1
The Graetz Problem 
 R. Shankar Subramanian  
As a good model problem, we consider steady state heat transfer to fluid in steady flow through a 
tube.    The  fluid  enters  the  tube  at  a  temperature 
0
T   and  encounters  a  wall  temperature  at 
w
T , 
which can be larger or smaller than 
0
T .  A simple version of this problem was first analyzed by 
Graetz (1883).  A sketch of the system is shown below.                
Objective  
To obtain the steady temperature distribution  (   ) , T   r   z  in the fluid, and to calculate the rate of heat 
transfer from the wall to the fluid  
Assumptions  
1. Steady fully developed laminar flow; steady temperature field.  
2.  Constant  physical  properties  , , ,
  p
k   C     --  This  assumption  also  implies  incompressible 
Newtonian flow.  
3. Axisymmetric temperature field    0
T
. 
 
4. Negligible viscous dissipation 
 
 
Fluid at
0
T
r
z
R
(   )
,
  w
T   R  z   T =
 
 
 
2 
Velocity Field 
 
Poiseuille Flow 
 
0; 0
r
v   v
=   =  
 
( )
2
0 2
1
z
r
v   r   v
R
|   |
=   
   |
\   .
   
0
: v  Maximum velocity existing at the centerline 
 
Energy Equation 
 
Subject to assumption (2), Equation (B.9.2) from Bird et al. (page 850) can be written as follows. 
 
   
1 0 0
r
p
v   v
T
C
t
=   =
  r
v +
  v
T
r
 
  +
2
2 2
3 4
1 1
z
T   T
v
r   z
T   T
k   r
r   r   r   r
   (
   
+    (
   
   (
   
       |   |
=   +
   |
      
\   .
2
2   v
T
z
  
   (
+   +   
   (
   (
   
 
 
and therefore, simplified to 
 
2 2
0 2 2
1
1
  r   T   T   T
v   r
R   z   r   r   r   z
|   |      (           |   |
   =   +
   |      |    (
         
\   .
\   .      
 
 
where 
(   )
/
  p
k   C     =  is the thermal diffusivity of the fluid. 
 
Boundary Conditions 
 
Inlet:      (   )
0
, 0 T   r   T =  
 
Wall:      (   ) ,
  w
T   R  z   T =  
 
Centerline:    (   ) 0, T   z  is finite  
    or   (   ) 0, 0
T
z
r
 
 
 
 
3 
Because  of  the  appearance  of  the  axial  conduction  term  in  the  governing  differential 
equation,  we  should  write  another  boundary  condition  in  the  z coordinate.    But  actually,  the 
inlet  condition  written  above  is  incompatible  with  the  inclusion  of  axial  conduction  in  the 
problem, because conduction will lead to some of the information about the step change in wall 
temperature  at  the  inlet  to  propagate  backward.    As  we  shall  see  shortly,  well  neglect  axial 
conduction, which will obviate the need for writing a second condition in the  z coordinate. 
 
Non-Dimensionalization 
 
We shall use the following scheme for scaling (or non-dimensionalizing) the variables. 
 
0
w
w
T   T
T   T
 , 
r
Y
R
= ,     
z
Z
R Pe
= ,   where  the Pclet Number 
0
Rv
Pe
= . 
 
This permits us to transform the governing differential equation and boundary  conditions to the 
following form. 
 
(   )
2
2
2 2
1 1
1   Y   Y
Z   Y   Y   Y   Pe   Z
                 |   |
   =   +
   |
         
\   .
 
 
          (   ) , 0 1 Y    =  
          (   ) 1, 0 Z    =  
          (   ) 0, Z   is finite 
or  (   ) 0, 0 Z
Y
 
  =
 
 
The Pclet Number 
 
The  Pclet  number  plays  the  same  role  in  heat  transport  as  the  Reynolds  number  does  in  fluid 
mechanics.    First,  we  note  that  the  Pclet  number  is  the  product  of  the  Reynolds  and  Prandtl 
numbers. 
 
