You are on page 1of 29

Chapter 4: Modeling Fluid and Thermal Systems

4.1 The tank has the pressure-mass relationship: P  1.013(10) 5  3.6m (in Pa)

Fluid capacitance is C  dV / dP .

Fluid mass is m  V so taking a differential yields dm  dV (incompressible fluid)

Taking differentials of the pressure equation we obtain dP  3.6dm  3.6 dV

Solving for C  dV / dP yields dV / dP  1 /(3.6  ) , or C  2.778(104 ) m3/Pa

Copyright © 2016 John Wiley & Sons


Chapter 4

4.2 For laminar flow we have the linear flow equation: P  RLQ

For turbulent flow we have the nonlinear flow equation: P  RT Q 2

Using the hydrostatic equation, tank pressure is P  Patm  gh . The pressure difference is

P  P  Patm  gh

Assuming laminar flow, we can use the data points to solve for laminar flow resistance RL

Data point #1: RL  gh / Q  (1000 kg/m3)(9.81 m/s2)(4.1 m)/8.9689E-4 m3/s = 4.4845(107) Pa-s/m3

Data point #2: RL  gh / Q  (1000 kg/m3)(9.81 m/s2)(3.24 m)/7.9730E-4 m3/s = 3.9865(107) Pa-s/m3

Therefore, we do not obtain the same laminar flow resistance RL  flow Q vs. h is not linear

Assuming turbulent flow, we can use the data points to solve for turbulent flow resistance RT

Data point #1: RT  gh / Q 2  5(1010) Pa-s2/m6


Data point #2: RT  gh / Q 2  5(1010) Pa-s2/m6 … etc

A single turbulent flow resistance RT holds for all data points. Hence the valve flow is turbulent.

(we could also plot flow Q vs. liquid height h and observe that the plot is nonlinear  turbulent )

2
Copyright © 2016 John Wiley & Sons
Chapter 4

4.3 Because we have two tanks (or, two fluid capacitances C1 and C2) we require two first-order
ODEs for the pressures P1 and P2

Tank 1: C1 P1  Qin  Q1 Tank 2: C2 P2  Q1  Q2

a) For laminar valve flow we can write Q  P / R for each valve

P1  Patm P2  Patm
Valve 1: Q1  Valve 2: Q2 
R1 R2

Substituting the flow rates Q1 and Q2 and C1 = A1/g and C2 = A2/g into the two first-order
ODEs yields (after some algebra)

A1  1 1
P1  P1  Qin  Patm
g R1 R1

A2  1 1  1 1
P2  P2  P1     Patm Mathematical model (part a)
g R2 R1  R2 R1 

b) Using the hydrostatic equation, tank pressures are P1  Patm  gh1 and P2  Patm  gh2 .
Taking derivatives, we obtain P1  gh1 and P2  gh2 . Substitute these expressions for P1, P2,
P1 , and P2 into the modeling equation derived in part a:

gh1  Patm  gh1   Qin  Patm


A1 1 1
Tank 1:
g R1 R1

 1 1
gh2  Patm  gh2   Patm  gh1      Patm
A2 1 1
Tank 2:
g R2 R1  R2 R1 

The terms involving atmospheric pressure Patm cancel and the final result is

g
A1h1  h1  Qin
R1

g g
A2 h2  h2  h1  0 Mathematical model (part b)
R2 R1

3
Copyright © 2016 John Wiley & Sons
Chapter 4

4.4 Because we have two tanks (or, two fluid capacitances C1 and C2) we require two first-order
ODEs for the pressures P1 and P2 :

Tank 1: C1 P1  Q1  Q2 Tank 2: C2 P2  Q2  Qout

where Q1 is flow through valve R1, Q2 is flow through valve R2, and Qout is flow through valve R3

The pump raises the water pressure from Patm (sump or reservoir) to Pin = gH + Patm, where Pin is
the absolute (input) pressure upstream of valve R1

The three laminar valve flows are

Pin  P1 P1  P2 P2  Patm
Valve 1: Q1  Valve 2: Q2  Valve 3: Q3 
R1 R2 R3

Substituting for the flow rates (and Pin = gH + Patm ) and capacitances C1 = A1/g and C2 = A2/g
into the first-order ODEs yields the model

