Anda di halaman 1dari 197

DISTILLATION VARIABILITY PREDICTION

by
SATISH ENAGANDULA, B.Tech.
A THESIS
IN
CHEMICAL ENGINEERING
Submitted to the Graduate Faculty
of Texas Tech University in
Partial Fulfillment of
the Requirements for
the Degree of
MASTER OF SCIENCE
IN
CHEMICAL ENGINEERING

Approved

-r^

Chairperson of the ConplWttee

Accepted

InterimDean of the Graduate Schfco


December, 2000

ACKNOWLEDGEMENTS

I would like to take this opportunity to sincerely thank my advisor. Dr. James B.
Riggs for his constant support and encouragement throughout the project. I have
thoroughly enjoyed working on this project. I am highly indebted to him for his timely
guidance and direction without which this project could not have been realized. I also
appreciate the financial support provided by the Texas Tech Process Control and
Optimization Consortium members. I would also like to thank Dr. Karlene A. Hoo for
being a part of my thesis committee. I also want to thank Marshall Duvall for his help.
Personally, I would like to thank all my friends, Mukund, Namit, Kishor, Alpesh,
Govindhakannan, and Rohit for making my stay in Lubbock a memorable one. I will miss
you all. I also want to thank Matt, Erik, Daguang, Meisong, Andrei, Xuan Li, and Rodney
Thompson for transforming the department office into a fun place to work in. Finally, I
would like to thank my parents, my sister, Vedashree and my brother, Vineet for their
love and support, which has kept me going throughout these years.

TABLE OF CONTENTS

ACKNOWLEDGEMENTS

ii

ABSTRACT

vii

LIST OF TABLES

viii

LIST OF HGURES

LIST OF NOMENCLATURE

xiv

1. INTRODUCTION

2. LITERATURE REVIEW

2.1 Distillation

2.1.1 Distillation Dynamics

2.1.2 Dual-Ended Composition Control

2.1.3 Configuration Selection

2.1.4 Decentralized PI Controller Tuning

2.1.5 Inferential Composition Control

2.2 Product Variability

2.3 Signal Processing

3. PRODUCT VARIABILITY PREDICTION APPROACH


3.1 Why a Disturbance Test?

10
12

4. LINEAR DYNAMIC MODEL DEVELOPMENT

14

4.1 C3 Splitter-Binary Distillation Column

14

4.1.1 Modeling Assumptions

14

4.1.2 Vapor Liquid Equilibrium

15

4.1.3 Steady State Designs

17

4.1.4 Linear Dynamic Modeling

17

4.1.4.1 Invariant Structure

20

4.1.4.2 Interior Trays of the Distillation Column

20

4.1.4.3 Accumulator

23

4.1.4.4 Reboiler

23

4.1.5 Level Controllers

24
iii

4.1.6 Dynamic Simulation Development

26

4.1.7 Linear Model Benchmarking

27

4.2 Depropanizer - Multicomponent distillation column

30

4.2.1 Modeling Assumptions

30

4.2.2 Vapor Liquid Equilibrium

31

4.2.3 Depropanizer Steady State Designs

33

4.2.4 Linear Modeling

35

4.2.4.1 Interior Trays of distillation column

35

4.2.4.2 Accumulator

37

4.2.4.3 Reboiler

38

4.2.5 Depropanizer Level Controllers

38

4.2.6 Inferential Composition Control

39

4.2.7 Dynamic Simulation Development

40

4.2.8 Linear Model Benchmarking

41

5. DUAL-ENDED COMPOSITION CONTROL

44

5.1 Configuration Selection

44

5.2 Invariant Structure of a Distillation Column

45

5.3 Composition Controller Tuning Criteria

46

5.4 Composition Control Results

48

5.4.1 Base Case C3 Splitter

48

5.4.1.1 Setpoint Control Results

50

5.4.1.2 Closed-Loop Bode Plots

52

5.4.2 High Purity C3 SpUtter

57

5.4.2.1 Setpoint Control Results

57

5.4.2.2 Closed-Loop Bode Plots

60

5.4.3 Low Purity C3 Splitter

65

5.4.3.1 Setpoint Control Results

65

5.4.3.2 Closed-Loop Bode Plots

68

5.4.4 Inverted Purity C3 SpUtter

73

5.4.4.1 Setpoint Control Results


IV

73

5.4.4.2 Closed-Loop Bode Plots


5.4.5 Base Case Depropanizer

76
77

5.4.5.1 Setpoint Control Results

81

5.4.5.2 Closed-Loop Bode Plots

84

5.4.6 High Purity Depropanizer

89

5.4.6.1 Setpoint Control Results

89

5.4.6.2 Closed-Loop Bode Plots

91

5.4.7 Low Purity Depropanizer

96

5.4.7.1 Setpoint Control Results

98

5.4.7.2 Closed-Loop Bode Plots

98

5.4.8 Asymmetric Purity Depropanizer

103

5.4.8.1 Setpoint Control Results

104

5.4.8.2 Closed-Loop Bode Plots

106

6. SIGNAL PROCESSING TECHNIQUES

111

6.1 Discrete Fourier Transforms

112

6.2 The Sampling Theorem and Signal Aliasing

114

6.3 Digital Filtering in Time Domain

115

6.4 Treatment of End Effects by Zero Padding

116

6.5 Industrial Feed Composition Signal

117

6.5 Results of Signal Processing Analysis

118

7. PRODUCT VARIABILFFY PREDICTION

122

7.1 Prediction Technique

122

7.2 Closed-Loop Product Variability Prediction

124

7.2.1 Base Case C3 Splitter

125

7.2.1.1 Results

125

7.2.1.2 Discussion

130

7.2.2 High Purity C3 Splitter

131

7.2.2.1 Results

131

7.2.2.2 Discussion

136

7.2.3 Low Purity C3 Splitter

137
\'

