
perturbation series solution that retains the full complement of momentum and viscous effects to
consistent levels of approximation in the series solution. This more rigorous approach indicates all
dimensionless parameters that affect flow and transport and allows a direct assessment of the relative
importance of viscous, pressure, and momentum effects in different directions in the flow. The
resulting lowest-order equations are solved explicitly and higher order terms in the series solutions
are determined numerically. Enhancement of rotor drag in this type of turbine enhances energy
conversion efficiency. We also developed a modified version of the integral perturbation analysis that
incorporates the effects of enhanced drag due to surface microstructuring. Results of the model
analysis for smooth disk walls are shown to agree well with experimental performance data for a
prototype Tesla turbine and predictions of performance models developed in earlier investigations.
Model predictions indicate that enhancement of disk drag by strategic microstructuring of the disk
surfaces can significantly increase turbine efficiency. Exploratory calculations with the model
indicate that turbine efficiencies exceeding 75% can be achieved by designing for optimal ranges of
C 2011 American Institute of Physics.
the governing dimensionless parameters. V
[doi:10.1063/1.3624599]

I. INTRODUCTION

Because multiple disk drag-type turbines, like the Tesla


turbine,1 are generally simple to manufacture and robust,
they are now being reconsidered as an expander option for
renewable energy applications such as solar Rankine combined heat and power systems, and geothermal power systems. To achieve optimized designs for applications of this
type, models of the flow, momentum transport, and energy
conversion in such devices are needed that are accurate
and that illuminate the parametric trends in expander
performance.
Multiple-disk Tesla-type drag turbines rely on a mechanism of energy transfer that is fundamentally different from
most typical airfoil-bladed turbines or positive-displacement
expanders. A complete understanding of turbine operation
requires an analytical treatment of the unique fluid mechanics processes that effect energy transfer from the fluid to the
rotor. A schematic of the Tesla turbine can be found in
Figure 1. The turbine rotor consists of several flat, parallel
a)

Electronic mail: vince.romanin@berkeley.edu.


Electronic mail: vcarey@me.berkeley.edu.

b)

1070-6631/2011/23(8)/082003/11/$30.00

disks mounted on a shaft with a small gap between each


disk; these gaps form the cylindrical microchannels through
which flow will be analyzed. Exhaust holes on each disk are
placed as close to the center shaft as possible. Flow from the
nozzle enters the cylindrical microchannels at an outer radius
ro where ro  DH (DH is the hydraulic diameter of the
microchannels). The flow enters the channels at a high speed
and a direction nearly tangential to the outer circumference
of the disks and exits through an exhaust port at a much
smaller inner radius ri. Energy is transferred from the fluid to
the rotor via the shear force at the microchannel walls. As
the spiraling fluid loses energy, the angular momentum drops
causing the fluid to drop in radius until it reaches the exhaust
port at ri. This process is shown in Figure 2.
Several authors have studied Tesla turbines in order to
gain insight into their operation. In the 1960s, Rice2 and
Breiter and Pohlhausen3 conducted extensive analysis and
testing of Tesla turbines. However, Rice did not directly compare experimental data to analytical results, and lacked an analytical treatment of the friction factor. Breiter et al. provided
a preliminary analysis of pumps only and used a numerical solution of the energy and momentum equations. Hoya4 and
Guha5 extensively tested sub-sonic and super-sonic nozzles

23, 082003-1

C 2011 American Institute of Physics


V

082003-2

V. D. Romanin and V. P. Carey

FIG. 1. Schematic of a Tesla turbine.

with Tesla turbines, however, their analysis was focused on


experimental results and not an analytical treatment of the
fluid mechanics that drive turbine performance. Carey6 proposed an analytical treatment that allowed for a closed-form
solution of the fluid mechanics equations in the flow in the
rotor; however, Careys model analysis invoked several idealizations that neglected viscous transport in the radial and axial
directions, and treated tangential viscous effects using a friction factor approach.
While previous investigations of momentum transport in
Tesla turbines described above have generally developed
models based on idealizations of the flow in these circumstances, here, an integral perturbation analysis framework
was explored as a means of providing a more rigorous fluid
mechanical treatment of the flow that also quantifies the
effects of relevant dimensionless parameters on performance. Beginning with the momentum and continuity equations for incompressible and steady flow in cylindrical
coordinates, an integral solution scheme was developed that
leads to a dimensionless perturbation series solution that
retains the full complement of momentum and viscous
effects to consistent levels of approximation in the series solution. This more rigorous approach directly indicates the
dimensionless parameters that affect flow and transport and
allows a direct assessment of the relative importance of pressure, viscous, and momentum effects in different directions
in the flow.
The performance analysis in the previous investigations described above suggest that enhancement of rotor
drag in this type of turbine generally enhances energy conversion efficiency. Information obtained in recent fundamental studies indicates that laminar flow drag can be
strongly enhanced by strategic microstructuring of the wall

FIG. 2. (Color online) Schematic of flow through a Tesla turbine


microchannel.