0 0
Re Pr
Rv   Rv
Pe
  
      
=   =      =     
 
The physical significance of the Pclet number can be inferred by recasting it slightly. 
 
0
Rate of energy transport by convection
Rate of energy transport by conduction
p
v C   T
Pe
T
k
R
   
=   =
 
 
 
 
4 
Note  that  the  numerator  represents  the  order  of  magnitude  of  the  convective  flux  in  the  main 
flow direction, whereas the denominator stands for the order of magnitude of the conduction flux 
in  the  radial  direction.    If  we  wish  to  compare  the  rates  of  energy  transport  by  these  two 
mechanisms  in  the  same  direction,  we  can  multiply  the  Pclet  number  by  / L   R  where  L   is  a 
characteristic length in the axial direction. 
For large values of  Pe , we can see that 
2
1
1
Pe
   .  Therefore, in the scaled energy equation, the 
term involving axial conduction can be safely neglected.  Physically, there are  two mechanisms 
for  transporting  energy  in  the  axial  direction,  namely,  convection  and  conduction.    Because  the 
Pclet  number  is  large,  we  are  able  to  neglect  transport  by  conduction  in  comparison  with 
transport  by  convection.    On  the  other  hand,  in  the  radial  direction,  there  is  only  a  single 
mechanism  for  transport  of  energy,  namely  conduction.    By  performing  calculations  including 
conduction in the axial direction, it has been established that it is safe to neglect axial conduction 
for  100. Pe        To learn  about  how  to  include  axial  conduction,  you  can  consult  the  articles  by 
Davis (1973), Acrivos (1980), and Papoutsakis et al. (1980).  
 
Let us make a sample  calculation  of the Pclet number for laminar  flow  heat transfer in a tube.   
The thermal diffusivity of common liquids is typically in the range 
2
7 7
10 2 10
  m
s
   
  , and well 
use the larger limit.  Choose 
 
2
7
0
10 , 0.05 , 2 10
m   m
R   mm   v
s   s
  
=   =   =     
 
This  yields,  2, 500 Pe = ,  which  is  much  larger  than  100.    We  can  check  to  see  if  the  flow  is 
laminar by calculating the Reynolds number.   If the fluid is water, 
2
6
10
  m
s
  
 , which  yields a 
Prandtl number  Pr 5
=   = .  Therefore, the Reynolds number is  Re 500 = , which is comfortably 
in the laminar flow regime.    
The final version of the scaled energy equation is   
(   )
2
1
1   Y   Y
Z   Y   Y   Y
           |   |
   =
     |      
\   .  
We  can  solve  this  equation  by  separation  of  variables,  because  the  boundary  conditions  in  the 
Y  coordinate are homogeneous.  The method of separation of variables yields an infinite series 
solution for the scaled temperature field. 
(   )   (  )
2
1
,
  n
Z
n   n
n
Y  Z   A e   Y
=
=
 
 
 
 
5 
In  the  above  solution,  the  functions  (  )
n
  Y    are  the characteristic functions or eigenfunctions of 
a proper Sturm-Liouville system. 
 
(   )
2 2
1
1 0
d   d
Y   Y
Y  dY   dY
     
|   |
+      =
   |
\   .
 
 
( )   ( ) 0 0 or 0
d
dY
   = is finite 
( ) 1 0    =  
 
The  above  ordinary  differential  equation  for  (  ) Y    can  be  solved  by  applying  the  following 
transformations  to  both  the  dependent  and  the  independent  variables  (Lauwerier,  1951,  Davis, 
1973). 
(   )   (  )
2
2
X
X   Y   W   X   e   Y     =   =  
 
This leads to the following differential equation for  (   ) W   X . 
 
(   )
2
2
1
1 0
4 2
d  W   dW
X   X   W
dX   dX
 |   |
+      +      =
   |
\   .
 