7.2.3.1 Results

137

7.2.3.2 Discussion

141

7.2.4 Inverted Purity C3 Splitter

143

7.2.4.1 Results

143

7.2.4.2 Discussion

147

7.2.5 Base Case Depropanizer

148

7.2.5.1 Results

148

7.2.5.2 Discussion

153

7.2.6 High Purity Depropanizer

154

7.2.6.1 Results

154

7.2.6.2 Discussion

155

7.2.7 Low Purity Depropanizer

159

7.2.7.1 Results

159

7.2.7.2 Discussion

164

7.2.8 Asymmetric Purity Depropanizer

165

7.2.8.1 Results

165

7.2.8.2 Discussion

166

7.3 Summary

171

8. CONCLUSIONS AND RECOMMENDATIONS

173

8.1 Conclusions

173

8.2 Recommendations

176

LFFERATURE CITED

177

VI

ABSTRACT

A novel technique is proposed to predict product variabilities for distillation


columns. The technique uses industrial disturbance data and applies signal processing
techniques to extract its amplitude and frequency information. This information is
combined with the closed-loop Bode plot for the same disturbance as a function of
frequency to predict closed-loop product variabilities for the column. The closed-loop
Bode plot is obtained using a linear dynamic model of the process. The approach is
demonstrated using a binary distillation colunm, a C3 splitter and a multicomponent
distillation column, a depropanizer. Four different designs of both columns were
considered. A thorough study of the approach is carried out to verify the accuracy and the
shortcomings of the approach. The potential of the approach as a quantitative tool for
configuration selection was also explored. For this purpose, nine different distillation
configurations were analyzed which indicated that this approach can be successfully used
for distillation configuration selection.

Vll

LIST OF TABLES

3.1

Steady state and dynamic characteristics of a typical distillation column

12

4.1

C3 Splitter Steady State Design Parameters

18

4.2

4.3
4.4

ALmax and ^ chosen for calculating the level controller tuning parameters
for C3 splitter

26

Depropanizer Steady State Design Parameters

34

ALmax and ^ chosen for calculating the level controller tuning parameters
for Depropanizer

40

5.1

Controlled and Manipulated Variable pairings for dual PI composition control

45

5.2

Base case C3 splitter dual PI composition controller tuning parameters

49

5.3

Base case C3 Splitter lAE indices for overhead impurity setpoint control

51

5.4

High Purity C3 splitter dual PI composition controller tuning parameters

58

5.5

High Purity C3 splitter lAE indices for overhead impurity setpoint control

60

5.6

Low Purity C3 splitter dual PI composition controller tuning parameters

66

5.7

Low Purity C3 splitter LAE indices for overhead impurity setpoint control

68

5.8

Inverted Purity C3 splitter dual PI composition controller tuning parameters

74

5.9

Inverted Purity C3 splitter LAE indices for overhead impurity setpoint control

76

5.10 Base case Depropanizer dual PI composition controller tuning parameters

82

5.11 Base case Depropanizer lAE indices for overhead impurity setpoint control

84

5.12 High Purity Depropanizer dual PI composition controller tuning parameters

91

5.13 High Purity Depropanizer LAE indices for overhead impurity setpoint control

91

5.14 Low Purity Depropanizer dual PI composition controller tuning parameters

96

5.15 Low Purity Depropanizer LAE indices for overhead impurity setpoint control
99
5.16 Asymmetric Purity Depropanizer dual PI composition controller tuning
parameters
104
5.17 Asymmetric Purity Depropanizer LAE indices for overhead impurity setpoint
control
106
7.1

Base case C3 Splitter LAE indices for Product Variability Prediction

130

7.2

High Purity C3 Splitter lAE indices for Product Variability Prediction

136

7.3

Low Purity C3 Splitter lAE indices for Product Variability Prediction

142

viii

7.4

Inverted Purity C3 splitter lAE indices for Product Variability Prediction

147

7.5

Base case Depropanizer lAE indices for Product Variability Prediction

153

7.6

High Purity Depropanizer lAE indices for Product Variability Prediction

159

7.7

Low Purity Depropanizer lAE indices for Product Variability Prediction

164

7.8

Asymmetric Purity Depropanizer LAE indices for Product Variability


Prediction

167

IX

LIST OF HGURES

1.1

Schematic of proposed approach for predicting product variability

11

4.1

Relative volatility variation of the propylene-propane system at 211 psia

16

4.2

Typical Structure of a distillation column

19

4.3

Invariant Structure of a distillation column

21

4.4

Distillation Tray Schematic

21

4.5

Comparison of open loop responses for C3 splitter

28

4.6

Comparison of open loop responses for depropanizer

42

5.1

Comparison between the closed-loop responses of the linear and non-linear


model for setpoint tracking of the [D,V] configuration of the Base case C3
Splitter

50'

Base case C3 Splitter Closed-Loop Bode plot for feed composition


disturbance rejection for [L,B] configuration

52

Base case C3 Splitter Closed-Loop Bode plot for feed composition


disturbance rejection for [L/D,V] configuration

53

Base case C3 Splitter Overhead Impurity Amplitude ratio plots for feed
composition disturbance rejection

55

Base case C3 splitter Bottoms Impurity Amplitude ratio plots for feed
composition disturbance rejection

56

5.2
5.3
5.4
5.5
5.6

Comparison between the closed-loop responses of the linear and non-linear


model for setpoint tracking of the [D,V] configuration of the High Purity C3
Splitter
59

5.7

High Purity C3 Splitter Closed-Loop Bode plot for feed composition


disturbance rejection for [L,V] configuration

62

High Purity C3 Splitter Closed-Loop Bode plot for feed composition


disturbance rejection for [L/D,V/B] configuration

62

High Purity C3 Splitter Overhead Impurity Amplitude ratio plots for feed
composition disturbance rejection

63

5.8
5.9
5.10

High Purity C3 Splitter Bottoms Impurity Amplitude ratio plots for feed
composition disturbance rejection

5.11

Comparison between the closed-loop responses of the linear and non-linear


model for setpoint tracking of the PD,V] configuration of the Low Purity C3
Splitter
67

5.12
5.13
5.14
5.15

Low Purity C3 Splitter Closed-Loop Bode plot for feed composition


disturbance rejection for [L,V] configuration

70

Low Purity C3 Splitter Closed-Loop Bode plot for feed composition


disturbance rejection for [L/D,V/B1 configuration

70

Low Purity C3 Splitter Overhead Impurity Amplitude ratio plots for feed
composition disturbance rejection

71

Low Purity C3 Splitter Bottoms Impurity Amplitude ratio plots for feed
composition disturbance rejection

72

5.16

Comparison between the closed-loop responses of the linear and non-linear


model for setpoint tracking of the [D,V] configuration of the Low Purity C3
Splitter
75

5.17

Inverted Purity C3 Splitter Closed-Loop Bode plot for feed composition


disturbance rejection for [L,V] configuration

78

Inverted Purity C3 Splitter Closed-Loop Bode plot for feed composition


disturbance rejection for [L/D,V/B] configuration

78

Inverted Purity C3 Splitter Overhead Impurity Amplitude ratio plots for


feed composition disturbance rejection

79

Inverted Purity C3 Splitter Bottoms Impurity Amplitude ratio plots for feed
composition disturbance rejection

