Sei sulla pagina 1di 37

MHMT14

Momentum Heat Mass Transfer

D
     source
Dt

Mass transfer

Fick´s law. Molecular mass transfer. Mass transfer


with chemical reactions. Unsteady mass transfer.
Convective mass transfer. Axial dispersion model.

Rudolf Žitný, Ústav procesní a


zpracovatelské techniky ČVUT FS 2010
MHMT14
Mass Transfer - diffusion
General transport equation for property P can be applied for the mass
transport too
P  
   (uP)    P  P( g )
t
In case of mixture of several components (we shall consider as an example
binary mixture consisting in components A and B) the transported properties
P are mass concentrations A of components proportional to mass fraction A
and density 
P   A  A 
Remark: components can be for example water vapour and air, or chemical
species, CH4, O2, N2,…

Mass flux of a component A is directly proportional to the driving potential –


gradient of concentration (or mass fraction), this is Fick’s law
r r
P  j A   DAB  A   DAB   A
{ { {
kgA diffusion gradient of
[ 2
] coefficient mass fraction
m s
[m2 / s ]
MHMT14
Mass Transfer – transport eqs.
Substituting into the general transport equation we obtain transport equations
for each component separately

 A r
   (u  A )    ( DAB A ) S
1t4 4 2 4 4 3
14 2 43 {A
diffusion flux. source term due to evaporation
compare with the continuity D AB is coefficient kgA
 r or chemical reaction [ 3 ]
of binary diffusion of A
equation  g( u  )  0 m s
t in mixture A+B [m2 / s ]

D  r
Using identity    g(u ) it is possible to write the transport equation as
Dt t

D A  A r
  (  u g A )    (  DAB A )  S A
Dt t
MHMT14
Mass Transfer – transport eqs.
So far everything seems to be as usual. For example exactly the same equation
(Fourier Kirchhoff) holds if you substitute T for the mass fraction A. However,
what does it mean the velocity u in the case of mixture? Individual components
are moving with different velocities and resulting mass flux (kg/(m2.s)) is the sum
of the component fluxes r r r r r r
u  {
 Au A   BuB  u   Au A  BuB
r r r
jA   A (u A  u )
r
n mass
A
flux of A

Especially at gases the concentrations are expressed in terms of molar


concentrations and molar fractions y A  cA / { c  pA / p
{
{ { kmol A+B
{
molar kmol A partial overall pressure
fraction m 3 pressure p A  pB
m3
r (m) r r r ( m) r r
cu  cAu A  cBuB  u  y Au A  yBuB
r
{ r r r ( m)
N A molar
flux of A J A  cA (u A  u )

Dy A y A r ( m)
c  c(  u{ gy A )    (cDABy A )  S A( m)
Dt t mean molar velocity
MHMT14
1D – steady diffusion
General transport equation for steady state, constant density and DAB, without
source term reduces to
r   2
A
u g A  DAB  A 
2
 uz A
 DAB
for 1D case
transport in z  z 2
z-direction

For gases it is more suitable to assume constant overall pressure and to use
molar fractions (yA) instead mass fractions
r (m) y  2
yA
u gy A  DAB y A 
2
 uz
( m) A
 DAB
for 1D case
transport in z  z 2
z-direction

A1 yA1 A2 yA2


z

L
MHMT14
1D – steady diffusion
Let us consider the case that u(m)=0, i.e. the same number of
molecules A is moving in one direction as the number of B molecules
in the opposite direction. uz( m)  0 z
This is so called equimolar diffusion and concentration profile is linear yA(z)

L
2 yA y A ( z )  y A1 z
0  DAB   yA1
z 2 y A2  y A1 L

Different case is for example evaporation of water vapors (component A) into air
(component B). Air cannot be absorbed in water and therefore mass or molar flux
is zero (uB=0) and mean molar velocity is determined by the velocity of vapors uA
N zA
uz( m )  uzA y A  (how to calculate the molar flux NzA will be shown in the next slide)
c
Equation of transport (steady state, without sources and unidirectional diffusion)
u (m) yA2
1  exp( z
z) u z( m)  0
y A  yA
2
y A ( z )  y A1 DAB
u z( m )  DAB   z yA(z) L
z z 2 y A2  y A1 uz( m )
1  exp( L)
DAB yA1
MHMT14
1D – steady diffusion
The mean molar velocity (convective velocity) for uB=0 and the resulting molar
flux NA follow from the definition of molar flux and the Fick´s law
J zA  N zA  c Au z( m )  N zA  y A N zA  N zA (1  y A )
y A
J zA  cDAB
z
cD y
N zA   AB A
1  y A z
Substituting for yA the previously calculated exponential yA profile gives
cDAB y A cDAB 1  y A2
N zA    ln
1  y A z L 1  y A1