 
This is known as Kummers equation.   It has two linearly independent solutions, but only one is 
bounded  at  0. X =     Because  ( ) 0    must  be  bounded,  we  must  require  that  ( ) 0 W   also  remain 
bounded.  This rules out the singular solution, leaving us with the regular solution 
 
(   )
1
, 1,
2 4
W   X   c M   X
 |   |
=   
   |
\   .
 
where  c  is  an  arbitrary  multiplicative  constant.  The  function  (   ) , , M  a b  X   is  the  confluent 
hypergeometric function, or Kummer function, and is discussed in Chapter 13 of the Handbook 
of  Mathematical  Functions  by  M.  Abramowitz  and  I.  A.  Stegun,    It  is  an  extension  of  the 
exponential function, and is written in the form of the following series. 
(   )
  (   )
(   )
(   )   (   )
(   )   (   )
2
1
, , 1
1 2!
1 1
1 1 !
n
a   a
a   X
M  a b  X   X
b   b  b
a   a   a   n
  X
b  b   b   n   n
+
=   +   +   +
+
+   +   
+   +
+   +   
 
You can see that when  a   b = ,  
 
(   ) , ,
  X
M  a  a  X   e =  
 
 
 
6 
 
Application  of  the  boundary  condition  at  the  tube  wall,  ( ) 1 0    = ,  leads  to  the  following 
transcendental equation for the eigenvalues. 
 
1
, 1, 0
2 4
M
  
  
|   |
   =
   |
\   .
 
 
The above equation has infinitely many discrete solutions for   , which we designate as 
n
 , with 
n  assuming positive integer values beginning from 1.  Corresponding to each value 
n
 , there is 
an eigenfunction  (  )
n
  Y   given by 
(  )   (   )
2
2
2
n
Y
n   n   n
Y   e   W   Y
=  
 
The first few eigenvalues are reported in the table. 
 
n  
n
  
2
n
  
1  2.7044  7.3136 
2  6.6790  44.609 
3  10.673  113.92 
4  14.671  215.24 
5  18.670  348.57 
Note  that  technically 
{   }
2
n
   is  the  set  of  eigenvalues,  even  though  we  use  the  term  loosely  to 
designate  {   }
n
 as that set for convenience.  
 
The  most  important  property  of  a  proper  Sturm-Liouville  system  is  that  the  eigenfunctions  are 
orthogonal  with  respect  to  a  weighting  function  that  is  specific  to  that  system.    In  the  present 
case, the orthogonality property of the eigenfunctions can be stated as follows. 
 
(  )   (  )   (   )
1
2
0
1 0,
m   n
Y   Y   Y   Y   dY   m   n          =   
 
 
Using  this  orthogonality  property,  it  is  possible  to  obtain  a  result  for  the  coefficients  in  the 
solution by separation variables. 
 
(  )   (   )
(  )   (   )
1
2
0
1
2 2
0
1
1
n
n
n
Y   Y   Y   dY
A
Y   Y   Y   dY
 
 
 
 
 
7 
The Heat Transfer Coefficient 
 
The  heat  flux  from  the  wall  to  the  fluid,  ( )
w
q   z   is  a  function  of  axial  position.    It  can  be 
calculated directly by using the result 
 
( )   (   ) ,
w
T
q   z   k   R  z
r
 
but as we noted earlier, it is customary to define a heat transfer coefficient  ( ) h   z  via 
 
( )   ( )(   )
w   w   b
q   z   h   z   T   T =     
 
where the bulk or cup-mixing average temperature 
b
T  is introduced.   The way to experimentally 
determine the bulk average temperature is to collect the fluid coming out of the system at a given 
axial  location,  mix  it  completely,  and  measure  its  temperature.    The  mathematical  definition  of 
the bulk average temperature was given in an earlier section. 
 