80

Comparison between the closed-loop responses of the linear and non-linear


model for setpoint tracking of [L/D,V/B] configuration of base case
depropanizer

83

Base case Depropanizer Closed-Loop Bode plot for feed composition


disturbance rejection for [L,V] configuration

86

Base case Depropanizer Closed-Loop Bode plot for feed composition


disturbance rejection for [L/D,V/B] configuration

86

Base case Depropanizer Overhead Impurity Amplitude ratio plots for feed
composition disturbance rejection

87

Base case Depropanizer Bottoms Impurity Amplitude ratio plots for feed
composition disturbance rejection

88

Comparison between the closed-loop responses of the hnear and non-linear


model for setpoint tracking of [L/D,V/B] configuration of high purity
depropanizer

90

High Purity Depropanizer Closed-Loop Bode plot for feed composition


disturbance rejection for [L,V] configuration

93

High Purity Depropanizer Closed-Loop Bode plot for feed composition


disturbance rejection for [L/D,V/B] configuration

93

5.18
5.19
5.20
5.21

5.22
5.23
5.24
5.25
5.26

5.27
5.28

xi

5.29
5.30
5.31

5.32
5.33
5.34
5.35

High Purity Depropanizer Overhead Impurity Amplitude ratio plots for


feed composifion disturbance rejection

94

High Purity Depropanizer Bottoms Impurity Amplitude ratio plots for


feed composition disturbance rejection

95

Comparison between the closed-loop responses of the linear and non-linear


model for setpoint tracking of [L/D,V/B] configuration of low purity
depropanizer

97

Low Purity Depropanizer Closed-Loop Bode plot for feed composition


disturbance rejection for [L,V] configuration

100

Low Purity Depropanizer Closed-Loop Bode plot for feed composition


disturbance rejection for [L/D,V/B1 configuration

100

Low Purity Depropanizer Overhead Impurity Amplitude ratio plots for


feed composition disturbance rejection

101^

Low Purity Depropanizer Bottoms Impurity Amplitude ratio plots for


feed composition disturbance rejection

102

5.36

Comparison between the closed-loop responses of the linear and non-linear


model for setpoint tracking of |L/D,V/B] configuration of high purity
depropanizer
105

5.37

Asymmetric Purity Depropanizer Closed-Loop Bode plot for feed


composition disturbance rejection for [L,V] configuration

108

Asymmetric Purity Depropanizer Closed-Loop Bode plot for feed


composition disturbance rejection for [L/D,V/B] configuration

108

Asymmetric Purity Depropanizer Overhead Impurity Amplitude ratio


plots for feed composition disturbance rejection

109

5.38
5.39
5.40

Asymmetric Purity Depropanizer Bottoms Impurity Amplitude ratio


plots for feed composition disturbance rejection

110

6.1

Signal representation of feed disturbance entering a linear model

112

6.2

Signal Processing Procedure

117

6.3

C3 Splitter Feed Composition Signal

118

6.4

Depropanizer Feed Composition Signal

118

6.5

Signal Processing for C3 Splitter Signal

120

6.6

Signal Processing for Depropanizer Signal

121

7.1

Closed-Loop Rejection of disturbance of the Base case C3 Splitter

127

7.2

Closed-Loop Rejection of disturbance of the High Purity C3 Splitter

132

7.3

Closed-Loop Rejection of disturbance of the Low Purity C3 Splitter

138

xii

7.4

Closed-Loop Rejection of disturbance of the Inverted Purity C3 Splitter

144

7.5

Closed-Loop Rejection of disturbance of the Base case Depropanizer

150

7.6

Closed-Loop Rejection of disturbance of the High Purity Depropanizer

156

7.7

Closed-Loop Rejection of disturbance of the Low Purity Depropanizer

160

7.8

Closed-Loop Rejection of disturbance of the Asymmetric Purity


Depropanizer

168

Xlll

LIST OF NOMENCLATURE

Lower cutoff frequency for bandpass filter

An

Input Amplitude of signal n

An

Output Amplitude of signal n

AR(w)

Amplitude Ratio for frequency w

Upper cutoff frequency for bandpass filter

Bottoms Flowrate

Distillate Flowrate

EMV

Murphree Tray Efficiency

Feed Flowrate

Frequency

fc

Nyquist critical frequency

//.

Fugacity of component j in the vapor leaving tray i

f^\

Fugacity of component j in the liquid leaving tray i

(i, .

Liquid enthalpy of component j at tray i

/J<'

Ideal liquid enthalpy of component j at tray i

f{

Vapor enthalpy of component j at tray i

fj''

Ideal vapor enthalpy of component j at tray i

>]

.th

Hn

n frequency component of Discrete Fourier Transform (DFT)

h(t)

Time domain signal

1,1

Imaginary part of nth frequency component of DFT

Kc

Controller gain

KJ^

Controller gain for Tyreus Luyben settings

Kij

Vaporization Equilibrium ratio of component j at tray I

Ku

Ultimate gain from ATV test

Lj

Liquid flowrate of stream leaving tray i

LR

Reflux Flowrate
xiv

MACC

Holdup of the accumulator

Mi

Holdup of tray i

MREB

Holdup of the reboiler

Number of sample measurements

Pressure, psia

Pu

Ultimate Period from ATV test

PSDn

Power Spectral Density for n ^ frequency component

Reboiler Heat Duty

Rn

Real part of n^*' frequency component of D F T

XACC

Liquid composition at the accumulator

Xij

Liquid composition of component j at tray i

XREB

Liquid composition at the reboiler

yij

Vapor composition of component j at tray i

EMV

Murphree Tray Efficiency

Temperature

TBJ

Bubble temperature at tray i

Vapor Flowrate

Vi

Vapor flowrate of stream leaving tray i

var(t)

Product Variability Prediction

w,j

Angular frequency of n^ component

ZFJ

Feed Composition of component j

Greek Letters
a. ^

Relative volatility of component h to component j at tray i

AT

Sampling interval

6 (CO,,)

Phase Shift for the n^ frequency component

A/

Fugacity coefficient of component j in the vapor leaving tray i

Yi.j

AV

Fugacity coefficient of component j in the liquid leaving tray i

XV

0,,

Input phase angle of the n'*' frequency component

<pl

Output phase angle of the n frequency component

rj^

Controller integral time for Tyreus and Luyben tuning settings

T^

Controller integral time

T,^^ .