and the molar fraction profile


1  y A ( z ) 1  y A2 z / L
( )
1  y A1 1  y A1
… and you see that this profile is also independent of DAB
MHMT14
Analogy heat and mass transfer
We could continue the lecture by unsteady diffusion, in a similar way and
with similar results as in the heat transfer. For example the principle of
penetration depth remains
  DAB t
and thus it is possible to estimate the range of concentration changes
corresponding to duration of a concentration disturbance (typical problem:
calculate the depth of soil filled by petrol spilled on surface at a specified time
after accident).
Problem of mass transfer from a surface to flowing fluid (convection) is also
solved in a similar way. It is only necessary to use appropriate variables

Heat transfer Mass transfer


T  Tf y A  y Af
temperature mass or molar fraction
Tw  T f y Aw  y Af

a [m2 /s] temperature diffusivity DAB [m2 /s] diffusion coefficient

q heat flux [W/m2 ] nA mass flux [kgA/(m2 .s)]

q   (Tw  Tf ) heat transfer coef. [W/(m2 K )] nA   (  Aw   Af ) mass transfer coef. [m / s]


MHMT14
Analogy heat and mass transfer
Analogical criteria

Heat transfer Mass transfer


D D
Nu  Nusselt number Sh  Sherwood number
 DAB
 
Pr  Prandtl number Sc  Schmidt number
a DAB
at DABt
Fo  Fourier number FoD  diffusional Fourier number
D2 D2
uD uD
Pe  Peclet number PeD  diffusional Peclet number
a DAB

Analogical correlations (valid for low concentrations, close to equimolar diffusion)

Heat transfer Mass transfer


Parallel flow
around a plate Nu  0.664 Re L Pr1/3 Sh  0.664 Re L Sc1/3

Flow around a
sphere Nu  2  (0.4 Re  0.06 Re2/3 ) Pr 0.4 Sh  2  (0.4 Re  0.06 Re2/3 )Sc0.4

Flow around
cylinder Nu  (0.4 Re  0.06 Re2/3 ) Pr 0.4 Sh  (0.4 Re  0.06 Re2/3 )Sc0.4
MHMT14
RTD – axial dispersion

Do you remember lecture 8 (RTD)?


The case of injection a tracer into a
stream of fluid flowing through an
apparatus was analyzed with the aim
to identify the RTD – Residence Time
Distribution of particles.

This is an example of transient


convective diffusion problem
(distribution of tracer concentration). Weyden
MHMT14
RTD – axial dispersion
The problem can be formulated as follows: a liquid flows in a long pipe with a fully
developed velocity profile, for example u(r )  2U (1  r 2 ) in the laminar regime,
where r is dimensionless radial coordinate and U is mean velocity. At inlet (x=0) a
small amount of tracer is injected at an infinitely short time. The injection should
simulate the situation when all fluid particles passing through x=0 are “labeled”
during a short time interval dt. These labeled molecules are component A and
their mean concentration cmA(t) is recorded by a detector located at a distance L
behind the injection point
cA c (t) is impulse response
A

t
L
The mean concentration in a cross-section cmA can be defined either as the area
average 1
cmA  2 c A (r )rdr
0 1


or the mass average cmA  2( u (r )cA (r )rdr ) / U
0
(both definitions are the same for the case
when velocity and/or concentrations are
uniform at cross section of pipe)
MHMT14
RTD – axial dispersion
Distribution of tracer concentration is described by the transport equation
cA cA 1  cA  2cA
 u (r )  DAB ( 2 (r ) 2 )
t x R r r r x this term, axial diffusion, is in fact negligibly
small when compared with the radial diffusion

The solution cA(r,x,t) obtained in the lecture 8 assumed zero diffusion (DAB=0),
therefore purely convective transport of tracer cA 2 c A
 2U (1  r ) 0
t x
5

giving impulse response


4

1
L2
cmA (t )   4(1  r )cA (r , L, t )rdr 
3
2

c
2 3
0
2U t 2

Remark: this definition of mean concentration ensures that


0
the recorded impulse response is identical with the residence 0 1 2
time distribution