(   )
0
0
2 ( ) ,
2 ( )
R
b   R
rV  r T   r   z   dr
T
rV  r   dr
 
 
where  the  velocity  field  ( )   (   )
2 2
0
1 / V   r   v   r   R =    .    You  can  see  from  the  definition  of  the  heat 
transfer  coefficient  that  it  is  related  to  the  temperature  gradient  at  the  tube  wall  in  a  simple 
manner. 
( )
  (   )
(   )
,
w   b
T
k   R  z
r
h   z
T   T
 
 
We can define a dimensionless heat transfer coefficient, which is known as the Nusselt number. 
(   )
  (   )
(   )
1,
2
2
b
Z
hR
  Y
Nu   Z
k   Z
=   =   
where 
b
  is the dimensionless bulk average temperature. 
 
 
 
 
 
 
 
 
 
8 
By substituting from the infinite series solution  for both the numerator and the denominator, the 
Nusselt number can be written as follows. 
 
(   )
  ( )
(   )   (  )
2
2
1
1
2
1
0
1
2
4 1
n
n
Z
  n
n
n
Z
n   n
n
d
A e
dY
Nu   Z
A e   Y   Y   Y   dY
=
= 
  
 
 
The denominator can be simplified by using the governing differential equation for  (  )
n
  Y  , along 
with the boundary conditions, to finally yield the following result. 
( )
( )
2
2
1
2
1
1
2 1
n
n
Z
  n
n
n
Z
n
n
n   n
d
A e
dY
Nu
d e
A
dY
=
 
=
=
  
 
 
We  can  see  that  for  large  Z ,  only  the  first  term  in  the  infinite  series  in  the  numerator,  and 
likewise  the  first  term  in  the  infinite  series  in  the  denominator,  is  important.    Therefore,  as 
Z , 
2
1
3.656
2
Nu
  
   = . 
 
The  sketch  qualitatively  illustrates  the  behavior  of  the  Nusselt  number  as  a  function  of 
dimensionless axial position. 
 
 
 
 
 
3.656
Nu
Dimensionless Axial Position   Z
0
0
 
 
 
9 
A  similar  analysis  is  possible  in  the  case  of  a  uniform  wall  flux  boundary  condition.   
Extensions of the Graetz solution by separation of variables have been made in a variety of ways, 
accommodating  non-Newtonian  flow,  turbulent  flow,  and  other  geometries  besides  a  circular 
tube. 
 
The Lvque Approximation 
 
 The  orthogonal  function  expansion  solution  obtained  above  is  convergent  at  all  values  of  the 
axial position, but convergence is very slow at the inlet is approached.  The main reason for this 
is  the  assistance  provided  by 
2
n
Z
e
  
  in  accelerating  convergence  for  sufficiently  large  values  of 
Z .    Lvque  (1928)  considered  the  thermal  entrance  region  in  a  tube  and  developed  an 
alternative  solution,  which  is  useful  precisely  where  the  orthogonal  function  expansion 
converges too slowly.   
 
 
 
 
 
 
 
 
 
 
 
 
 
 
We shall now construct the Lvque solution which is built on the assumption that the thickness 
of the thermal boundary layer 
t
  R     .  This assumption leads to the following simplifications. 
 
1.  Curvature  effects  can  be  neglected  in  the  radial  conduction  term.    This  means  that  the 
derivative  
1   T
r
r   r   r
    |   |
   |
   
\   .
 can be approximated by 
2
2
1   T   T
R
R   r   r   r
       |   |
 =
   |
      
\   .
. 
 
2. Because we are only interested in the velocity distribution within the thermal boundary layer, 
we expand the velocity field in a Taylor series in distance measured from the tube wall and retain 
the first non-zero term. 
 
Defining  x   R   r =    , we can rewrite the velocity distribution as 
 
( )
  (   )
2
2
0 0 0 2 2
1 2 2
z
R   x
  x   x   x
v   r   v   v   v
R   R   R   R
|   |
   |   |
=      =          |
     |
   |
  \   .
\   .
 