A1   1 1 
P2  gH  Patm 
1 1
P1     P1 
g  R1 R2  R2 R1

A2   1 1  1 1
P2     P2  P1  Patm Mathematical model
g  R2 R3  R2 R3

4
Copyright © 2016 John Wiley & Sons
Chapter 4

4.5 Because we have one tank (single capacitance) we start with the single ODE

CP  Qin  Qout where C = A/g

The centrifugal pump takes liquid at atmospheric pressure Patm and increases the pressure by
amount P . Therefore the fluid pressure immediately upstream of valve R1 is Pin = Patm + a 2

a) For laminar flow (both valves) we have the linear flow equations (with R1 = RL)

Pin  Patm a 2 P  Patm


Valve 1: Qin   Valve 2: Qout 
R1 RL R2

Substituting the above flow equations into the single first-order ODE yields

A  1 a 2 1
P P  Patm Model for laminar valve R1 = RL
g R2 RL R2

b) For turbulent flow in valve 1 (laminar flow in valve 2) we have

Pin  Patm a 2 P  Patm


Valve 1: Qin   Valve 2: Qout  (same as a)
RT RT R2

Substituting the flow equations into the single first-order ODE yields

A  1 a 2 1
P P  Patm Model for turbulent valve 1
g R2 RT R2

5
Copyright © 2016 John Wiley & Sons
Chapter 4

4.6 The fundamental modeling equation for a pneumatic system with laminar flow is

P V
CP  win  where capacitance C  (rigid vessel)
R nRT

Volume V is clearly constant for the two processes, as is supply pressure PS, temperature T, and
isothermal expansion process (n = 1). However, gas constant R is different for the two processes
(air vs. H2) and therefore capacitance C is not the same for the two processes. Hence, the
modeling equations have different values of C and the pressure responses P(t) will be different.

6
Copyright © 2016 John Wiley & Sons
Chapter 4

4.7 The fundamental pressure-rate ODE for a compressible hydraulic fluid is


P  Qin  V 
V

Volume V is constant so V  0 .


a) For constant volume, the basic pressure-rate equation becomes P  Qin , or CP  Qin
V
V
Hence fluid capacitance is C  = 0.006 m3/1.2(109) Pa = 5(10-12) m3/Pa

b) The pressure rate at this instant is


P  Qin = [1.2E9 Pa][2.4E-6 m3/s]/0.006 m3 = 480,000 Pa/s = 480 kPa/s
V

7
Copyright © 2016 John Wiley & Sons
Chapter 4

4.8 The fundamental pressure-rate ODE for a compressible pneumatic fluid is

nRT  P 
P   win  V
V  RT 

Volume V is constant so V  0 .

a) For a rigid vessel (constant volume) the fluid capacitance for a pneumatic system is

V
C
nRT

where n = 1 (isothermal), R = 287 N-m/kg-K, and T = 298 K. Hence capacitance is

C = 7.015(10-8) kg/Pa

b) The pressure rate at this instant is

nRT
P  win = [0.0035 kg/s]/[7.015E-8 kg/Pa] = 49,890 Pa/s = 49.89 kPa/s
V

8
Copyright © 2016 John Wiley & Sons
Chapter 4

4.9 The fundamental pressure-rate ODE for a hydraulic system with compressible flow is


P  Qin  V  (A)
V

First, consider the free-body diagram of the mechanical component (accumulator piston)

PA
kx
Patm A

Summing forces (positive is to the right), we obtain PA  kx  Patm A  mx  0 (no inertia)

Or, the spring force balances the two pressure forces: ( P  Patm ) A  kx

Taking a derivative of the above equation yields P A  kx , or, solving for velocity x  P A / k

Because the accumulator volume is V  V0  Ax , the time-rate of volume change is V  Ax

Therefore, substituting the velocity from the mechanical force balance, we obtain V  P A2 / k

Finally, we can substitute the above result for V in Eq. (A):

 P A2   V A2  
P   Qin   or,    P  Qin
V k   k 

V A2
The bracketed term is the total capacitance: C  Accumulator capacitance (m3/Pa)
 k
(hydraulic system)

9
Copyright © 2016 John Wiley & Sons
Chapter 4

4.10 The fundamental pressure-rate ODE for a pneumatic system with compressible flow is

nRT  P 
P   win  V (A)
V  RT 

First, consider the free-body diagram of the mechanical component (accumulator piston)

PA
kx
Patm A

Summing forces (positive is to the right), we obtain PA  kx  Patm A  mx  0 (no inertia)