Validity of this convective solution is restricted to very short times, so short that the
penetration depth of tracer is less than the radius of pipe D t  R AB
MHMT14
RTD – axial dispersion
Transport equation taking into account radial velocity profile is very difficult to
solve (see later). However, for example at turbulent flow regime the velocity
profile is almost uniform and the transport equation is simplified
Colour
cA cA  2cA
injection
Ut U  DAB
t x x 2
And this is called axial dispersion model ADM
x R c (x,t)
A

The parabolic PDE should be completed by boundary conditions


A) Open/Open problem cA0 for x, x- x

B) Closed/closed pipe of a finite length (DAB=0 for x<0 and x>L)


Remark: The closed end means that a labeled particle A once entering the inlet of L
pipe at x=0 cannot move back due to random migration and also a particle once
leaving the outlet cannot be returned back into the system (in our case pipe).

Initial condition at time t=0: There is Zero concentration everywhere with the
exception of origin where a unit amount of tracer was injected c A ( x, t  0)   ( x)
Dirac delta
function
MHMT14
RTD – axial dispersion
It is not necessary but useful to simplify the diffusion equation by using the
transformation to a convected coordinate system (,t) moving with fluid
cA  2cA
  x  Ut   DAB
t  2
This is a linear parabolic equation, exactly the same as the equation for unsteady
heat transfer (distribution of temperature in a plate). Only the boundary and initial
conditions are different. The analytical solution based upon infinite series of terms
Fi(t)Gi() is suitable for the case of bounded regions (e.g. a plate of finite
thickness), while in infinite regions (-,) integral transforms are preferred. In
our case we try to use the Laplace transform of time ts

f%( s)   f (t )e  st dt
0

which transforms the time derivative to multiplication by Laplace variable s


 
 c  c
L  A    A e st dt  [cAe st ]tt 
0  s  c e  st
dt  cA (0)  sc%A (s)
 t  0 t
A
0
MHMT14
RTD – axial dispersion
Application of Laplace transform to partial differential diffusion equation gives
 2c%A ( s,  )
cA (0,  )  sc%A ( s,  )  DAB
 2
and this is only an ordinary differential equation with respect spatial coordinate 
 2c%A sc%A  ( )
  
 2 DAB DAB
Solution of this equation for 0 (in this region the delta function is zero) is easy
s
c%A ( s,  )  A( s) exp( |  |)
DAB
This a continuous function of  with discontinuous first derivative at =0

c%A c% s 1
| 0  A | 0   2 A( s)  
  D D
1 4 2 4 3AB 1 4 2 AB4 3
derivative of general solution integral of right hand side (delta function)
MHMT14
RTD – axial dispersion
Thus determined coefficient A(s) completes the solution (expressed in the
Laplace domain), satisfying open/open boundary conditions and also initial
condition
1 DAB s
c%A ( s,  )  exp( |  |)
2 DAB s DAB
Next step must be the inverse Laplace transform usually performed by using
tables of Laplace transforms. In this reference you find 1  e  s  1 2
L   exp(  )

 s 
 t 4t

and this is almost our case, the only difference is scaling of the s-variable by a
constant DAB. There is simple rule for scaling L f ( D t )  1 f%( s )
 AB  (Prove!)
DAB DAB
Using the formula for inverse transform we obtain final result, concentration at
a distance x and time t

1 2
cA (t ,   x  Ut )  exp( )
4 DABt 4 DABt
MHMT14
RTD – axial dispersion
Comparison of impulse response of the convective model (parabolic velocity profile)
and the axial diffusion model (constant velocity U) for open/open case

DAB=0.05 m2/s 4

U=1 m/s cA
3 x2
L=1 m cmA (t ) 
2U 2t 3
Please, note the fact, that the
value of diffusion coefficient 2
0.05 is absolutely unrealistic, 1 ( x  Ut ) 2
even at gases the molecular cA  exp( )
diffusion is of the order 10-5. 4 DABt 4 DABt
1
Much greater values (e.g. 0.05)
are effective dispersion
coefficients (see later)
0
0 0,5 1 1,5 t 2