Hydraulic Time Constant of tray i

vi'

Angular frequency of the n^ component

Symbols
Zf/,,

Phase angle of n^ frequency component of DFT

h{f)

Digital filter response function

I// I

Amplitude of n frequency component of DFT

XVI

CHAPTER 1
INTRODUCTION

Distillation is one of the most widely used processes in the chemical processing
industries worldwide. Its use ranges from separation of heavy crudes to separation of
liquefied air. In the U.S alone, 40,000 distillation columns are used and they comprise
95% of the separation processes throughout the chemical processing industries (CPI)
(Degolyer & McNaughton, 1989). Producing products with low variability and with
minimal energy consumption are the most crucial factors for the success of companies in
the CPI. For cases, where the column product is a high value added product (e.g.,
feedstock for polymers) low variability may be a primary customer application and as a
result, upper product variability limits are many times product specifications. Reduction
in product variability can easily translate into reduction in energy/utility usage and
increased production rates. It will also help in reducing the variability in the downstream
units and can provide safety advantages too.
Distillation is an energy intensive process. Energy must be supplied to the reboiler
and removed from the condenser. It is estimated that U.S consumes 80 quads ^ of energy
annually, of which 7.25% is consumed by the CPI. Separation processes, mostly
distillation, in the CPI account for 43% of energy consumption in the CPI. (U.S Dept. of
Energy, 1988; U.S Dept. of Energy, 1989; Chemical Manufacturer's Association, 1989).
Humphrey et al. (1991) showed for various distillation applications such as ethylbenzenestyrene, propylene-propane, and methanol-water that excess energy in the range of 1015% is typically consumed during the column operation. Most of this consumption
resulted from manual operation of the column, or operations at greater than specified
purities as a safety margin. Improper control strategies largely accounted for increased
variability in products and increased utility usages. It may be concluded that the potential
for economic savings from even a small increase in efficiency of operation of distillation
columns, is great.

' 1 quad = lO'"* BTU, or 170 million bbl. of oil.

Distillation control has a dominant effect on the economic performance of a plant


(Degolyer and McNaughton, 1989). It is a challenging problem due to non-linearity,
coupling between manipulated variables, severe disturbances, and non-stationary
behavior. Single ended composition control has been shown to result in higher energy
consumption as it allows composition at one end of the column to float. Dual ended
composition control on the other hand has been able to reduce energy consumption but
results in dynamic stability and interaction problems (Chiang and Luyben, 1985). These
problems may be reduced by selection of proper pairings between manipulated and
control variables, i.e., configuration selection. The correct configurations are shown to
have a more profound effect on distillation control performance than conventional or
advanced control strategies. In fact, Duvall (1999), Anderson (1998), and Hurowitz
(1998) have shown that a reasonable control configuration can result in product
variabilities that are an order of magnitude worse than the optimum configuration.
Distillation configuration selection is not straightforward; it generally requires control
studies based on rigorous non-linear tray-to-tray column simulations, which is a lengthy
procedure.
The objective of this research is to address the distillation configuration selection
problem using an approach based on product variability to compare different
configurations. The approach is demonstrated on two distillation columns, namely, C3
splitter and depropanizer. This approach is an extension of the work by Hurowitz (1998).
In that work, it was demonstrated that it accurately predicted product variability for an
[L,B] configuration of a C3 splitter.
The C3 splitter represents a class of columns known as superfractionators
(Luyben, 1992). These columns, generally binary, separate close boiling mixtures such as
propane-propylene, ethylbenzene-styrene etc. Since the column products represent final
products, frequent upstream disturbances need to be minimized to maintain high product
quality. C3 splitters are characterized by low relative volatility (less than 1.2), high reflux
ratios, and a large number of trays. They typically have long open loop response times in
the range of 5-30 hours. Superfractionators typically have flat temperature profiles,
thereby hmiting the usage of tray temperature for inferential composition control. As a
9

result, only direct composition measurements can be taken which introduce sampling
time and transport delay problem. These factors account for the difficulties encountered
in the superfractionator composition control (Luyben, 1992).
The depropanizer is in the class of high relative volatility distillation columns,
(greater than 1.5). Any disturbances in the column are passed to the downstream units.
These columns are generally multicomponent columns and have fast dynamics and are
highly non-linear. To compensate for fast dynamics, inferential composition control using
tray temperatures is used to speed up composition control. This may also address the
deadtime and sampling times associated with composition measurements (Carling and
Wood, 1986; Duvall, 1999).
The objective of this research addresses the issue of configuration selection. For
this purpose a novel approach of predicting product variability is proposed. The approach
identifies a linear dynamic model and predicts product variability. The procedure is
carried out in the fi-equency domain. The approach extracts frequency information from
the feed disturbances and combines it with the information in a closed-loop Bode plot for
the same feed disturbance generated using the linear dynamic model to predict product
variability. This research demonstrates that this approach can be successfully used to
identify the optimum configuration for a distillation column. To accomplish these
objectives, the following stages are followed: (I) develop linear dynamic models for the
C3 splitter and the depropanizer, (2) analyze signal processing techniques to extract
frequency information, and finally (3) compare product variabilities predicted by
different configurations to select the best configuration based on the product variabilities.
Reviews on distillation, C3 splitter, depropanizer and signal processing are
provided in Chapter 2. The description of the proposed approach to predict product
variability is presented in Chapter 3. Linear dynamic models of the C3 splitter and the
depropanizer needed for implementing the approach are developed in Chapter 4. Chapter
5 describes the signal processing techniques used to extract the frequency content from
industrial feed composition signals; Chapter 6 contains the results of the product
variability prediction and the impact on configuration selection. Lastly, Chapter 7
summarizes the results and proposes future research in this area.

CHAPTER 2
LFTERATURE REVIEW
Since distillation is one of the most common processes used in the chemical
processing industries, a large body of knowledge describing its dynamics, operation and
control can be found in the open literature (Luyben, 1992; Kister, 1992, 1990; Buckley et
al., 1985; Shinsky, 1984, Rademaker et al., 1975, Nissenfled and Seeman, 1981;
Deshpande, 1985).
In this chapter, the literature about the dynamics and control of distillation
columns are reviewed especially as it concerns C3 splitter and depropanizer. This is
followed by a discussion on dual-ended composition control, configuration selection and
controller tuning as they are relevant to this research. Section 3 contains some signal
processing information germane to the understanding of this research.

