Você está na página 1de 12

MACROSCOPIC MASS AND ENERGY BALANCES

Michael Caracotsios, Ph.D.


Clinical Associate Professor
Chemical Engineering Department
University of Illinois at Chicago
1 Introduction

In the study of chemical engineering systems, we find it necessary to write conservation laws for
“macroscopic systems” that include large pieces of equipment or parts thereof, like chemical
reactors, pumps, compressors, flash drums, distillation columns, pipes and other. The
conservation laws allow us to monitor the behavior of these systems and take corrective action if
necessary. They also allows us to size the equipment and to optimize the process behavior in
order to minimize costs of production or maximize revenues. In this monograph we will present
the generic form of these macroscopic balances for a general system depicted in Figure 1.
below. The conservation laws for mass, momentum, energy and mechanical energy will be
written and some examples will be presented to illustrate their use for practical situations.

Figure 1. A stationary open system and its surroundings

Michael Caracotsios Page 1


The general form of the conservations laws for mass, momentum and energy can be written in
qualitative terms as follows:

(ACCUMULATION)=(INPUT)-(OUTPUT)+(GENERATION)

Although this universal equation describes the conservation of a physical quantity, very often we
find it necessary to include various constraints that arise from numerical considerations, as well
as from thermodynamic principles. The most common constraints include:

1. Thermodynamic equilibrium constraints if multiple phases are present in the mixture.


2. The second law of thermodynamics or equivalently δ Q = TdS for systems undergoing a
reversible process.
3. Entropy and lost work macroscopic balances for irreversible processes.
4. Summation constraints if the state variables are the molar or mass fractions.
5. Numerical constraints that impose physical bounds on certain unknown variables, such as
non-negative flows, mole[mass] fractions in the range [0,1] and other similar constraints.

We are now going to describe in some detail the macroscopic balances for mass, momentum,
mechanical energy and total energy. In all balances we assume that the property profiles in the
direction perpendicular to the flow are flat so that p = p where p indicates the average
value of a property p , like fluid density, velocity, enthalpy etc.

2 Macroscopic Mass Balance

The macroscopic mass balance equation for a component i can be written as follows:

dM i
dt
= ( ρ υ A ) − ( ρ υ A ) + ( ℜ V =)
i
1 1 1
i
2 2 2
i
tot w1i − w2i + ℜiVtot

whereas the macroscopic total mass balance is given by the equation:

dM tot
( ρ1υ1 A1 ) − ( ρ2υ2 A2 ) =
= w1 − w2
dt

The following definitions and units apply to the symbols used in the above equation:

Michael Caracotsios Page 2


Symbol Meaning Units
Mi Mass of component i in the System kg
kg
ρ1i , ρ 2i Component i density at planes 1 and 2
m3
w1i , w2i Component i mass rate at planes 1 and 2
M tot Total Mass of the System kg
t Time s
kg
ρ1 , ρ 2 Fluid density at planes 1 and 2
m3
m
υ1 ,υ2 Fluid velocity at planes 1 and 2
s
A1 , A2 Flow cross section area at planes 1 and 2 m2
kg
ℜi Component i generation term (for Reactive systems)
m3 ⋅ s
Vtot Total Volume of the System m3
kg
w1 , w2 Fluid mass rate at planes 1 and 2
s

Under steady-state conditions and if no reactions are present the macroscopic component and
total mass balance reduce to the following simple equations:

ρ1υ1 A1 − ρ 2υ2 A2 =w1 − w2 =0 ⇒ w1 =w2 =w


ρ1iυ1 A1 − ρ 2iυ2 A2 =w1i − w2i =0 ⇒ w1i =w2i =wi

3 Macroscopic Momentum Balance

Recognizing the υ represents momentum per unit mass the steady-state macroscopic momentum
balance is given by the following equation:

d
( M υ=) (υ1 ρ1υ1 A1 + P1 A1 ) − (υ2 ρ 2υ2 A2 + P2 A2 ) + ( M tot g + F )
dt

where the symbol definitions presented above apply also here and in addition we have the
following

Symbol Meaning Units


g Gravity acceleration m / s2
P1 , P2 Pa [
Fluid pressure at planes 1 and 2 = ] kg / ( m ⋅ s 2 )
F Extrenal force acting on the fluid =[
N ] kg ⋅ m / s 2