Or, the spring force balances the two pressure forces: ( P  Patm ) A  kx

Taking a derivative of the above equation yields P A  kx , or, solving for velocity x  P A / k

Because the accumulator volume is V  V0  Ax , the time-rate of volume change is V  Ax

Therefore, substituting the velocity from the mechanical force balance, we obtain V  P A2 / k

Finally, we can substitute the above result for V in Eq. (A):

nRT  P P A2   V PA2  
P   win   or,    P  win
V  RT k   nRT RTk 

V PA2
The bracketed term is total capacitance: C  Accumulator capacitance (kg/Pa)
nRT RTk
(pneumatic system)

10
Copyright © 2016 John Wiley & Sons
Chapter 4

4.11 We can equate the Darcy equation with the basic laminar flow relation:

1 2L
P  v f  RLQ
2 d

For pipe flow, volumetric flow-rate is Q = Av = (d/2)2v . Substituting this expression for Q in
the above equation and solving for laminar resistance RL we obtain

L
RL  2 v f
d 3

Substitute for the Darcy friction factor, f = 64/Re and use the definition of Reynolds number Re

vd
Re 

L 64 128L
RL  2 v or RL  Hagen-Poiseuille law (Eq. 4.3)
d 3 vd d 4

11
Copyright © 2016 John Wiley & Sons
Chapter 4

1 2L
4.12 The Darcy equation is P  v f
2 d

For turbulent flow, pressure drop is P  RT Q 2 .

We know that the volumetric flow rate is Q  Av   (d / 2) 2 v . Therefore, substituting this


expression for Q in the turbulent-flow pressure-drop equation yields P  RT  2 (d / 2) 4 v 2 .
Equating the Darcy equation with the general turbulent-flow equation we obtain

4
1 L d 
P  v 2 f  RT  2   v 2
2 d 2

Solving for the turbulent-flow resistance yields

L
RT  8  f = 6.0159(108) Pa-s2/m6
d 2
5

12
Copyright © 2016 John Wiley & Sons
Chapter 4

4.13 Laminar flow rate is linearly proportional to pressure drop: Q  P / RL .

Laminar-flow fluid resistance RL can be computed from any data trials in Table P4.13. Using
the measurements from the first trial:

P1
RL   7.2 Pa/6E-5 m3/s = 120,000 Pa-s/m3
Q1

The other trials yield the same result for RL.

13
Copyright © 2016 John Wiley & Sons
Chapter 4

4.14 The manufacturer’s pressure-flow relationship is P  2.1646(1011 )Q 2 (Pa)

a) The basic hydraulic orifice equation for valve flow is

2
Q  Cd A0 P

Squaring both sides, solving for P, and equating with the manufacturer’s equation we obtain

Q2
2 2
 P  2.1646(1011 )Q 2
2Cd A0

Solving the above expression for area A0 yields (after substituting numerical values for  and Cd)


A0  = 7.2055(10-5) m2
(2)( 2.1646 E11)Cd2

b) The MATLAB M-file for plotting Q vs. P is below, along with the plot (next page):

% Problem 4.14b
%
% constants
rho = 864; % density, kg/m^3
A0 = 7.2055e-005; % orifice area, m^2
Cd = 0.62; % discharge coefficient

DelP = linspace(0,1.4e6,500);
for i=1:500;
DP = DelP(i);
Q1(i) = sqrt(DP/2.1646e11);
Q_full(i) = Cd*A0*sqrt(2*DP/rho);
Q_half(i) = Cd*(A0/2)*sqrt(2*DP/rho);
Q_quart(i) = Cd*(A0/4)*sqrt(2*DP/rho);
end

plot(DelP,Q_full,DelP,Q_half,DelP,Q_quart)
grid
xlabel('Pressure drop, \DeltaP, Pa')
ylabel('Volumetric flow-rate, Q, m^3/s')

14
Copyright © 2016 John Wiley & Sons
Chapter 4

-3
x 10
3

fully open valve


2.5

Volumetric flow-rate, Q, m3/s


half open valve
1/4 open valve
2

1.5

0.5

0
0 2 4 6 8 10 12 14
Pressure drop, P, Pa x 10
5

15
Copyright © 2016 John Wiley & Sons
Chapter 4

4.15 a) For an incompressible fluid the in/out flow to the chamber is Q  Ax ; in other words
the volumetric fluid flow-rate Q is equal to the change in volume. Therefore, the pressure-
rate equation for either chamber of the hydraulic system is