2.1 Distillation
2.1.1 Distillation Dynamics
Fuentes and Luyben (1983) first studied dynamics of high purity distillation
columns. They studied the impact of relative volatility on these columns. They found that
with higher relative volatility the dynamics becomes faster and non-linear and hence,
difficult to control. They recommend using faster composition measurements using
inferential measurements such as tray compositions for high relative volatility columns.
Skogestad and Morari (1988) studied the dynamic behavior of distillation columns by
developing simple and analytical models. They found that high purity columns with large
reflux rates would be most difficult to control. They proposed using logarithmic
compositions in the model to reduce the non-linearity of the column. Carling and Wood
(1986) and Wang and Wood (1985) studied the behavior of multicomponent columns
such as the depropanizer and observed that they exhibit non-minimum phase behavior
(inverse responses) in the light key (propane) and heavy key (iso-butane). Non-minimum
phase behavior further contributes to the difficulties encountered during multicomponent
distillation control (Carling and Wood, 1986).

2.1.2 Dual-Ended Composition Control


Distillation accounts for a large percentage of energy used in the chemical
processing industries. Hence a reduction in energy consumption could mean significant
economic savings. Luyben (1975) and Chiang and Luyben (1985) based on steady state
calculations showed that dual composition control considerably reduces energy
consumption as compared to single ended control. They, however, pointed out the
dynamic stability and interaction problems posed by dual composition control. Ryskamp
(1980) and Stanley and McAvoy (1985) reported industrial energy savings of 10-30% on
the use of dual composition control. They observed that dynamic simulations showed
much larger energy savings as compared to those predicted by steady state analysis.
However, Freuhauf and Mahoney (1994) recommended usage of single
composition control because of the difficulty involved in implementing and maintaining
dual composition control. For heat integrated distillation columns, Hansen et al. (1998)
showed that dual-ended composition control provided better results than single-ended
control.

2.1.3 Configuration Selection


Plenty of literature has been published addressing the issue of configuration
selection. It is widely accepted that no particular configuration is the best for all
distillation columns.
Skogestad and Morari (1987), Skogestad (1990) and Shinsky (1984) proposed a
set of guidelines to identify the best configuration for a distillation column. Shinsky
(1984) approached the configuration selection problem using a steady state relative gain
array (RGA). Skogestad and Morari (1987) also recommended an RGA analysis for
choosing the best configuration but based their studies on a linear model of the
distillation column. Skogestad (1990) highlighted the usefulness of frequency dependent
RGA for configuration selection especially with respect to the [D,B] configuration.
In general, it has been found that P^/D,V/B] configuration was the most favored
control configuration for distillation control (Skogestad and Morari, 1987; Shinsky, 1984;
Skogestad, 1990). Skogestad and Morari (1987) recommended avoiding configurations

with large values of the entries of the RGA matrix because it may signal ill-conditioning.
They found that material balance configurations resulted in poor dynamic response and
poor disturbance rejection if one of the loops was not functional. This is consistent with
Shinsky's (1984) recommendation of avoiding material balance configurations.
Skogestad (1990) and Skogestand and Morari (1988) studied the same seven
distillation columns for configuration selection. For dual-ended composition control, the
[L/D,V/B] was found to be the best configuration for all seven columns. This agreed with
Finco (1987) and Finco et al. (1989) on the performance of the [D,B] configuration for
columns with high purity and/or large reflux rate.
Finco (1987) and Finco et al. (1989) carried out an extensive study of a C3 spHtter
and concluded that [L/D,V/B] and [D,B] structures were the best configurations. They
based their analysis on a non-linear dynamic model of an industrial C3 splitter. They also
pointed out that the [D,B] configuration lacked integrity and needed restructuring in the
case of valve saturation and sensor failures. They also recommended that tight level
control be maintained when using [D,B] configuration.
Gokhale et al. (1994) also studied an industrially benchmarked non-linear C3
splitter model using decoupled PI controllers. They concluded that the [L,B]
configuration and [D,B] with tight level control were the best configurations. Hurowitz
(1998) carried out an extensive analysis of an industrially benchmarked C3 splitter and
found [L,B], [D,B] with tight level control and IL,V/B] configurations to be the most
suitable configurations for their C3 splitter.
Carling and Wood (1986) and Freitas et al. (1994) found the [L,V] to perform
better than the [L,B1 configuration for a depropanizer. Duvall (1999) reconmiended
[L/D,V/B] and [L,V/B] configurations based on his analysis using an industrially
benchmarked non-linear tray to tray depropanizer model. He also found [L,V] to provide
a reasonable control performance.

2.1.4 DecentraUzed PI Controller Tuning


A distillation column typically has two level control loops for controlling levels in
the reboiler and accumulator and two composition control loops for controlling overhead

and bottom compositions. Tuning of both level and composition control loops is crucial
for good distillation control. Since the level control loops are much faster than
composition control loops, the level control loops are tuned first followed by composition
control. Rademaker et al. (1975) recommend using tight level control for energy balance
configurations and detuned level control for material balance configurations. Marlin
(1995) has provided a detailed study of level control dynamics and tuning. For distillation
columns, he recommended sluggish and detuned level control. He observed that
overshoot or oscillatory behavior of level control loops might have an adverse effect on
composition control loops.
For tuning composition control loops, the interaction between the two loops
should be considered. Luyben (1986) demonstrated a procedure of tuning composition
loops using Ziegler and Nichols settings (Ziegler and Nichols, 1942). The initial tuning
parameters for both composition loops are obtained by treating both loops as independent
loops and obtaining Ziegler and Nichols tuning parameters for each based on transfer
function model of the process. The two loops are then detuned using a common detuning
factor determined by the Biggest Log Modulus Testing (BLT) method. A drawback with
the BLT method is that it requires all transfer functions in the models which may be
difficult to estimate using the pulse testing method (Moudgalya, 1987). Finco (1987), in
fact, showed that these identified transfer functions could be in error with the actual
process. Instead of transfer function models of the process, Astrom and Hagglund (1984)
suggested using a relay feedback technique, Autotone Variation (ATV) method.
Recently, Tyreus and Luyben (1992) proposed a new set of tuning rules for
processes with large time constants. They used the ATV method to find the Tyreus and
Luyben tuning parameters which they promptly detuned using a modified version of the
BLT method. They concluded that the new tuning settings offered better results than
Ziegler and Nichols settings.