Michael Caracotsios Page 3


4 Macroscopic Energy Balance

The total energy of the systems consists of the sum of its internal energy plus kinetic energy plus
potential energy. The unsteady-state macroscopic energy balance is described by the following
equation:

d  1 2 P1    1 2 P2  
(U tot + K tot + Φ tot ) =
 U1 + υ1 + gh1 +  ρ1υ1 A1  −  U 2 + υ2 + gh2 +  ρ 2υ2 A2  + Q + Ws
dt  2 ρ1    2 ρ2  
For stationary systems the kinetic and potential terms are not present and therefore the
macroscopic energy balance reduces to the following equation:

dU tot   1 P   1 P  
=  U1 + υ12 + gh1 + 1  ρ1υ1 A1  −  U 2 + υ22 + gh2 + 2  ρ 2υ2 A2  + Q + Ws
dt  2 ρ1    2 ρ2  

This equation for a closed system with no input and output reduces to

dU tot    + Wdt
 ⇒ dU = δ Q + δ W ⇔ ∆U = Q + W
= Q + W ⇒ dU = Qdt
dt

The last equation of course is the mathematical rendition of the first law of thermodynamics.
P
Recognizing now that U + = H the macroscopic energy balance can be written as:
ρ

dU tot   1    1  
=   H1 + υ12 + gh1  ρ1υ1 A1  −   H 2 + υ22 + gh2  ρ 2υ2 A2  + Q + Ws
dt  2    2  

Usually the kinetic and potential terms are small compared to the enthalpy and internal energy
terms, so the above equation can further be simplified to:

([ H ] ρ υ A ) − ([ H ] ρ υ A ) + Q + W
dU tot
= 1 1 1 1 2 2 2 2 s
dt

Under steady-state conditions there is a further simplification that leads to:

∆H = H 2 − H1 ⇒ ∆H = Q + Ws

which is similar to the first law of thermodynamics but for open systems.

In the above equations we have made use of the mathematical relation: δ Q = Qdt  . The Greek
symbol δ is used for the heat and work differentials, instead of the standard differential symbol

Michael Caracotsios Page 4


d, since δ Q is not a true differential. As an example, and to clarify the two differential symbols
consider a stationary system that undergoes a change from a state 1 to a state 2. Then the
following mathematical relations hold for the internal energy and the heat input or output to or
from the system:

2 2

∫ dU =
1
∆U and ∫ δ Q =
Q
1

The symbol definitions presented in the mass, momentum and mechanical energy balances apply
also here and in addition we have the following:

Symbol Meaning Units


U tot Fluid total Internal Energy J
K tot Fluid total Kinetic Energy J
Φ tot Fluid total Potential Energy J
J
H1 , H 2 Fluid specific Enthalpy at planes 1 and 2
kg
J
Q Heat input into the system
s
J
Ws Shaft work done on the fluid
s
J
W Compression work on the fluid
s
J
Q Heat input into the system per unit mass
kg
J
Ws Shaft work done on the fluid per unit mass
kg
J
W Compression work on the fluid per unit mass
kg

Michael Caracotsios Page 5


5 Macroscopic Mechanical Energy Balance

The mechanical energy of a system is the sum of its kinetic and potential energy. The mechanical
energy balance is not a fundamental principle.. It is rather a corollary of the equation of motion.
1
Recognizing the υ 2 represents kinetic energy per unit mass and gh represents potential energy
2
per unit mass the macroscopic mechanical energy balance for constant fluid density is given by
the equation:

d 1 2 P1  1 2 P2 
( K tot + Φ=
tot )  υ1 + gh1 +  ρ1υ1 A1 −  υ2 + gh2 +  ρ 2υ2 A2 − Eυ +=
Ws 0
dt  2 ρ1   2 ρ 2 

Applying the steady-state macroscopic mass balance to the above momentum equation we obtain
under steady-state conditions:

1 2  1 2 
2
dP
 υ1 + gh1  −  υ2 + gh2  − ∫ − Eυ + Ws =
0
2  2  1 ρ

where the symbol definitions presented above apply also here and in addition we have the
following:

Symbol Meaning Units


h1 , h2 Fluid inlet and outlet heights from a reference plane m
J
Eυ Friction loss
s
J
Ws Shaft work done on the fluid
s
J
Eυ Friction loss per unit mass of fluid
kg
J
Ws Shaft work done on the fluid per unit mass
kg