Phys. Fluids 23, 082003 (2011)

surfaces in microchannels.79 The conventional Moody diagram shows that for most channels, surface roughness has
no effect on the friction factor for laminar flow in a duct.
However, in micro-scale channels several physical nearsurface effects can begin to become significant compared to
the forces in the bulk flow. First, the Moody diagram only
considers surface roughnesses up to 0.05, which is small
enough not to have meaningful flow constriction effects. In
microchannels, manufacturing techniques may often lead to
surface roughnesses that comprise a larger fraction of the
flow diameter. When the reduced flow area becomes small
enough to affect flow velocity, the corresponding increase
in wall sheer can become significant. Secondly, the size,
shape, and frequency of surface roughness features can
cause small areas of recirculation, downstream wakes, and
other effects which may also impact the wall shear in ways
that become increasingly important in smaller size channels, as the energy of the perturbations become relevant
compared to the energy of the bulk flow.
In 2005, Kandlikar et al.7 modified the traditional
Moody diagram to account for surfaces with a relative
roughness higher than 0.05, arguing that above this value
flow constriction becomes important. Kandlikar proposes
that the constricted diameter be simplified to be
Dcf Dt  2e, where e is the roughness height, Dt is the base
diameter, and Dcf is the constricted diameter. Using this formulation, the Moody diagram can be re-constructed to
account for the constricted diameter, and can thus be used
for channels with relative roughness larger than 5%. Kandlikar conducts experiments which match closely with this prediction and significantly closer than the prediction of the
classical Moody diagram. Kandlikar, however, only conducts
experiments on one type of roughness element, and does not
analyze the effect of the size, shape, and distribution of
roughness elements, although he does propose a new set of
parameters that could be used to further characterize the
roughness patterns in microchannels.
Croce et al.8 used a computational approach to model
conical roughness elements and their effect on flow through
microchannels. Like Kandlikar, he also reports a shift in the
friction factor due to surface roughness and compares the
results of his computational analysis to the equations proposed by several authors for the constricted hydraulic diameter for two different roughness element periodicities. While
the results of his analysis match Kandlikars equation
(Dcf Dt  2e) within 2% for one case, for a higher periodicity Kandlikars approximation deviates from numerical
results by 10%. This example, and others discussed in Croces paper, begins to outline how roughness properties other
than height can effect a shift in the flow Poiseuille number.
Gamrat et al.9 provides a detailed summary of previous
studies reporting Poiseuille number increases with surface
roughness. He then develops a semi-empirical model using
both experimental data and numerical results to predict the
influence of surface roughness on the Poiseuille number.
Gamrats analysis, to the best of the authors knowledge, is
the most thorough attempt to predict the effects of surface
roughness on the Poiseuille number of laminar flow in
microchannels.

082003-3

An integral perturbation model of flow

Phys. Fluids 23, 082003 (2011)

There appears to have been no prior efforts to model


and quantitatively predict the impact of this type of drag
enhancement on turbine performance. The integral perturbation analysis can be modified to incorporate the effects of
enhanced drag due to surface microstructuring. The goal of
this analysis is to model surface roughness effects on momentum transport in drag-type turbines in the most general
way; therefore, surface roughness is modeled as an increase
in Poiseuille number, as reported by Croce and Gamrat. The
development of the integral perturbation analysis and evaluation of its predictions are described in the following sections.

and the isentropic efficiency of the nozzle is defined as


gnoz

v2o =2
:
v2o;i =2

(5)

It follows from the above relations that for a perfect gas


flowing through nozzles with efficiency gnoz, the tangential
velocity of gas into the rotor at r ro, taken to be equal to
the nozzle exit velocity, is given by
p
vh rro vo gnoz vo; i ;
(6)
where vo, i is computed using Eq. (4), and for choked flow
the nozzle efficiency is given by

II. ANALYSIS

An analysis will now be outlined that describes first the


flow through the nozzle of the turbine and then the flow
through the microchannels of the turbine, while incorporating a treatment of microstructured walls. The resulting equations for velocity and pressure can be used to solve for the
efficiency of the turbine. The closed form solution of the
fluid mechanics equations allows a parametric exploration of
trends in turbine operation.

c1=c

gnoz

cRPt =Pnt crit


h
i:
2cp 1  Po =Pnt c1=c

(7)

Treating the gas flow as an ideal gas with nominally


constant specific heat, Eq. (6) provides the means of determining the rotor gas inlet tangential velocity (vh)rro given
the specified flow conditions for the nozzle.

A. Treatment of the nozzle delivery of flow to the rotor

Before considering the flow in the rotor, a method for


predicting the flow exiting the nozzle in Figure 1 must be
considered. For the purposes of this analysis, the tangential
gas velocity (vh) at the outer radius of the rotor (ro) is taken
to be uniform around the circumference of the rotor and
equal to the nozzle exit velocity determined from one dimensional compressible flow theory.
In expanders of the type considered here, the flow
through the nozzle is often choked. This was the case in expander tests conducted by Rice,2 who reported that virtually
all the pressure drop in the device is in the nozzle and little
pressure drop occurs in the flow through the rotor. The pressure ratio Po=Pnt across the nozzle for choked flow must be
at the critical pressure ratio (Pt=Pnt)crit at the nozzle inlet
temperature. For a perfect gas, this is computed as


Pt
Pnt

crit

2
c1

c=c1
;

(1)

((Pt=Pnt)crit is about 0.528 for air at 350 K (Ref. 10)).