2.1.5 Inferential Composition Control


Fuentes and Luyben (1983) showed that high relative volatility columns exhibit
fast dynamics and recommended the use of inferential measurements, such as tray

temperatures, to infer compositions. The relative volatility for a depropanizer ranges


between 1.5 and 2.0. Hence, tray temperatures were used to estimate compositions.
Inferential control can be accomplished using a single control loop or a cascaded loop.
Marlin (1995) and Riggs (1998,1999) provide a detailed description about inferential
measurement and control. Riggs (1998) also provided a simple procedure to identify the
best trays for inferring compositions. Wolf and Skogestad (1996) considered a cascade
implementation and described several effects that should be taken into consideration for
selection of trays.
Shen and Lee (1998) successfully applied multivariable adaptive inferential
control to the Berry and Wood column to improve product quality. Joseph and Brosilow
(1978) carried out a thorough analysis of inferential control systems and demonstrated the
concepts using an industrial debutanizer. They reported improvements in steady state
performance by as much as 400% and reported that an inferential control systems
response to various disturbances were superior to that of tuned composition feedback
systems. McAvoy et al. (1996) considered a non-linear inferential control scheme by
applying a non-hnear inferential parallel cascade control (NIPCC) to the Tennesse
Eastman Challenge problem. For random feed fluctuations, they found the variance in
product flow and composition reduced significantly by using NIPCC.

2.2 Product Variability


Instrumentation such as transmitters and valves are recognized as key elements to
reducing product variability (Nelson, 1997; Weldon, 1999; Pyotsia et al., 1996; Blevins et
al., 1995). Few researchers have proposed novel tools to study process variability. CoUani
et al. (1989) first proposed a procedure to monitor and control process variability using
standard deviation control charts. They proposed process designs based on the economics
of cost inspection, collecting and analyzing data, and the cost and profit of renewals in
process design. A controllability index was developed by Zheng et al. (1999) based on
economic and product variability considerations to identify the best process design and
control structure. They demonstrated the concept on a binary distillation column. Tseng
et al. (1999) proposed a quantitative design tool based on frequency response of
8

individual parallel paths between inputs and outputs of a process to identify sources of
product variabihty and to make necessary design changes.

2.3 Signal Processing


A large body of material has been written describing different signal processing
techniques such as sampling, filtering, integral transforms, spectral analysis, etc. (Press et
al., 1991; Kamen, 1990). Walker (1991) has provided a good detailed description on fast
Fourier transforms such as radix-2-fast Fourier transform. Press et al. (1992) has provided
codes in Fortran 77 for implemention procedures for the radi?c2-fast Fourier transform.
Hurowitz (1998) has provided a detailed description of employing signal processing
techniques on discrete signals.

CHAPTER 3
PRODUCT VARIABILITY PREDICTION APPROACH
Product variability is a concern for many industries. Thus success relies on their
ability to minimize the variability in the products and to maintain the product quality
within the specified limits. Since distillation columns represent 95% of the separation
processes using in the chemical processing industries (CPI), the variability introduced by
distillation processes plays a major role in the ultimate product quality. This work
proposes to address product variability for distillation columns. The approach is
described below.
The approach to predict product variability in distillation columns is shown
schematically in Figure 3.1. The main idea is to address the problem in the frequency
domain rather than in the time domain. The process to be studied is characterized by its
steady state and dynamic process conditions. These characteristics are used to generate a
linear model for the process. The steady state and dynamic operating conditions
characterizing a typical distillation column are summarized in Table 3.1. Based on a user
selected controller tuning criteria, the controller tuning parameters are identified and
tuned online. Using these tuning parameters, the linear model is executed under closedloop conditions at different frequencies of the feed disturbances. From the resulting
responses at these various frequencies, amplitude ratios and phase angle shifts are
calculated to develop a closed-loop Bode plot for the feed disturbance. Industrial
disturbance data can be processed to extract the amplitude and frequency components of
the signal. These are combined with the closed-loop Bode plot to predict product
variability.
This research is an extension to the previous work by Hurowitz et al. (1998).
There it was demonstrated for product variability prediction for [L,B] configuration on a
C3 splitter. The schematic of that approach is analogous to the approach shown in Figure
3.1 except that the current approach uses a linear dynamic model. The signal processing
techniques were improvement to the signal processing techniques used (Hurowitz, 1998).

10

W)
C3

<L)

^ i:J 3

c
O

O
OH

C/0

o
U

73
O

oLH

Xi

o
O
OH

ci
13
<U
Vi

O
DO

bX) <i\
1-H

cj

00

I>

o
C/5

c^
cd

* - >

a Q

OH

(U

73
(1)

c3
P X)
c/) V i
C3 P
a^ +-

WH

C/5

-H

amic
terist

-4-^

a
> o
cd
Q
^

11

Table 3.1 Steady state and dynamic characteristics of a typical distillation column
State

Characteristics

Steady State

Total Number of Trays


Feed Tray location
Murphree Tray Efficiency
Column pressure
Reflux Ratio
Boilup ratio
Feed Row Rate
Feed condition
Feed composition
Distillate Flow Rate
Distillate Composition
Bottoms Flow Rate
Bottoms Composition

Dynamic

Accumulator and Reboiler Residence Time


Tray Hydraulic Time Constant
Composition Analyzer Deadtime
Composition Analyzer Sampling Rate

3.1 Why a Disturbance Test?


A disturbance test is one of the most severe tests for any controller with fixed
parameters. In the case of a distillation column, a feed composition disturbance is not
uncommon and usually has a significant impact on the column's performance. Hence a
change in feed composition can be used to evaluate the performance of different control
configurations. Moreover, feed compositions can sometimes be measured online.
The approach shown in Figure 3.1 is demonstrated using two distillation columns
namely, a C3 spHtter and a depropanizer. The potential of the approach as a quantitative
12

tool for configuration selection is also evaluated. The C3 spHtter can be classified as a
binary distillation column, which separates a mixture of predominantly propane and
propylene. It represents a class of columns known as superfractionators, which separate
close boiling mixtures (Luyben, 1992). They are characterized by low relative volatility
(less than 1.2), high reflux ratios and a large number of trays. The depropanizer is a
multicomponent distillation column separating a mixture of propane, ethane, n-butane,
hexane, i-butane, and pentane. The mixture has a high relative volatility (greater than 1.5)
and is usually an intermediate separation process in an industrial plant. The depropanizer
and the C3 splitter represent two very different column types and can be used to
demonstrate the approach. Linear dynamic tray-to-tray simulators under dual PI
composition control are developed for both distillation columns based on their steady
state and dynamic operating conditions (Gokhale et al., 1995; Duvall, 1999). The
composition and level controllers are tuned based on a preset tuning criteria and arc used
to generate a closed-loop Bode plot for feed composition disturbances. Industrial feed
composition disturbance data for the two columns are used to generate the required
frequency information. This frequency information is then combined with closed-loop
Bode plot information of the disturbance to predict product variability. The results will be
compared with rigorous non-linear tray-to-tray simulations.