Gaussian concentration distribution along a pipe at a fixed time


1,4
t=1
cA 1,2
1 Width of concentration pulse
is very well characterised by
A small amount of tracer 0,8
the penetration depth
diffused before the injection 0,6
cross section (open/open case)
0,4    DABt  0.4
0,2
0
-0,5 0 0,5 1 1,5 2
x
MHMT14
RTD – axial dispersion
The closed/closed problem can be solved in a similar way, using Laplace transform
 2
L e  
and the inverse transform 1  s
4 t 3
exp(
4t
)

1 ( L  Ut ) 2
giving similar result cA (t ,   x  Ut )  exp( )
4 DABt 3 4 DABt
Both models of axial dispersion (open/open, closed/closed) are frequently used
in practice for modeling RTD in apparatuses like tubular reactors, packed beds,
extruders, fluidised beds, bubble columns and many others. However to match
experimental results it is necessary to use experimentally determined coefficient
De which is usually much greater than the molecular diffusion coefficient DAB
evaluated from tables or correlations. This is the same situation like with the
turbulent viscosity which is much greater than the molecular viscosity. And the
same reason: effective diffusion coefficient (called dispersion coefficient) is not a
material parameter and its value is affected by macroscopic motion - convection.
Laminar flow in pipe with nonuniform velocity profile is a good example: axial
dispersion is determined first of all by convection (by the parabolic velocity
profile). This problem was first solved by G.I.Taylor, see next slides…
MHMT14
Axial dispersion model ADM
G.I.Taylor (do you remember his analysis of large bubbles?) performed experiment
with injection of colour tracer into laminar flow in pipe.
Taylor G.I.: Dispersion of soluble matter in solvent flowing slowly through a tube. Proc.Roy.Soc. A, 219, pp.186-203 (1953)
Experimental setup consists in a long glass tube with small boring (alternatively 0.5 and 1 mm). Water flows inside the tube
very slowly (U1 mm/s) and thus perfect fully stabilised parabolic velocity profile exists in the whole tube. Diluted potassium
permanganate was used as a tracer and its concentration was evaluated visually, comparing colour in the test tube A with
color of prepared samples with precisely determined concentrations in the tube B. Flowrate was controlled by the valve N
and measured from the motion of meniscus it the tube T.

During flow an expanded blob of


tracer, moving with the mean liquid
velocity, was observed. When the
Capillary d=1 mm flow was stopped the expansion of
L=152 cm t=11000 s blob was stopped also. The axial
dispersion of tracer was observed
Distilled water Illuminated only during flow.
glass plate Theoretical explanation presented
by Taylor (1953, and 1954) gives
very surprising result:

Dispersion increases
Comparison tube with the decreasing
diffusion coefficient!!!
KMnO4 Meniscus velocity
MHMT14
ADM laminar/turbulent flow
Theoretical models for axial dispersion in a pipe developed by Taylor 1953 (very
slow laminar flow), 1954 (turbulent flow) can be summarized in this way
tracer injection
Ut

x R
cA(x,t)
R 2U 2
laminar De  [m2/s]
cA cA  cA 2
48DAB
u  De 2
t x x w
turbulent De  10.1Rv  10.1R
*


Example (corresponding to the Taylor’s experiment) R=0.0005 m, U=0.001 m/s, D AB=10-9 m2/s.
De=5e-6 (dispersion coefficient is 500 times greater than diffusion coefficient), Re=1 (laminar flow, stabilisation of
parabolic velocity profile almost immediately, at a distance from inlet less than 0.1 mm), minimum time corresponding
to equilibrations of radial concentration profile according to penetration theory 80 s (therefore axial dispersion model
can be used only for times longer than 80 seconds)..
MHMT14
ADM - restrictions
Axial dispersion model can be applied either at very high flowrates (at turbulent
flow regime) or at very small flowrates, when radial diffusion has got enough
time to equilibrate transversal concentration profile. There is a gap for
intermediate flowrates, where a numerical solution is still necessary.

DAB -diffusion coefficient, De-dispersion coefficient [m2/s]

Q [m3/s]
L [m]
De 
R 2U 2 w Q
De  10.1R
48DAB 

Q<12 DAB L Q>100 DAB L Q>3600 R


Taylor’s dispersion Hic sunt leones Purely convective transport Turbulent flow (R-radius, -kinematic viscosity)
(see next slide)

Very small flowrates Very large flowrates


MHMT14
ADM - restrictions
The following cA(r,x,t) profiles were calculated numerically for different values of
diffusion coefficient. Solution for DAB=1e-7 can be approximated quite well by
convective model, but the ADM is not a very good approximation in both cases.