If the nozzle exit velocity is the sonic speed at the
nozzle throat, it can be computed for a perfect gas as
p
(2)
vo;c at cRTt ;
where Tt the nozzle throat temperature for choked flow, is
given by
c1=c

Tt Tnt Pt =Pnt crit

(3)

For isentropic flow through the nozzle, the energy equation dictates that the exit velocity would be
vo;i

r
h
i
c1=c
2cp Tnt 1  Po =Pnt

(4)

B. Analysis of the momentum transport in the rotor

For steady incompressible laminar flow in microchannels between the turbine rotor disks, the governing equations
for the flow are
Continuity:
r  v 0:

(8)

Momentum:
v  rv 

rP
r2 v f:
q

(9)

Treatment of the flow as incompressible is justified by


the observation of Rice2 that minimal pressure drop occurs
in the rotor under typical operating conditions for this type
of expander. For this analysis, the following idealizations are
adopted:
(1) The flow is taken to be steady, laminar, and two-dimensional: vz 0 and the z-direction momentum equation
has a trivial solution.
(2) The flow field is taken to be radially symmetric. The
inlet flow at the rotor outer edge is uniform, resulting in
a flow field that is the same at any angle h. All h derivatives of flow quantities are, therefore, zero.
(3) Body force effects are taken to be zero.
(4) Entrance and exit effects are not considered. Only flow
between adjacent rotating disks is modeled.
With the idealizations noted above, the governing Eqs.
(8) and (9), in cylindrical coordinates, reduce to
continuity:
1 @rvr
0;
r @r

(10)

082003-4

V. D. Romanin and V. P. Carey

Phys. Fluids 23, 082003 (2011)

r-direction momentum:
 
@vr v2h
1 @P
 
vr
@r
r
q @r




1@
@vr
@ 2 vr vr
2  2 ;
r

@r
@z
r
r @r
h-direction momentum:




@vh vr vh
1@
@vh
@ 2 vh vh
vr


2  2 ;
r
@r
r
@r
@z
r
r @r

(11)

(12)

z-direction momentum:
0

 
1 @P
:
q @z

Equation (13) dictates that the pressure is uniform across


the channel at any (r, h) location. For the variations of the radial and tangential velocities, the following solution forms
are postulated:
vr vr r/z;

(14)

vh v^h r/z Ur;

(15)

where


/z


 n 
n1
2z
1
n
b

v^h r

1
b

1
b

b=2
vr dz;

(17)

vh  Udz;

(18)

b=2

sw f

q^
v2h
:
2

l@vh  U=@zzb=z
s

:
2
q^
v2h
q^
vh =2

(19)

where Rec is the Reynolds number defined as

(23)

FPo Po=24;

(24)

n 1 Po=8 3FPo :

(25)

It follows that:

for laminar flow over smooth walls: n 2, Po 24,


FPo 1,
for laminar flow over walls with drag enhancing roughness: n > 2, Po > 24, FPo > 1.

The variation of the velocity profile with n is shown in


Fig. 3.
C. Radial velocity solution from the continuity
equation

Substituting Eq. (14) into Eq. (10) and integrating with


respect to r yields
(20)

For the purposes of this analysis, the tangential shear


interaction of the flow with the disk surface is postulated to
be equivalent to that for laminar Poiseuille flow between
parallel plates,
Po
;
Rec

(22)

and Po is a numerical constant usually referred to as the Poiseuille number. For flow between smooth flat plates, the wellknown laminar flow solution predicts Po 24. For flow
between flat plates with roughened surfaces, experiments79
indicate that a value other than 24 better matches pressure loss
data. We, therefore, define an enhancement number FPo as

For a Newtonian fluid, it follows that

DH 2b;

b=2

b=2

q^
v h DH
;
l

which quantifies the enhancement of shear drag that may


result from disk surface geometry modifications. Note that
Eqs. (19)(23) dictate that for the postulated vh form (15),

where b is the gap distance between disks.


For laminar flow in the tangential direction, the wall
shear is related to the difference between the mean local gas
tangential velocity and the rotor surface tangential velocity
v^h vh  U through the friction factor definition,

Rec

(16)

and vr and v^h are mean velocities defined as


vr r

FIG. 3. (Color online) Variation of velocity profile with n.

(13)

vr / constant Cr :
rvr r

(26)

Integrating Eq. (16) across the channel and using the


fact that
b=2
b=2
/dz 2
/dz b
(27)
b=2

yields
(21)

b=2
b=2

rvr dz

b=2
b=2

r
vr /dz r
vr b bCr C0r :

(28)

082003-5

An integral perturbation model of flow

Phys. Fluids 23, 082003 (2011)

Mass conservation requires that


2pro q

b=2

vr dz 2pro q
vr ro

b=2

b=2
/dz
b=2

vr ro b m_ c ;
2pro q

(29)

where m_ c is the mass flow rate per channel between rotors.