Fluid at
0
T
r
z
R
(   )
,
  w
T   R  z   T =
t
 
 
 
10 
 
Recall  that  a  power  series  obtained  by  any  method  is  a  Taylor  series.    The  above  approach  is 
simpler  than  working  out  the  derivatives  of  ( )
z
v   r   in  the  x  coordinate,  evaluating  them  at  the 
wall, and constructing the Taylor series. 
 
3. Because the conditions outside the thermal boundary layer are those in the fluid entering the 
tube,  we  shall  use  the  boundary  condition  (   )
0
T   x   T     instead  of  the  centerline  boundary 
condition employed in obtaining the Graetz solution. 
 
Beginning  with  the  simplified  energy  equation  in  which  axial  conduction  has  been  neglected 
already, and invoking the above assumptions, we have the following governing equation for the 
temperature field. 
 
 
2
0 2
2
  x   T   T
v
R   z   x
   
=
   
 
 
where  the  chain  rule  has  been  used  to  transform  the  second  derivative  in  r   to  the  second 
derivative in  x .    
 
The temperature field  (   ) , T   x  z  satisfies the following boundary conditions. 
    (   )
0
, 0 T   x   T =  
    (   ) 0,
  w
T   z   T =  
(   )
0
, T   z   T    =  
We  shall  work  with  a  dimensionless  version  of  these  equations.    For  consistency,  we  scale  the 
temperature and axial coordinate in the same manner as before. 
 
0
w
w
T   T
T   T
   
z
Z
R Pe
=  
 
We  define  a  new  scaled  distance  from  the  wall  via  / X   x   R = .    The  scaled  governing  equation 
and boundary conditions are given below. 
 
2
2
2X
Z   X
       
=
   
 
 
(   ) , 0 1 X    =  
(   ) 0, 0 Z    =  
(   ) , 1 Z     =  
 
 
 
 
11 
The  similarity  of  this  governing  equation  and  boundary  conditions  to  those  in  the  fluid 
mechanical  problem  in  which  we  solved  for  the  velocity  distribution  between  two  plates  when 
one of them is held fixed and the other is moved suddenly is not a coincidence.   For small values 
of time in the fluid mechanical problem, we replaced the boundary condition at the top plate with 
one at an infinite distance from the suddenly moved plate, and used the method of combination 
of  variables  to  solve  the  equations.    It  would  be  worthwhile  for  you  to  go  back  and  review  the 
notes on combination of variables at this stage. 
 
By invoking ideas very similar to those used in the fluid mechanical problem, we postulate that a 
similarity  solution  exists  for  the  temperature  field  in  the  present  problem.    That  is,  we  assume 
(   )   (  ) , X  Z   F     =  where the similarity variable  (   ) / X   Z     =  .  The variable  (   ) Z   represents the 
scaled thermal boundary layer thickness, and is unknown at this stage.  We make the necessary 
transformations using the chain rule. 
 
2
dF   X  d   dF   d   dF
Z   Z  d   dZ   d   dZ  d
            
            
      |   |
=   =      =  
   |
   
  \   .
 
 
1 dF   dF
X   X  d   d
   
      
   
=   =
   
 
 
(   )
2 2
2 2 2
1 1 1 1 dF   dF   d   dF   d  F
X   X   Z   d   X   d   X  d   d   d
   
                        
   (
     (      (          
=   =   =   =
   (
     (      (
         
         
   
 
 
Using  these  results,  the  partial  differential  equation  for  (   ) , X  Z    is  transformed  to  an  ordinary 
differential equation for  (  ) F  . 
 
2
2 2
2
2 0
d  F   d   dF
d   dZ   d
   
   
|   |
+   =
   |
\   .
 
 
It  is  evident  that  the  similarity  hypothesis  will  fail  unless  the  quantity  inside  the  parentheses  is 
required  to  be  independent  of  Z ,  and  therefore,  a  constant.    For  convenience,  we  set  this 
constant to 3/2.  Therefore, we have an ordinary differential equation for  (  ) F   and another for 
(   ) Z  . 
 