2s 4s 6s 8s 10s
DAB =10-6m2/s

2s 4s 6s 8s 10s
DAB =10-7m2/s R

Pipe axis

cA
DAB =10-6 m2/s DAB =10-7 m2/s
L=1m, Numerical solution
R=0.01m,
U=0.1 m/s ADM (Taylor)

Numerical solution
ADM (Taylor)

t t
MHMT14 1886–1975
Diffraction-quant. m.(24 years old)
Motion of shocks (25 years old)
Instabilities T.C. (38 years old)

Statistical theory of turbulence


Drops and bubbles
...
Taylor dispersion (68 years old)
...
Classical Physics Through the Work
of GI Taylor MIT Course Electrohydrodynamics (83 years old)
MHMT14
ADM - derivation
Taylor (1953) demonstrated and experimentally verified that the transport
equation
cA c 1  cA  cA 2
 2U (1  r 2 ) A  DAB ( 2 (r ) 2 )
t x R r r r x
can be substituted by the axial dispersion equation (U-mean velocity)
cmA cmA  2cmA
U  De
t x x 2
where cmA is area average of concentration, and De is dispersion coefficient
1
R 2U 2
cmA  2 c A (r )rdr De 
0
48DAB
Taylor used the area average because it corresponded to the used experimental
technique (area averaged colour of tracer). The solution also assumed
impermeable wall, therefore zero concentration gradient at wall was applied as
a boundary condition.
MHMT14
ADM - derivation
In the following I shall try to follow the Taylor’s derivation with only slight changes:
Instead of the area average the mass average will be used. The reason why is in
the fact that only then it is possible to interpret the impulse response (response to
infinitely short injection of tracer) as the RTD = residence time distribution.
The second modification concerns the boundary condition at wall; instead of
impermeable wall (zero gradient) Newton’s boundary condition will be used. Why?
Exactly the same transport equation holds also for temperature and the only
difference is temperature diffusivity a replacing diffusion coefficient DAB. Similar
stimulus response experiments are carried out with a temperature marking instead
of the tracer injection (the temperature marking can be realised by short heating of
incoming fluid by an ohmic or dielectric heater, and response can be easily
recorded by thermocouples). However, in the case that the tube will not be perfectly
insulated, the dispersion of temperature T differs from the dispersion of a
component cA.
In the following equations the symbol for temperature T(r,x,t) will be used as an
alternative of concentration cA(r,x,t).
MHMT14
ADM - derivation
Transport equation for T (temperature or concentration) and boundary
condition at wall can be written in terms of dimensionless variables
T T 1  T
 2 Pe (1  r 2 )  (r )
 x r r r
T
|r 1  kT
r
where r-radius/R, x-distance/R. Pe is Peclet number defined as
UR UR
Pe  (for heat transfer) or Pe  (for mass transfer)
a DAB
 is dimensionless time (Fourier number)
at DABt
 (for heat transfer) or  (for mass transfer)
R2 R2
The coefficient k can be interpreted as the Biot number
R R
k (for heat transfer) or k (for mass transfer)
 DAB
where  represents heat transfer coefficient from the pipe wall to environment.
MHMT14
ADM - derivation
Transport equation can be transformed to the moving coordinate system
(moving right with the mean fluid velocity U)

T T 1  T (1)
 Pe(1  2r 2 )  (r )
  r r r
  x   Pe
Mass average Tm for parabolic velocity profile is the integral
1
Tm ( , )  4 r (1  r 2 )T (r ,  , )dr (2)
0

Approximation of solution of PDE is suggested in the form


Tm  2Tm
T (r ,  , )  Tm  h(r )  g (r ) 2 (3)
 
based upon assumption that a radial profile exists only if there are some
changes in the axial direction. The functions h(r) and g(r) are to be specified.
MHMT14
ADM - derivation
Substituting approximation (3) into (1)
Tm h' Tm g'  2Tm
 [ Pe(1  2r )   h '']
2
 [ Pe(1  2r )h   g ''] 2
2
(4)
 1 4 4 42 4 4 43r  1 4 4 4 4 2 4 4 r4 43 
this term can be eliminated by suitable De this term should be independent of r
choice of function h(r)

The function h(r) should be an even function h(r) =h1+h2r2+h3r4


Pe 2 r 4
h(r )  h1  (r  )
4 2
The coefficient h1 follows from definition of average temperature
Tm  2Tm
1 1 1 1

0 4r (1  r )Tdr  0 4r (1  r )drTm  0 4r (1  r )hdr   0 4r (1  r ) gdr  2


2 2 2 2

                      
Tm 1 ! !