P  Q  V  = 0  No pressure change in either chamber
V

The orifice volume-flow rate for x  0 (i.e., piston moves to the right) is

Q  Cd A0
2
P2  P1   Ax (fluid flow from chamber 2 to 1)

Squaring both sides and solving for the pressure difference yields

A2 x 2  A2 x 2
P2  P1   0 if x  0 and P2  P1   0 if x  0
2Cd2 A02 2Cd2 A02

Or, we can use the following pressure-difference equation that accounts for x  0 and x  0

A2
P2  P1  x x (A)
2Cd2 A02

Next, consider the free-body diagram (FBD) of the mechanical component (piston)

P1A P2A
Fa
bx

Summing forces (positive is to the right), we obtain P1 A  Fa  P2 A  bx  mx

Finally, substitute Eq. (A) for the pressure difference in the above mechanical model:

A3
mx  bx  x x  Fa Mathematical model of hydraulic damper
2Cd2 A02

16
Copyright © 2016 John Wiley & Sons
Chapter 4

b) For compressible flow we must use the pressure time-rate ODEs

 
Chamber 1: P1  Q  Ax  Chamber 2: P2   Q  Ax 
V1 V2

where the chamber volumes are V1  V0  Ax and V2  V0  Ax . The above pressure-rate


equations assume that Q > 0 is flow from chamber 2 to chamber 1, or P2  P1  0 .

We can use the following general volumetric flow-rate equation which holds for P2  P1  0
or P2  P1  0 :
2
Q  Cd A0 sgn( P2  P1 ) P2  P1

Finally, substitute the above expression for Q and the expressions for V1 and V2 in the
pressure-rate ODEs. Also, include the mechanical model of the piston (see FBD in part a):

  2 
P1   Cd A0 sgn( P2  P1 ) P  P  Ax 
V0  Ax   2 1 

  2 
P2    Cd A0 sgn( P2  P1 ) P2  P1  Ax  Mathematical model

V0  Ax   

mx  bx  ( P1  P2 ) A  Fa

17
Copyright © 2016 John Wiley & Sons
Chapter 4

4.16 The chamber volumes are V1  V0  Ax and V2  V0  Ax , and the volumetric rates
are V1  Ax and V2   Ax .

The complete mathematical model of the pneumatic piston damper consists of two pressure
time-rate ODEs and the mechanical model (see related steps in Problem 4.15)

nRT  P 
P1   w  1 Ax 
V0  Ax  RT 

nRT  P 
P2    w  2 Ax  Mathematical model
V0  Ax  RT 

mx  bx  ( P1  P2 ) A  Fa

where the pneumatic orifice mass-flow rate is either choked or unchoked, depending on the
pressure ratio between the chambers:

 1
 2

2    
 lo    lo  
P P  
w  Cd A0 Phi sgn( P2  P1 ) if Plo / Phi  Cr (unchoked)
(  1) RT  Phi   Phi  
 

 1
 
w  Cd A0 Phi sgn( P2  P1 ) Cr if Plo / Phi  Cr (choked)
RT

The “high” pressure is Phi = max(P1,P2) and the “low” pressure is Plo = min(P1,P2).

Note that w > 0 if P2 > P1 (flow from chamber 2 to 1) and w < 0 if P2 < P1

18
Copyright © 2016 John Wiley & Sons
Chapter 4

4.17 The fundamental first-order ODE for a hydraulic system with compressible flow is


P  Q  V 
V
where Q is the input or output volumetric flow-rate, and hence depends on whether the chamber
is connected to the supply pressure PS or reservoir Pr .

Because we have two pressure chambers (two fluid capacitances) we begin with the appropriate
first-order ODEs for a hydraulic system with compressible flow:

 
Chamber 1: P1 
V
Q  V 
1 1 Chamber 2: P2 
V2
Q 2  V2 
1

Because the chamber volumes are V1  A1 x and V2  A2 ( L  x) , the volume time-rates are
V1  A1 x and V2   A2 x . The orifice (valve) flow rates for chambers 1 and 2 are (using orifice
area A0  z h )

Q1  Cd h z
2
PS  P1  for z > 0 or Q1  Cd h z
2
P1  Pr  for z < 0 (A)
 

Q2  Cd h z
2
P2  Pr  for z > 0 or Q2  Cd h z
2
PS  P2  for z < 0 (B)
 