Combining Eqs. (28) and (29) yields the following solution
for the radial velocity:
ro vro
;
(30)
vr 
r
where
vro

m_ c
:
2pro qb

(31)

D. Solution of the tangential and radial momentum


equations

2
^ 0 dW=dn
^
^ 00 d 2 W=dn
^
where W
and W
. Solution of these
equations requires boundary conditions on the dimensionless
^ and P).
^
relative velocity and the dimensionless pressure (W
Here, it is assumed that the gas tangential velocity and the
^ at the outer radius
disk rotational speed are specified, so W
of the disk is specified. It follows that

at n 1;

/ dz 2

b=2

b=2
0

2n 1
b;
/ dz
2n 1
2

(32)

b=2  2 
b=2  2 
d /
d /
dz

2
dz
2
dz
dz2
b=2
0
4n 1
;

b

(33)

^ 0:
P1

^ v^h =Uo 
vh  U=Uo ;
W

(35)

P^ P  Po =qUo2 =2;

(36)

Vro vro =Uo ;

(37)

e 2b=ro ;

(38)

Rem DH =ro

DH m_ c
m_ c DH
2 ;
2pro bl
pro l

(44)

P^ P^0 eP^1 e2 P^2    :

(45)

Substituting results in Eqs. (46)(54).


O(e0):


6FPo  1
^ 0 1  86FPo  1 n W
^ 0;

W
0
3FPo
n
Rem

(46)

12FPo 1 2
^ 2 n2 4W
^0 2n V 2 FPo 96 ;
P^00
V W
0
ro
6FPo  1 n3 ro
Rem n
(47)
P^0 0;

(48)

O(e1):


1
n
0
^
^1 ;
 86FPo  1  W
0 W1
n
Rem

(49)

12FPo 1 ^ ^
^1 ;
2W0 W1 4W
P^01
6FPo  1 n

(50)

^1 0;
at n 1 : W

(39)

P^1 0;

(51)

O(e2):

converts Eqs. (11) and (12) to the forms,

2
Vro
;
Rem n


1 82n 1n ^

W2
n
Rem
 00

^
^0
^0
6FPo  1 nW
W
W
0
0
;


Rem Rem Rem n
6FPo

^20 
W

@ P^ ^0
4n 1 2
^ 2 n2
P
Vro W
3
@n
2n 1n
^ 2n 32n 1
4W

^ 1 e2 W
^W
^ 0 eW
^2    ;
W

^0 W
^0;ro ;
at n 1 : W
(34)

(43)

Equations (42)(43) provide boundary conditions for solution of the dimensionless tangential momentum Equation
(41) and the radial momentum Equation (40), which predicts
the radial pressure distribution.
Since e 2b=ro is much less than 1 in the systems of interest here, we postulate a series expansion solution of the form,

doing so and introducing the dimensionless variables


n r=ro ;

(42)

^ it follows that
In addition, from the definition of P,

The next step is to substitute the postulated solutions


from Eqs. (14) and (15) into the tangential and radial momentum equations (12) and (11), integrate each term across
the microchannel, and use Eq. (27) together with the results,
b=2

^
^o :
W1
W

(40)





2n 1
2n 1 e2
2n 1 e2
00
^
^0
nW


1 W
n1
2n 1 Rem
2n 1 Rem



2n 1 e2 1 82n 1n ^

1
W;
2n 1 Rem n
Rem
(41)

12FPo 1 ^ ^
^2 ;
P^02
2W0 W2 4W
6FPo  1 n
at n 1 :

^2 0; P^2 0:
W

(52)

(53)
(54)

In solving Eqs. (46)(54), the dimensionless parameters


in the equations,

082003-6

V. D. Romanin and V. P. Carey

ni ri =ro ;
^0 W
^0:ro vh;ro  Uo ;
W
ro
Uo

Phys. Fluids 23, 082003 (2011)

(55)
f n
(56)
(57)

Vro vro =Uo ;

(58)

e 2b=ro

(59)

grm

are dictated by the choices for the following physical


parameters:

gi

ri, ro: the inner and outer radii of the disks


b: the gap between the disks, from which we can compute
DH 2b
m_ c : the mass flow rate per channel between rotors
x Uo=ro: the angular rotation rate
vh;ro : the mean tangential velocity at the inlet edge of the
rotor
Po=Pnt: pressure ratio
Tnt: nozzle upstream total temperature.

Also, for choked nozzle flow, the tangential velocity at


the rotor inlet will equal the sonic velocity (vh, ro a). The
^ require that the choices for Mo and
definitions of Mo and W
^0; ro satisfy
W
p Pt =Pnt c1=2c
crit
:
Mo  Uo = cRTt
^0;ro 1
W

(60)

Solving Eqs. (46), (49), and (52) with boundary conditions


(48), (51), and (54) gives the following solutions:
^0
W

 f n
Rem
e
Rem
^
W0;ro 

;
24FPo nef 1 24FPo

^1 0;
W
f n n

^2 e
n ef n gn dn ;
W
n 1

(61)
(62)
(63)

where
gn

^00  nW
^000
^0 =n  W
6FPo  1 W
;

Rem
6FPo

(64)

(65)

And n* is a dummy variable of integration.