The friction loss Eυ consists of two parts: (a) the loss in all sections of straight conduits and (b)
the loss due to fittings(round and square elbows), valves, entrance, contraction and expansion
effects. The following formula gives the friction loss in a piping system:

Michael Caracotsios Page 6


1 2 L 1 A
Eυ =υ f + υ 2 eυ where Rh =
2 Rh 2 Z
Rh hydraulic radius
A cross section area of conduit
Z wetted perimeter
L pipe length
f Fanning friction factor
Disturbance eυ
Entance to pipe 0.05
Small area
Sudden contraction 0.45 (1 − β ) β =
Large area
2
1 
Sudden expansion  − 1 If β ≈ 0 then eυ =
1
β 
90 rounded elbow 0.4 − 0.9
0

900 square elbow 1.3 − 1.9


0
45 elbow 0.3 − 0.4
Globe valve (open) 6 − 10
Gate valve (open) 0.2

The friction factor can be calculated from the Colebrook equation which is given by the formula:

1  1.256 k  ρυ D
+ 1.737 ln  + =0 Re = > 2100
f  Re f 3.707 D  µ

An explicit formula for the friction factor is given by the Blasius equation:

ρυ D
=f 0.0791Re −0.25 =
Re > 2100
µ

An approximation to Colebrook’s equation is given by the Haaland explicit equation:

1  6.9 k  ρυ D
=
−3.6 log10  +  Re = > 400 0
f  Re 3.707 D  µ

where in the above equations k is the tube roughness, D is the tube diameter and Re is the
Reynolds number. For laminar flow ( Re < 2100 ) the friction factor is given by the equation:

Michael Caracotsios Page 7


16
f =
Re

For packed tubes the friction factor is given by the Ergun equation:

1− ε  1− ε  ρυ d p
f = 3  75.0 + 0.875  Re = υ [=] superficial velocity
ε  Re  µ

where in the above equation ε is the porosity of the tube packing, and d p is the particle
diameter. Remember that the Darcy friction factor is four times the Fanning friction factor.

6 Macroscopic Entropy Balance

The derivation of the macroscopic entropy balance is similar to the one for the energy balance,
however one needs to consider the entropy of the system and that of the surroundings.
Recognizing that there is a generation term, the macroscopic entropy balance can be written as
follows:

Q 
([ S1 ] ρ1υ1 A1 ) − ([ S2 ] ρ2υ2 A2 ) +
dS tot
= + Sgen
dt To

Under steady-state conditions the above equation reduces to:

Q 
∆S = S 2 − S1 ⇒ w ∆S = + Sgen where w = ρ1υ1 A1 = ρ 2υ2 A2
To

Dividing the left and right hand side of the above equation by the mass flow rate we obtain an
expression for the entropy generation for irreversible processes:

Q Q
∆S = + Sgen ⇒ Sgen =∆S −
To To

And the lost work due to the irreversibility of the process can be calculated from the formula:

Wlost= T0 ⋅ Sgen

For reversible process, since the entropy generation is equal to zero, we obtain the second law of
thermodynamics:

Q δQ
∆S = S 2 − S1 ⇒ ∆S = and in differential form dS =
To To

Michael Caracotsios Page 8


Furthermore for reversible adiabatic processes the change in entropy is equal to zero dS = 0 .

The symbol definitions presented in the mass, momentum, mechanical and total energy balances
apply also here and in addition we have the following:

Symbol Meaning Units


J
S tot Fluid total Entropy
K
J
S1 , S 2 Fluid specific Entropy at planes 1 and 2
K ⋅ kg
T0 Bath Temperature K
J
Sgen Rate of Fluid Entropy Generation
K ⋅s
J
Sgen Fluid Specific Entropy Generation
K ⋅ kg

Note: Specific thermodynamic properties, such as specific enthalpy, entropy, internal energy and
volume are often given per mole (mol) of substance instead of per mass (kg) of substance.