Next, consider the free-body diagram (FBD) of the mechanical component (piston + load)

P1A1 P2A2
FL
bx

Summing forces (positive is to the right), we obtain P1 A1  P2 A2  FL  bx  mx

Finally, substitute for the volumes and volume time-rates. The complete mathematical model is


P1  Q1  A1 x 
A1 x

P2  Q2  A2 x  Mathematical model of hydraulic servo
A2 ( L  x)
mx  bx  P1 A1  P2 A2  FL

The flow rates Q1 and Q2 are determined by Eqs. (A) and (B) and depend on the sign of the spool-
valve displacement z, which is an input to the system.

19
Copyright © 2016 John Wiley & Sons
Chapter 4

4.18 Because we have a single fluid capacitance (single vessel), we begin with the appropriate
first-order ODE for a pneumatic system with compressible flow:

nRT  P 
P   win  wout  V
V  RT 

a) Since volume is V = 25 m3 = constant then V  0 . The oven has no in-flow, so win = 0. The
(out) mass-flow through the valve is either choked or unchoked depending on the ratio Patm/P.
Therefore, the complete model is

 1
 2

2  P    P 
 atm    atm  

CP  Cd A0 P if Patm / P  Cr (unchoked)
(  1) RT  P   P  
 

 1

CP  Cd A0 P Cr  if Patm / P  Cr (choked)
RT

V
where the pneumatic capacitance is C 
nRT

b) Pneumatic capacitance can be computed using the above expression with n = 1 (isothermal),
V = 25 m3, T = 373 K, and R = 300 N-m/kg-K. The result is C = 2.2341(10-4) kg/Pa.


 2   1
The critical pressure ratio is Cr     0.5266 for coke gas.
  1

Because the initial pressure ratio is Patm/P = (1.103E5 Pa)/(9E5 Pa) = 0.1126 < Cr the valve flow
is initially choked.

 1
 
Initial (choked) mass-flow rate: wout  Cd A0 P Cr = 0.7387 kg/s
RT

20
Copyright © 2016 John Wiley & Sons
Chapter 4

c) The following M-file computes the mass-flow rate for pressures ranging from 0.9 MPa to Patm

% Problem 4.18c

% constants
R = 300; % gas constant, N-m/kg-K
T = 373; % temperature, K
gamma = 1.41; % ratio specific heats
Cd = 0.8; % discharge coefficient
A0 = 5e-4; % orifice area, m^2
P_atm = 1.013e5; % ambient pressure, Pa
Cr = (2/(gamma+1))^(gamma/(gamma-1));
pow1 = 2/gamma;
pow2 = (gamma+1)/gamma;

P_gas = linspace(9e5,P_atm,500);
for i=1:500
P = P_gas(i); % coke-gas pressure, Pa
P_ratio = P_atm/P;
c = Cd*A0*P; % common constant
if P_ratio > Cr
% unchoked flow
w(i) = c*sqrt(2*gamma*(P_ratio^pow1 - P_ratio^pow2)/((gamma-1)*R*T));
else
% choked flow
w(i) = c*sqrt(gamma*Cr^pow2/(R*T));
end
end

plot(P_gas,w)
title('Mass-flow rate through valve')
grid
xlabel('Oven pressure, P, Pa')
ylabel('Mass-flow rate, w, kg/s')

Mass-flow rate through valve


0.8

0.7

0.6
Mass-flow rate, w, kg/s

0.5

0.4

0.3

0.2

0.1

0
1 2 3 4 5 6 7 8 9
Oven pressure, P, Pa 5
x 10

21
Copyright © 2016 John Wiley & Sons
Chapter 4

4.19 Because the pneumatic “air spring” force is F = kairx, a positive displacement dx will
result in a positive differential pressure force dF = AdP. Therefore, we can write kair as

dF AdP
kair  
dx dx

A positive change in displacement dx > 0 produces a negative change in volume: dV   Adx .