With this result, the energy efficiency of the rotor and of
the turbine, respectively, can be computed using

DH m_ c
;
pro2 l

Rem

46FPo  1n2
:
Rem

vh;o Uo  vh;i Ui
;
vh;o Uo

(66)

vh;o Uo  vh;i Ui
;
Dhisen

(67)

which rearrange to
grm 1 

^i ni ni
W
;
^o 1
W

(68)

^nn r =r ;
^i W
W
i o
i

(69)


^o 1  W
^i ni ni c  1M2
W
o
"
:
gi
 c1=c #
Pi
1
Pnt

(70)

The baseline case for comparison with rough wall solutions is that for a smooth wall, or FPo 1. The solution for
W^0 (Eq. (61)) reduces to
^0
W

 20 n2 1
Rem eRem
Re
^
m:
W0;ro 
24
n
24

(71)

The pressure distribution can be found numerically or


analytically by integrating equations (47), (50), and (53).
E. Higher order (e1, e2) terms

In order to evaluate the significance of the higher order


^ and P,
^ results are plotted for two different
solutions of W
operating conditions in Figures 4 (velocity) and 5 (pressure).
^2 and P^2 ) are
Under both scenarios, the 2nd order terms (W
shown to be the same order of magnitude as the 0th order
^0 and for
^0 and P^0 ). Velocity plots for W
terms (W
^ 1 e2 W
^ 0 eW
^2 fall nearly directly on top of each other
W
^ For values of e as high as 1=10, much
(similarly for P).
^2
larger than are found in most systems of interest, both e2 W
and e2 P^2 are less than 0.1% of the value of the 0th order

FIG. 4. (Color online) Comparison of


velocity plots for 0th order and 2nd
order velocity solutions. In both (a) and
^0 (solid line) and for
(b), the plots for W
^1 e2 W
^2 (dashed line) are
^0 eW
W
nearly coincident. The dotted-dash line
^2 ) is shown to be the same order of
(W
^0 , thus making it negligimagnitude as W
ble when multiplied by e2. (a) Case 1:
^0 2, Re 10, ni 0.2, Vro 0.05,
W
m
e 1=20; choked flow and (b) case 2:
^0 1:1, Rem 5, ni 0.2, Vro 0.05,
W
e 1/20; choked flow.

082003-7

An integral perturbation model of flow

Phys. Fluids 23, 082003 (2011)

FIG. 5. (Color online) Comparison of


pressure plots for 0th order and 2nd
order velocity solutions. In both (a) and
(b), the plots for P^0 (solid line) and for
P^0 eP^1 e2 P^2 (dashed line) are nearly
^2 ) is
coincident. The dotted-dash line (W
shown to be the same order of magni^0 , thus making it negligible
tude as W
when multiplied by e2. (a) Case 1:
^0 2, Re 10, ni 0.2, Vro 0.05,
W
m
e 1=20; choked flow and (b) case 2:
^0 1:1, Re 5, ni 0.2, Vro 0.05,
W
m
e 1=20; choked flow.

term for the two cases shown. Note that Eq. (62) along with
^1 P^1 0. Henceforth, we
Eqs. (50) and (51) show that W
can neglect the 1st and 2nd order terms, and only the 0th
^0 and P^0 ) will be considered.
order terms (W
III. COMPARISON OF SMOOTH WALL CASE WITH
EARLIER MODEL PREDICTIONS

^0 solution corresponds closely with the solution


The W
developed by Carey,6 only differing by numerical constants.
A comparison of results with the model from Careys earlier
model is shown in Figure 6. Carey6 made several assumptions, including ignoring radial pressure effects, treating the
flow as inviscid with a body force representation of drag, and
ignoring z-derivatives of velocity. In the present analysis,
initial assumptions were more conservative and terms were
removed based on the arguments of the perturbation analysis.
The similarities in the results of this analysis with that of
Carey verify that the assumptions made were valid. Additionally, Careys analysis was compared extensively with
experimental data in Romanin and Carey,11 so a close correlation between the two approaches is encouraging.
A comparison with previous experimental data11 can be
seen in Table I. The agreement between test data and the
model predicted efficiency is reasonable considering the
uncertainty of the test data, and is similar to the accuracy of
the earlier model developed by Carey.6 However, due to the
small range of nondimensional parameters explored in the
experimental data, a more extensive comparison of test data
with the model may generate additional insights.