13

CHAPTER 4
LINEAR DYNAMIC MODEL DEVELOPMENT

In this chapter, we study two distillation columns namely the C3 splitter and the
depropanizer. The C3 spUtter is a binary distillation column whereas the depropanizer is a
multicomponent distillation column. The models for the C3 splitter and the depropanizer
considered here are adopted from Hurowitz (1998) and Duvall (1999), respectively. They
developed rigorous non-linear dynamic models for these columns and benchmarked them
against industrial data. These rigorous non-linear models are linearized around their
steady state operating conditions to obtain linear dynamic models. The linear models are
tested by comparing their open-loop responses to that of the rigorous non-linear models.
Throughout the chapter, the condenser and accumulator refer to as the combined system
of condenser and accumulator.

4.1 C3 Splitter-Binary Distillation Column


4.1.1 Modeling Assumptions
The C3 splitter is one of the most important and widely used distillation columns in
the industry. The assumptions made in the development of this model are as follows:
1. The input feed stream is a binary system of propane and propylene.
2. The column has a single feed stream and two product streams, the bottoms and the
distillate.
3. Pressure remains constant throughout the column.
4. Since propane and propylene have identical molecular weights and heats of
vaporization, an equimolal overflow assumption is valid.
5. Vapor-holdup is negligible.
6. At each tray of the column, the vapor is in equilibrium with the liquid. The same is
true at the reboiler.
7. Vapor-liquid equilibrium (VLE) is represented by a relative volatility model. The
relative volatility is a function of pressure and liquid phase propylene mole fraction.

14

8. The trays are non-ideal and can be described by incorporating a single Murphree tray
efficiency (Murphree, 1925) for all trays except the reboiler. The reboiler is treated as
an ideal stage.
9. The column uses a reboiler and a total condenser.
10. Tray Hquid hydrauHcs is calculated using the hydraulic time constant approach
presented by Franks (1972) and Luyben (1990).
11. Heat-transfer dynamics in the reboiler and condenser are not considered in the model.
12. The liquid levels and compositions at the reboiler and accumulator are controlled
using PI controllers.

4.1.2 Vapor Liquid Equilibrium


Propane and propylene form a binary mixture. The vapor liquid equilibrium
between these components can be described using a relative volatility model. For a binary
system, it can be represented using the following equation,
y, =

-^

(4.1)

l + {a^,-i)x^
where yi, the vapor composition of component 1 is represented as a function of jci, the
liquid composition of component 1 and a , , , the relative volatility of component 1 to
component 2. In general, the relative volatility could be a function of composition, and
pressure or temperature. The correlation for relative volatility for propane-propylene
mixture was developed by Finco (1987), using the thermodynamic data provided by Hill
(1959). The correlation treated the relative volatility as a function of liquid composition
and pressure and is given as,

a^ = 1.285500 - 0.00044600P

(4.3a)

a^ = 0.008008-0.00010350P

(4.3b)

a._ = 0.052215 - 0.00014607?

(4.3c)

where P is the pressure in psia, x is Uie hquid phase propylene mole fraction and a is the
relative volatility of propylene to propane. The relative volatility is a measure of the
15

degree of separation, higher relative volatility (> 2) implies an easier separation. The
closer the value of the relative volatility to I, the more difficult the separation. The
relative volatility predicted by the above equation. Figure 4.1 shows that that the relative
volatihty for propane-propylene mixture at 211 psia lies between 1.1 and 1.2 indicating
that the separation is very difficult.

0.2

0.4

0.6

0.8

Liquid Phase Propylene Mole Fraction

Figure 4.1 Relative volatility variation of the propylene-propane system at 211 psia
In the dynamic model of the C3 splitter, VLE equations are solved on each tray.
Hence, in order to generate a linear dynamic model, the VLE equations need to be
hnearized at each tray. That is.
/

da 1.2./
1.2./ +

yu =

dx,

(l-h(a,.,,-l)x,.,)
Jss

da.,1.2.1
dx\.i

(4.4)

'1./

(4.5)

= - ( ! . / + 2 f l , , . A-,,,) 55
/55

where, tiie subscript / refers to tiie i^ tray, around whose steady state (SS) tiiese
equations are computed. The variables with a bar represent deviations from their steady
state. This convention will be used throughout this chapter.
16

4.1.3 Steady State Designs


Five different C3 splitter steady state designs developed by Hurowitz (1998) were
studied. These designs differed from each other with respect to product purity
specifications. The basic procedure followed for developing these designs can be
summarized as:
1. The minimum number of theoretical trays required for the desired separation were
calculated using Fenske's equation (Fenske, 1932).
2. The minimum reflux ratio was calculated using Underwood's (1948) equation.
3. The actual reflux ratio was taken as 1.2 times the minimum reflux ratio.
4. The number of theoretical trays was determined using Eduljee's equation (1975).
5. The actual number of total trays was determinedusing Lockett's (1986) equation.
6. Finally, the feed tray location was determined by minimizing the operating reflux
ratio.
The design parameters for the five different designs are sunmiarized in Table 4.1
4.1.4 Linear Dynamic Modeling
The structure of a typical distillation column, such as a C3 splitter, is shown in
Figure 4.2. A saturated Hquid feed stream with flowrate F and propylene mole fraction ZF
is introduced into the distillation column at tray Np. The exit liquid stream from the
bottom tray of the column is partially recycled back to the column through the reboiler.
The remaining stream represents the bottoms product B, rich in heavier component
propane. The overhead vapor from the top tray is condensed in the condenser. Part of the
condensed Hquid is recycled to the column and is called the reflux, R. The remaining
Hquid is the distillate product D, rich in the lighter component propylene. The reflux
combined with the liquid feed stream and the vapor generated by the reboiler creates the
vapor/Hquid traffic throughout the column.