Therefore, the “air spring” constant becomes

dP
kair   A2
dV

For a polytropic process we have


n
m 
P  a  a  air   amair
n n
V n
 V 

So, taking differentials yields

dP  namair
n
V  n 1dV or dP  nPV 1dV

Forming the ratio dP/dV and using the nominal pressure and volume we obtain

dP P
 n 0
dV V0

Finally, substituting the above expression into the “air spring” constant equation yields

P0
kair  nA2 “air spring constant” (N/m)
V0

22
Copyright © 2016 John Wiley & Sons
Chapter 4

4.20 Because we have a single fluid capacitance (single chamber), we begin with the
appropriate first-order ODE for a pneumatic system with choked flow:

nRT  P 
P   win  V if supply valve is open
V  RT 

nRT  P 
P    wout  V if exhaust valve is open
V  RT 

The chamber volume is V  V0  A1 x and therefore the volume time-rate is V  A1 x .

The incoming air mass-flow when the supply valve is open is

 1
 
win  Cd A0 PS Cr (choked flow) (A)
RTS

 2   1
where Cr is the critical pressure ratio, C r    , and TS is the supply tank temperature.
   1

The exhaust air mass-flow when the exhaust valve is open is

 1
 
wout  Cd A0 P Cr (choked flow) (B)
RT

Next, consider the free-body diagram (FBD) of the mechanical component (piston)

PA1 PatmA2
FL
bx

Summing forces (positive is to the right), we obtain PA1  Patm A2  FL  bx  mx

Finally, substitute for the volume and volume time-rate. The complete mathematical model is

nRT  P  nRT  P 
P   win  A1 x  (supply) or P    wout  A1 x  (exhaust)
V0  A1 x  RT  V0  A1 x  RT 

mx  bx  PA1  Patm A2  FL

where win and wout are determined by Eqs. (A) and (B).

23
Copyright © 2016 John Wiley & Sons
Chapter 4

4.21 a) The thermal system consists of a single chamber (thermal capacitance C), so we begin
with a single first-order ODE for closed thermal systems (neglecting the mass-transfer terms)

CT  qin  qout   qout


T  Ta
where the input heat-transfer is zero and output heat-transfer rate is qout 
R
Hence, the model is
RC T  T  Ta Thermal system model

b) The initial output heat-transfer rate to the ambient surroundings is

T (0)  Ta
q(0) 
R

We know that the initial chamber temperature is T(0) = 35 deg C = 308.15 K and the ambient
temperature is Ta = 25 deg C = 298.15 K. Therefore, we need to compute the thermal resistance
R. For conduction, the thermal resistance is

x
R where x = thickness, A = area, and k = thermal conductivity coefficient
kA

Searching the literature, we find that k = 0.21 W/m-K for cardboard. Hence, using x = 0.015 m
and A = 0.22 = 0.04 m2, we find that R = 1.7857 K/W. Using this value of R and the initial
temperature difference of 10 K, the initial heat-transfer rate is q(0) = 5.6 W.

c) Using the thermal modeling equation, the initial time-rate of temperature is

T  T ( 0)
T (0)  a (A)
RC

Now we need thermal capacitance, C = mcp. The air mass is computed from its density and
chamber volume: m = V, where volume is V = (0.2 m)(0.2 m)(0.4 m) = 0.016 m3. We can
compute air density using the perfect gas law:

P

RairT

Assuming air pressure is P = 1.013(105) N/m2, gas constant Rair = 287 N-m/kg-K, and temperature
T = 308.15 K, we obtain density = 1.1454 kg/m3. Hence, the air mass is m = 0.0183 kg. The
specific heat at constant pressure for air is cp = 1010 J/kg-K. Finally, the thermal capacitance is
computed to be C = 18.51 J/K.

The denominator in Eq. (A) is RC = (1.7857 K-s/J)(18.51 J/K) = 33.0536 s. Since the initial
temperature difference is –10 K, the initial time-rate of temperature is T (0) = –0.3025 K/s.

24
Copyright © 2016 John Wiley & Sons
Chapter 4

4.22 The thermal system consists of a single chamber (thermal capacitance C), so we begin
with a single first-order ODE for closed thermal systems (neglecting the mass-transfer terms)

CT  qin   qout

where the input heat flow is qin and the summation of the five output heat-transfer rates is

5
T  Ta 5(T  Ta )
q out 
i 1 R

R

There are five output heat-transfer rates through the five surfaces of the kiln; each surface has the
same thermal resistance R.