Figure 7 shows a 3D plot of turbine efficiency as a func^0;ro for the operating parameters in the first
tion of Rem and W
four lines of Table I. The data points are overlaid on top of
the surface plot, which shows how the data compares to the
predictions of efficiency. The figure shows that the analytical
^0;ro will increase
model correctly predicts that decreasing W
^0;ro
efficiency, and suggests that decreasing both Rem and W
can dramatically improve performance.
A 3D plot of turbine efficiency with typical operating
^0;ro , is shown
parameters, and over ideal ranges of Rem and W
^0;ro , the analysis
in Figure 8. At very low values of Rem and W
shows that very high turbine efficiencies can be achieved.
Practical issues arise when generating power in microchan^0;ro and Rem .
nels such as these at very low values of W
^0;ro requires the rotor to be spinning at speeds
Reducing W
very close to the air inlet speeds. This is difficult to achieve
because it requires very low rotor torque and high speeds,
which may require high gear ratios to achieve in some applications. Also, lower Reynolds numbers require very small
disk spacings (b) and larger disk radii (ro).
IV. MODELING OF FLOW VELOCITY WITH
ROUGHENED OR MICROSTRUCTURED SURFACES
(FPO > 1)

Now that the perturbation analysis has resulted in equations that define the operating conditions and efficiency of
the turbine as a function of FPo, we can analyze the effect of
surface roughness on turbine performance. Developing a
direct correlation between surface roughness and FPo is a

FIG. 6. (Color online) Comparison of


solutions from the perturbation method
and the model developed by Carey.6 (a)
^0 2, Rem 10, ni 0.2,
Case 1: W
Vro 0.05; choked flow. The analysis
predicts a turbine isentropic efficiency of
gi 26.1% while the analysis by Carey6
predicts gi 27.0% (b) case 2:
^0 1:1, Rem 5, ni 0.2, Vro 0.05;
W
choked flow. The analysis predicts a turbine isentropic efficiency of gi 42.3%
while the analysis by Carey6 predicts
gi 42.5%.

082003-8

V. D. Romanin and V. P. Carey

Phys. Fluids 23, 082003 (2011)

TABLE I. Comparison of analysis with experimental data from Romanin


and Carey.11
m_ c
g/s

x
rad/s

Rem

^0;ro
W

gi;exp
(%)

gi;model
(%)

1.64
1.66
1.65
1.65
2.37
2.40
2.38
2.38

450
784
953
1110
708
1110
1370
1590

47.5
48.1
47.8
47.8
68.6
69.5
68.9
68.9

18.2
10.0
8.1
6.8
11.2
6.8
5.3
4.4

3.1
4.9
5.9
6.8
3.0
4.3
5.3
6.0

2.3
3.6
4.4
5.1
2.3
3.3
4.1
4.8

detailed process that involves characterizing specific geometric properties of the roughness features and is beyond the
scope of this analysis. Here, we will only discuss the effects
of increasing FPo. Kandlikar8 reported values for FPo as high
as 3.5 for roughened surfaces in microchannels, so values up
to FPo 3.5 will be considered.
A. Discussion of the velocity and pressure fields

Figure 9 shows that the velocity profile is significantly


altered by using a roughened surface. Equation (70) shows
^i should be minimized to increase efthat the exit velocity W
ficiency and indeed the efficiency does increase with FPo.
FPo 2 results in a turbine isentropic efficiency of
gi 45.1%, compared to an efficiency of gi 42.3% for
FPo 1.
Figure 10 shows the dimensionless pressure P^ as a function of n for several values of FPo. The figure shows that the
dimensionless pressure decreases with increasing FPo. This
can be attributed to the competing effects of centripetal force
and radial pressure. Increasing surface roughness decreases
the velocity and, therefore, the centripetal force is decreased.
The required pressure field to balance the centripetal force
on the fluid is, therefore, also decreased. It is important to
note that this does not contradict the conventional knowledge
that the pressure drop increases along the direction of the

FIG. 7. (Color online) A plot of experimental data from the first four lines
of Table I with a surface plot of efficiency (gi) from Eq. (71) (FPo 1
(smooth wall), c 1.4 (air), ni 0.45, Pi=Pnt 0.4; choked flow).

FIG. 8. (Color online) A plot of efficiency (gi) as a function of dimension^0; ro ) and modified Reynolds
less tangential velocity difference at the inlet (W
number (Rem ) for typical operating conditions: FPo 1 (smooth wall),
c 1.4 (air), ni 0.2, Pi=Pnt 0.5; choked flow.

flow as the surface roughness is increased. The pressure drop


described here is in the radial direction, while the fluid flow
has both a radial and circumferential component.
B. Performance enhancement due to mictrostructured
surfaces

Over the entire range of values for FPo discussed by


Croce,8 Figure 11 shows that efficiency increases a total of
3.8 percentage points, which amounts to a 9.2% improvement in performance over a smooth wall.
Figure 12 shows a surface plot of efficiency as a func^0;ro . It is
tion of two non-dimensional parameters, Rem and W
shown that increasing surface roughness can yield especially
significant performance improvements for higher Reynolds
numbers rather than lower. Similar trends to those reported
by Carey6 and Romanin11 can be seen in Figure 12; it is clear
that high efficiency turbine designs should strive for Reynolds numbers and dimensionless inlet velocity differences to
be as small as possible. It is also shown that penalties due to

^0 1:1,
FIG. 9. (Color online) Velocity vs. n for several values of FPo. W
Rem 5, ni 0.2; choked flow.

082003-9

An integral perturbation model of flow

Phys. Fluids 23, 082003 (2011)

^ vs. n for several values


FIG. 10. (Color online) Dimensionless pressure (P)
^0 1:1, Re 5, ni 0.2, Vro 0.05; choked flow.
of FPo. W
m

FIG. 11. (Color online) Efficiency (gi) vs. FPo for ni 0.2 and choked flow.