Note: SI Fundamental Units

7 Macroscopic Lost Work(LW) Balance

The derivation of the macroscopic Lost Work (LW) balance is similar to the one for the energy
balance. We first define the concept of the Exergy Function (equivalent to available work) Ek
which applies to a process stream k as E= k H k − T0 S k where T0 is known as bath or sink
temperature (usually set at ambient conditions or approximately 300 o K ). Then the steady-state
Lost Work balance is given by:

Michael Caracotsios Page 9


   T 
LW = −∆  ∑ mk Ek  + ∑ Ws + ∑ 1 − 0 Q
 k   T 

where the summation indicates all possible source of shaft (mechanical) work done on the
system and all possible heat loads on the system. The Lost Work is usually an indication of the
overall efficiency of the process in considerations. Minimization of the lost work is a continuous
task and part of the discipline of process intensification. For a closed system with no mechanical
work done on it the above equation reduces to the well know Carnot efficiency formula.

8 Differential Forms of Macroscopic Balances

When the properties of a system change continuously differential forms of the macroscopic
balances must be used. Consider the flow of a compressible fluid in a horizontal pipe. The
pressure, temperature and specific volume of the fluid change continuously along the length of
the pipe. In this particular case the macroscopic balances must be written in a differential form as
1
indicated below. In the equations below we make use of the formulas ρ = and
V
ρυ = G ⇒ υ = GV where ρ is the compressible fluid mass density, G is the mass velocity,
which is constant under steady-conditions and constant cross section area for flow, V is the
mass specific volume of the fluid, and υ is the linear velocity of the fluid.

Macroscopic Mass Balance:

dυ dV
υ = GV ⇒ =G or dυ = GdV
dz dz

Macroscopic Energy Balance:

dT dυ dT dV
dH + υ dυ =
qdz ⇒ C p +υ =⇒
q Cp + G 2V =
q
dz dz dz dz

where q is the linear heat flux per unit length of pipe, per unit mass of fluid flowing through the
pipe and C p is the heat mass capacity of the fluid.

Mechanical Energy Balance

dP1 4f dV dP 2f
υ dυ + + υ2 dz =0 ⇒ G 2V +V + G 2V 2 =0
ρ 2 D dz dz D

Michael Caracotsios Page 10


For incompressible fluids the density is constant and the Mechanical Energy Balance equation
becomes after some algebraic modifications:

∆P 1 4f
− = ρυ 2
L 2 D

For compressible fluids the above equations must be combined with an equation of state in order
to create a well-defined system of coupled differential equations. If we assume ideal gas
behavior then we have PVM w = RT ,where M w is the molecular weight of the fluid, and
therefore we can replace the temperature derivative with:

dT PM w dV VM w dP
= +
dz R dz R dz

Combining the ideal gas law with the differential forms of the macroscopic energy and
mechanical energy balances, the equations that describe flow of a compressible ideal gas fluid in
linear pipe can be written as:
dV dP 2f
G2 + + G 2V = 0
dz dz D
 2 PM w  dV VM w dP
 G V + Cp  + Cp =q
 R  dz R dz

These two coupled differential equations can then be solved simultaneously using available
commercial software, like Athena Visual Studio.

An alternative method is to start with the macroscopic mechanical energy balance and replace
T +T
the specific volume with a value calculated at the average temperature Ta = 1 2 . Then we
2
have:

dV dP 2f dV PM w 2f
G2 + + G2 dz =
0 ⇒ G2 + dP + G 2 dz =
0
V Va D V RTa D

Integrating the last equation between the inlet and the outlet of the pipe and assuming that the
Reynolds number remains constant (i.e. the viscosity of the fluid does not change appreciably
over the length of the pipe) we obtain:

G 2 ln
V2
+
Mw
V1 2 RTa
( P22 − P12 ) + G 2
2 Lf
D
=
0

Applying the ideal gas law the above equation becomes:

Michael Caracotsios Page 11


G 2 ln
PT
1 2
+
Mw
P2T1 2 RTa
( P22 − P12 ) + G 2
2 Lf
D
=
0

This last equation can be solved for the outlet pressure using a numerical method, like secant, or
Newton or Goalseek in Microsoft Excel. Once the outlet pressure is calculated, one can also
calculate the specific mass volume as well as the outlet velocity of the fluid. An easier equation
to solve, can be derived by introducing a new variable x = P2 / P1 . Then the equation for the
pressure drop becomes:

 T  M P2
−G 2 ln  x 1  + w 1 ( x 2 − 1) + G 2
2 Lf
=
0
 T2  2 RTa D

We also know that the unknown variable lies in the open interval ( 0,1) because the pressure
always drops and therefore we can easily come up with an initial guess like ( x0 = 0.5 ) for the
solution of the above equation.

Michael Caracotsios Page 12

Você também pode gostar