2
2
2
3 0
d  F   dF
d   d
   
+   =  
 
2
3
2
d
dZ
   =  
 
 
 
12 
 
To derive the boundary conditions on these functions, we must go to the boundary conditions on 
(   ) , X  Z  .   In  a straightforward way, we see that  (   ) 0, 0 Z    =   yields  ( ) 0 0 F   = ,  and  (   ) , 1 Z     =  
leads to  (   ) 1 F  = .  The remaining (inlet) condition gives 
 
(   )
  ( )
, 0 1
0
X
X   F 
|   |
=   =
   |
   |
\   .
 
 
By  choosing  ( ) 0 0    = ,  this  condition  collapses  into  the  condition  (   ) 1 F  =   obtained  already 
from  the  boundary  condition  on  the  scaled  temperature  field  as  X  .  Summarizing  the 
boundary conditions on  (  ) F   and  (   ) Z  , we have 
 
( )   (   ) 0 0, 1 F   F =    = , and  
( ) 0 0    =  
Integration yields the following solution for the scaled boundary layer thickness  (   ) Z  . 
 
(   )
1/ 3
9
2
Z   Z 
  |   |
=
   |
\   .
 
 
The solution for  (  ) F   is 
(  )
  (   )
3
3
3
0
0
0
1
4/ 3
e   d
F   e   d
e   d
=   =
 
 
Here,  ( ) x    represents  the  Gamma  function,  discussed  in  the  Handbook  of  Mathematical 
Functions by Abramowitz and Stegun.  The numerical value of  (   ) 4/ 3 1.120     . 
 
Heat Transfer Coefficient 
 
In the thermal entrance region, when the thermal boundary layer is thin, we can approximate the 
bulk average temperature 
b
T  by the temperature of the fluid entering the tube 
0
T .  Therefore, we 
define the heat transfer coefficient in this entrance region by 
 
(   )   (   )
0
,
w   w
T
q   k   R  z   h  T   T
r
=   =   
 
 
 
 
 
13 
Transforming  to  dimensionless  variables,  and  defining  a  Nusselt  number  2 / Nu   hR   k = ,  we 
can write 
 
(   )   (   )
  (   )
  ( )
2
2 0, 0
dF
Nu   Z   Z
X   Z   d
=   =
 
 
By  substituting  for  (   ) Z    and  ( ) 0
dF
d
,  we  obtain  the  following  approximate  result  for  the 
Nusselt number in the thermal entrance region. 
 
(   )
1/ 3
1/ 3
1.357
  R
Nu   Z   Pe
z
|   |
     |
\   .
 
 
Comparison with the exact solution shows this result is a good approximation in the range  
 
2500 50
Pe   z   Pe
R
|   |
   
   |
\   .
 
 
References 
 
M.  Abramowitz  and  I.  Stegun,  Handbook  of  Mathematical  Functions,  Dover,  New  York 
(1965). 
 
A. Acrivos, The extended Graetz problem at low Pclet numbers, Appl. Sci. Res. 36, 35 (1980). 
 
E.J. Davis, Can. J. Chem. Engg. 51, 562 (1973). 
 
L. Graetz, Ueber die Wrmeleitungsfhigkeit von Flssigkeiten, Annalen der Physik und Chemie 
18, 79 (1883). 
 
H.A. Lauwerier, Appl. Sci. Res. A2, 184 (1951). 
 
M.A.  Lvque,  Les  lois  de  la  transmission  de  chaleur  par  convection,  Annales  des  Mines, 
Memoires,  Series  12,  13,  201-299,  305-362,  381-415  (1928)  {as  cited  by  J.  Newman,  Trans. 
ASME J. Heat Transfer, 91, 177 (1969)} 
 
E. Papoutsakis, D. Ramkrishna, and H.C. Lim, The extended Graetz problem with Dirichlet wall 
boundary conditions, Appl. Sci. Res. 36, 13 (1980).