^0;ro are less dramatic for roughhigher values of Rem and W


ened surfaces.
By taking advantage of microstructured surfaces, larger
disk gaps, and smaller disks can be used while limiting penalties to efficiency. For example, the nondimensional turbine parameters outlined in case 2 (see Figure 4(b)) can be used to
deduce the physical parameters in the right hand side of Eqs.
(55)(59). Using this set of physical parameters, the Poiseuille
number can be doubled (FPo 2), and the radius can be
decreased while keeping other parameters constant until the
efficiency is equivalent to that achieved by the parameters
from case 2 (Figure 4(b)). This process results in a turbine radius of ro 18.6 cm, down from ro 34.7 cm in the smooth
wall case. In other words, doubling the Poiseuille number, in
this case, allowed for a 46% reduction in turbine size with
equivalent performance. Similar trade-offs with other physical
parameters can be explored, allowing greater flexibility in
high-efficiency turbine design. The values outlined in this
example are not universal, however, as Figure 13 shows that
performance increases due to roughened surfaces vary with
non-dimensional parameters (e.g., performance increases are
less significant at low modified Reynolds (Rem ) numbers).

any h location a the rotor inlet (n r=ro 1), over time, the
fluid traces an (r, h) path through the channel between adjacent disks that is determined by integrating the differential
relations,

V. STREAMLINE VISUALIZATION

The model theory developed here also provides the


means to determine the trajectory of streamlines in the rotor
using the h and r direction velocity components. Starting at

rdh vh dt;

(72)

dr vr dt:

(73)

Combining the above equations yields the following differential equation that can be integrated to determine the dependence of h with r along the streamline,
 
dh
vh
:
(74)

dr st vr r
Note that since the velocities are functions only of r, the
entire right side of the above equation is a function of r. In
terms of the dimensionless variables described above, the
streamline differential Equation (74) can be converted to the
form,
 
^
dh
nW

;
(75)
dn st
Vro
where Vro is the ratio of radial gas velocity to rotor tangential
velocity at the outer edge of the rotor,
Vro

vro
m_ c

:
Uo 2pro bqo Uo

(76)

FIG. 12. (Color online) A 3D surface


plot of efficiency (gi) as a function of the
inlet dimensionless tangential velocity
^0;ro ) and Reynolds number
difference (W
(Rem ) for typical operating parameters
(c 1.4 (air), ni 0.2, Pi=Pnt 0.5;
choked flow). (a) FPo 1 and (b)
FPo 2.

082003-10

V. D. Romanin and V. P. Carey

FIG. 13. (Color online) A 3D surface plot of


the percent increase in efficiency

resulting from increasing FPo from 1 to 2 gi;FPo 2  gi;FPo 1 =gi;FPo 1 as a


^
function of the inlet dimensionless tangential velocity difference (W0;ro ) and
Reynolds number (Rem ) for typical operating parameters (FPo 1 and FPo 2,
c 1.4 (air), ni 0.2, Pi=Pnt 0.5, choked flow).

Rotor streamlines determined by integrating Equation


^o 3:0, (DH=ro)Rem 5.0, Vro 0.05, and
(75) for W
ni 0.2 are shown in Figure 14. Flow along one streamline
enters the rotor at h 0 , whereas the other streamline
begins at h 180 . The model can be used to predict how
the inward spiral path of the flow changes as the governing
parameters are altered. Figure 14 shows streamlines from the
roughened surfaces have a larger radial component than
those generated with a smooth surface. An analytical method
for predicting streamlines can be useful in designing complex disk surface geometries that consider flow direction,
such as surface contours or airfoils.

VI. CONCLUSIONS

It has been shown that the use of an integral perturbation


analysis scheme allows construction of a series expansion solution of the governing equations for rotating microchannel
flow between the rotor disks of a Tesla-type drag turbine. Two
idealizations in the model may limit its accuracy. One is the
postulated tangential velocity profile used to facilitate the integral analysis. The other is the idealization of the inlet flow as
being uniform over the outer perimeter of the disk. Real turbines of this type have a discrete number of nozzles that

Phys. Fluids 23, 082003 (2011)