Finally, substituting the above result into the first-order thermal system ODE we obtain

RC T  5T  Rq in  5Ta Thermal system model

Note that we could define the model in terms of the equivalent thermal resistance, REQ = R/5:

REQCT  T  REQqin  Ta Thermal system model using REQ

25
Copyright © 2016 John Wiley & Sons
Chapter 4

4.23 The system consists of a single chamber (thermal capacitance C), so we begin with a
single first-order ODE for closed thermal systems (i.e., neglect the mass-transfer terms)

CT  qin  qout   qout

where the input heat-transfer is zero and output heat-transfer rate is

T  Ta
qout  where R is the thermal resistance of the thermos’ insulation
R

Hence, the model is

RC T  T  Ta Thermal system model

which is the same mathematical model as the thermal chamber in Problem 4.21.

26
Copyright © 2016 John Wiley & Sons
Chapter 4

4.24 The boundaries of the two thermal capacitances C1 and C2 are shown below

q1 C1, T1 C2, T2 q3
q2

qin (heater)

Our basic thermal modeling equations are two first-order ODEs for closed thermal systems (i.e.,
neglect the mass-transfer terms)

Chamber 1: C1T1  qin  q1  q2 Chamber 2: C2T2  q2  q3

The three heat-transfer rates depend on the temperature difference and thermal resistance:

T1  Ta T1  T2 T2  Ta
q1  q2  q3 
R1 R2 R1

Substituting the above expressions for the heat-flow rates in the two ODEs yields

1 1  1 1
C1T1    T1  T2  qin  Ta Thermal system model
 R1 R2  R2 R1
1 1  1 1
C2T2    T2  T1  Ta
 R1 R2  R2 R1

Chamber temperatures T1 and T2 are the dynamic variables and input heat flow rate qin and
ambient temperature Ta are the system input variables.

27
Copyright © 2016 John Wiley & Sons
Chapter 4

4.25 The single capacitance thermal boundary for the open system is shown below:

hin C, T
qout

hout
qin (heater)

Because we have one thermal capacitance we begin with the single first-order ODE:

CT  hin  hout  qin  qout

We have heat-energy transfer due to the mass flowing in and out of the system.

Input enthalpy rate: hin  win c pTin Output enthalpy rate: hout  woutc pTout

The mass-flow rates are equal, so w = win = wout, and the output stream temperature is Tout = T

T  Ta
The heat-transfer leaving the thermal boundary is qout 
R

Substituting the above expressions for the heat-transfer terms in the ODE yields

1 1
CT  wc pT  T  wc pTin  qin  Ta Mathematical model
R R

28
Copyright © 2016 John Wiley & Sons
Chapter 4

4.26 The heat-exchanger system consists of two thermal capacitances: the steam chamber C1
and water in the copper tubes, C2. The steam chamber has insulation with thermal resistance
R1 and the copper tubes have thermal resistance R2. The two thermal boundaries for the open
system are shown below:

hin,1
Steam chamber
hin,2
qout

C2, T2 C1, T1
q1

Water in tubes
hout,2 hout,1

The fundamental first-order thermal modeling equation is applied to each capacitance:

Steam chamber: C1T1  hin,1  hout,1  q1  qout


Water stream: C2T2  hin,2  hout,2  q1

We have heat-energy transfer due to the mass flowing in and out of the system (steam and water)

Steam: input enthalpy rate: hin,1  win,1c p ,1Tin,1 output enthalpy rate: hout,1  wout,1c p ,1Tout,1
Water: input enthalpy rate: hin,2  win,2 c p , 2Tin,2 output enthalpy rate: hout,2  wout,2 c p , 2Tout,2

The two respective mass-flow rates are equal, so w1 = win,1 = wout,1, and w2 = win,2 = wout,2.

The output steam temperature is Tout,1 = T1 and the output water temperature is Tout,2 = T2 . Note
that the steam and water have different ratios of specific heats: cp,1 (steam) and cp,2 (water).

The heat-flow rate from the steam chamber to the water stream is q1 and the heat-flow rate from
the steam chamber to the ambient surroundings is qout, as shown below:

T1  T2 T1  Ta
Crossing copper tubes (R2): q1  Crossing steam chamber (R1): qout 
R2 R1

Substituting the respective heat-transfer terms into the two ODEs yields the complete model:

1 1  1 1
Steam chamber: C1T1  w1c p ,1T1    T1  T2  w1c p ,1Tin,1  Ta
 R1 R2  R2 R1
1 1
Water stream: C2T2  w2 c p , 2T2  T2  T1  w2 c p , 2Tin,2
R2 R2

29
Copyright © 2016 John Wiley & Sons

You might also like