0 0
1
Pe 2 r 4 Pe
0       
2
r (1 r )(h1 ( r )) dr 0 h1
4 2 16
MHMT14
ADM - derivation
In a similar way the function g(r) is derived

g (r )  A  Br 2  Cr 4  Dr 6  Er 8

The five coefficients A,B,C,D,E are selected so that the radial dependence
will be eliminated (3 equations for coefficients at r2, r4, r6),


then the normalisation condition 0  r (1  r 2 ) g (r )dr
0

and finally required boundary condition at wall:


T
|r 1  kT
r
Tm  2Tm Tm  2Tm
h '(1)  g '(1) 2  k (Tm  h(1)  g (1) 2 )
   
MHMT14
ADM - derivation
After the suggested (little bit tedious) manipulations the following transport
equation can be obtained
Tm 6k Pe Tm Pe2 40  11k  2Tm
 (Tm  )
 3 k 16  640 3  k  2
and returning back to the fixed coordinate system x     Pe

Tm 24  11k Tm 6k Pe2 40  11k  2Tm


 Pe  Tm  1

 24  8k x 3 k 640 3  k x 2 Tm ( , )  4 r (1  r 2 )T (r ,  , )dr
0

If we repeat the whole procedure but now using the area average Tm
Tm 12  4k Tm 8k Pe2 60  17k  2Tm
 Pe  Tm  1

 12  3k x 4k 12 60(4  k ) x 2 Tm ( , )  2 rT (r ,  , )dr


0

The both equations reduces to the ADM for insulated (impermeable) wall (k=0)
Tm Tm Pe2  2Tm
 Pe 
 x 48 x 2
and this is the result
obtained by Taylor
MHMT14
ADM - derivation

Frankly, I am not sure, if the derivation is correct –


isn’t it surprising that the ADM model with insulated
walls is the same when using area and mass
averaged concentrations (temperatures)?
MHMT14
Mass transfer - reactions

Diffusion plays a dominant role in chemical


reactions and combustion, because species
Hockney
react only in the case that they are sufficiently
mixed to a molecular level (micromixing).
MHMT14
Mass transfer - reactions
i mass fraction of specie i in mixture [kg of i]/[kg of mixture]
i mass concentration of specie [kg of i]/[m3]

Mass balance of species (for each specie one transport equation)


 r
( i )    ( i u )    (ii )  Si
t
Because only micromixed Rate of production
Production of species is controlled by reactants can react of specie i [kg/m3s]

Diffusion of reactants (micromixing) – tdiffusion (diffusion time constant)


Chemistry (rate equation for perfectly mixed reactants) – treaction (reaction constant)
tdiffusion
Damkohler number Da 
treaction
E a
Da<<1 Si   A exp( )i Reaction controlled by kinetics (Arrhenius)
RT

Da>>1 Si   Ci Turbulent diffusion controlled combustion
k
MHMT14
Species transport - Fluent
This is example of 2 pages in Fluent’s manual (Fluent is the most frequently used program for Computer Fluid Dynamics modelling)
MHMT14 EXAM

Mass transfer
MHMT14 What is important (at least for exam)
Transport equation, written either in concentrations (mass or molar) or in
fractions
D A  A r
  (  u g A )    (  DAB A )  S A
Dt t
Fick law
r
jA   DAB A   DAB A
r r r
jA   A (u A  u )
Penetration depth

  DAB t
MHMT14 What is important (at least for exam)
Analogy and corresponding dimensionless criteria

D 
Nu  Nusselt number Pr  Prandtl number
 a
D 
Sh  Sherwood number Sc  Schmidt number
DAB DAB

Axial dispersion model and relationship between diffusion and dispersion


coefficients
cA cA  2cA
U  De 2
t x x

Potrebbero piacerti anche