deliver inlet flow at specific locations. The gap between the


outer edge of the rotor disk and the housing generally allows
the flow to distribute itself somewhat over the perimeter.
Thus, the idealization of uniform inlet flow may be a good
one if the flow is delivered by several nozzles around the perimeter. Clearly, however, the model developed here is
expected to be most accurate under conditions, where the
postulated tangential velocity profile and the idealization of
uniform inlet velocity are consistent with expected actual conditions in the flow.
Although the model may be somewhat limited by the
idealizations described above, it has several very useful
advantages. One is that it provides a rigorous approach that
retains the full complement of momentum and viscous
effects to consistent levels of approximation in the series
solution. Another is that by constructing the solution in
dimensionless form, the analysis directly indicates all the
dimensionless parameters that dictate the flow and transport, and, in terms of these dimensionless parameters, it
provides a direct assessment of the relative importance of
viscous, pressure, and momentum effects in different directions in the flow. Our analysis also indicated that closed
form equations can be obtained for the lowest order contribution to the series expansion solution, and the higher
order term contributions are very small for conditions of
practical interest. This provides simple mathematical relations that can be used to compute the flow field velocity
components and the efficiency of the turbine, to very good
accuracy, from values of the dimensionless parameters for
the design of interest. In addition, it has been demonstrated
here that this solution formulation facilitates modeling of
enhanced rotor drag due to rotor surface microstructuring.
The type of drag turbine of interest here is one of very few
instances in which enhancement of drag is advantageous in
fluid machinery. We have demonstrated that by parameterizing the roughness in terms of the surface Poiseuille number ratio (FPo), the model analysis developed here can be
used to predict the enhancing effect of rotor surface microstructuring on turbine performance for a wide variety of
surface microstructure geometries.
Predictions of the model analysis have been shown
to agree well with available experimental drag turbine
performance data. However, the available data are limited to conditions corresponding to low isentropic

FIG. 14. (Color online) Streamlines for


^0 1:1, Rem 5, ni 0.2, and
W
Vro 0.05 (a) FPo 1 and (b) FPo 2.

082003-11

An integral perturbation model of flow

efficiency. A particularly interesting prediction of the


model developed here is that low Reynolds numbers and
high rotor speeds result in the highest turbine isentropic
efficiencies. Specifically, for modified Reynolds numbers
(Rem ) less than 1.2 and dimensionless inlet velocity dif^0 ) less than 1.2 (or Mo > 0.41 for choked flow),
ference ( W
efficiencies up to and exceeding 80% can be achieved. In
addition to low Reynolds numbers and high rotor speeds,
roughened or microstructured surfaces can provide efficiency benefits that can further improve turbine performance. Surface roughness was shown to improve turbine
efficiency by 9.2% in one example case.
The results of this investigation clearly indicate a path
of design changes that can significantly improve the energy
efficiency performance of Tesla-type disk-rotor drag turbines. The trends that indicate this path are supported by
available experimental data. However, because the ranges of
experimental data are limited to low efficiency conditions,
we were not able to validate the model predictions into the
range of conditions predicted to produce high efficiencies.
Comparison of the predictions of the model presented here
with new performance data for Tesla-type drag turbines at
lower Reynolds numbers (Rem ) and inlet velocity difference
^0 ) conditions are needed to fully explore the accuracy of
(W
the model predictions. This model, nevertheless, offers a useful means to explore parametric trends in designs of Teslatype drag turbines, and it can be useful in comparisons with
predictions of more detailed computational fluid dynamics
models of the flow in these types of turbines.

Phys. Fluids 23, 082003 (2011)

ACKNOWLEDGMENTS

Support for this research by the UC Center for Information Technology Research in the Interest of Society (CITRIS) is gratefully acknowledged.
1

N. Tesla, Turbine, U.S. Patent No. 1,061,206 (May 1913).


W. Rice, An analytical and experimental investigation of multiple disk
turbines, J. Eng. Power 87, 29 (1965).
3
M. C. Breiter and K. Pohlhausen, Laminar flow between two parallel
rotating disks, Tech. Rep. ARL 62-318 (Aeronautical Research Laboratories, Wright-patterson Air Force Base, Ohio, 1962).
4
G. P. Hoya and A. Guha, The design of a test rig and study of the performance and efficiency of a tesla disc turbine, Proc. Inst. Mech. Eng.,
Part A 223, 451 (2009).
5
A. Guha and B. Smiley, Experiment and analysis for an improved design
of the inlet and nozzle in tesla disc turbines, Proc. Inst. Mech. Eng., Part
A 224, 261 (2010).
6
V. Carey, Assessment of tesla turbine performance for small scale rankine combined heat and power systems, J. Eng. Gas Turbines Power 132,
122301 (2010).
7
S. G. Kandlikar, D. Schmitt, A. L. Carrano, and J. B. Taylor, Characterization
of surface roughness effects on pressure drop in single-phase flow in
minichannels, Phys. Fluids 17, 100606 (2005).
8
G. Croce, P. Dagaro, and C. Nonino, Three-dimensional roughness
effect on microchannel heat transfer and pressure drop, Int. J. Heat Mass
Transfer 50, 5249 (2007).
9
G. Gamrat, M. Favro-Marinet, S. Le Person, R. Bavie`re, and F. Ayela,
An experimental study and modelling of roughness effects on laminar
flow in microchannels, J. Fluid Mech. 594, 399 (2008).
10
B. R. Munson, D. F. Young, and T. H. Okiishi, Fundamentals of Fluid
Mechanics, 5th ed. (Wiley, New York, 2006).
11
V. Romanin and V. Carey, Strategies for performance enhancement of
tesla turbines for combined heat and power applications, ASME 2010
Energy Sustainability Conference Proceedings, Phoenix, Arizona, USA, 2,
5764 (ASME, New York, 2010).
2

Physics of Fluids is copyrighted by the American Institute of Physics (AIP). Redistribution of journal material
is subject to the AIP online journal license and/or AIP copyright. For more information, see
http://ojps.aip.org/phf/phfcr.jsp

You might also like