17

c
3
OH

T3
u,

CO
Ol

^ ^

>

to
00

o
<N

(N

OH

O
en

'
00

ol

VO
00
ON

n
ON
o
r- Tf
d
d

^
^
Tt
Tj-

00

ON

ON

00

ON

ON

-^

(O

o
o
d

00

CO
CO

r>;

ON

oo
^

<N
H

'^

<N

>n

CO

in

CO

uo

Wi

CO

o
<N

ON

n
00

NO
^

ON

^
^
^
-^

to
00

U
Tt
^

VO

(N
<N
O
ON

NO

Vi

CO
<N

ON

<N
(N

o
00
<N
O
o
i^
'sf O N
<N
d
P
d
d d d d d d

iO
00

en

r->
d>

^
Tf
Tj-^

VO

oi

00
CO
NO

u-i
lO

o
00
o
o
NO
r^ o
ro
d d d d d d

^
^
^
"*
ON
NO

ON
CO
CO
CO
00
-^

o
rd d

ON
ON
ON

m
o
^
r-

ON
ON

^^
^
C4

o
<N
p
d d

Vi
(U

B
a
a

B
o

'vi

<->
41

B
"S

PQ

4>

Vi

>>
C3
<U

41

00

o
41

f. .
'4H

O
1J

00

u
x>

ed
on

CJ
C

Vi

c
o
41
C3

sD

:zi
'4->
rt

^VH

o
>>
a

73
O
(U

o
H tU

Vi

4_<

PH

X)

41

00

(/I
<D

OH

0^

ck
O

X
3

'o
U

*4I

o
a
V4

s s

..M

.2

Ui

a.

c
o

c
top

73

T3

ex
o
S-i

a
PC

o
Vi

OH

X)
o
4.^

a
PC

C3

^
o
c
o

C/3

V!

4>
4>

tin

o
X

c
4.^

C
O

CQ

.1.H
4-'

T3

OS
(U

00

O
N

VH

<u
N

C
<

C
<

"3

c
o

...^

4-1

o
o
<

a
o

4->

73

o
o

C3

1H

OH

Cu

13
O

.1^

PC

4>

Vi

4>

4>

'o

a.
o o

o
c

>^

'vi

cx

OH

o
U

Condenser
Duty (Qc)

Accumulator
N

Reflux (LR)

Distillate (D)

Feed (F, Zp)

Vapor Boilup (V)


Reboiler
Duty (QR)
Reboiler

Bottoms (B)

Figure 4.2 Typical Structure of a Distillation Column

19

From a process control viewpoint, the independent variables for the process are
feed flowrate, F, propylene feed composition, zp, bottoms flowrate, B, distillate flowrate,
D, reflux flowrate, LR, vapor flowrate, V, from the reboiler and the constant pressure, P
throughout the column. For the C3 spHtter, the F, zp and P form the disturbances to the
system whereas the remaining variables, namely B, D, R and V can be used as
manipulated variables. It should be realized that the steam used in the reboiler and the
cooHng water used in the condenser are not modeled for in the process. Instead, the reflux
R and vapor flowrate V are treated as direct inputs. Luyben (1972) has provided a
detailed non-Hnear model of an ideal binary distillation column. The C3 spUtter model is
developed along the similar lines.

4.1.4.1 Invariant Structure


For a distillation column, choosing which manipulated variable to control each
controlled variable is called configuration selection. It is one of the most important
assessments associated with distiUation control. Figure 4.3 represents a particular
invariant structure of distillation column, which faciHtates in addressing this issue. In this
structure, we assume that the distillation column is made up of three separate entities
namely, the accumulator, reboiler and the interior trays of the distillation column. These
entities remain the same irrespective of which configuration is used; only the mode of
interaction between them will differ from configuration to configuration. As a result, the
same model can be used for any configuration. The advantages of having such a structure
to assess various configurations are explained in the next chapter.

4.1.4.2 Interior Trays of the Distillation Column


Let's consider a schematic of a tray in Figure 4.4. For the C3 spHtter, since the
latent heats of vaporization of propane and propylene are almost equal, the equimolal
overflow assumption is found to be a good approximation. This assumption means that
we need not consider the energy balance equation for each tray. Thus,
V=Vi=constant

i = 0,l,...N

(4.6)

20

A material balance on each tray results in (Luyben, 1972),


dM:
dt ^ =
j ^

A>.-A

(4.7)

= Xi^i^M-x.L. + V(y._, - y.)

(4.8)

Accumulator

Distillate

Feed

1.

Bottoms
Reboiler
Figure 4.3 Invariant Structure of a distillation column

-i+l

Li

Vi

Vi

Figure 4.4 Distillation Tray Schematic

21

Each tray is treated as a non-ideal tray with a Murphree tray efficiency EMV (Murphree,
1925). Hence,
y> = y>-. + E^(yr-y<-.)-

(4.9)

Also, each tray is modeled as a continuous strirred tank (CST) with a hydrauHc time
constant, r,^^^. Hence, Equation (4.7) becomes.
^,=^..X,

(4.10)

dL,
dt

(4.11)

_L,,,-L,
tra\

For C3 splitter, the hydraulic time constant model is used for all trays throughout
the column (Hurowitz, 1998). To obtain a Hnear model for the C3 spHtter, the above
rigorous non-linear material balance equations at each tray are linearized around the
steady state operating conditions of that tray. The Hnearization is carried out using a first
order Taylor series of expansion, which results in.
dL-

_ Li+i Li

dt

(4.12)

tra\

dx.
dt

JC.(+1

-x,\
^i

-r
(L.A
"i+X
i,-.,-H
^m-1
K^'uss

Jss

fy

f V ^

M..

VJss

^L

^
X:

V ^ '

JSS

(4.13)

yi-1-1
V ^ . JSS

V, =yi-i + Em^(yJ'

' JSS

(4.14)

-yi-i)-

For the tray where the feed stream is introduced,

dL

A>.-A+^

dt

rrav

d(x.M-)

(4.15)
1 , \7(

\ , c

(4.16)

dt
(4.17)
yi = yi-i + EMv(yi'
dl,

dt

-yi-i)

(4.18)

L,>I-L/-HF
tray

ni

-^,>|-^/

dt

A>.+

-\',>1-1

/:,x
- k;
,x , 1 L.,
N*^/
^/-i-^/-i
"i+l + F )
V-

M:

V^'A.

/55

Jss

M:

( V ^
X- -

Jss

M.' JSS
,

yi

yi-i+iZr-Xi)ssF + (E)ss^^
K^'jss

(4.19)
y, =y,_,+^(y;'^-y._,)

(4.20)

4.1.4.3 Accumulator
For the accumulator, the material balance equations are
dM ACC
=
dt

(4.21)

V-L,-D

-l.^^-^l
at

(4.22)

y^,V-(L,+D)x ACC

The accumulator is treated as a well-mixed vessel. Since the cooling water is not
considered a part of the distillation column, the energy balance equation is not
considered. To generate a linear model for the accumulator, the above equations are
linearized around the steady state operating conditions of the accumulator
dM ACC
dt

d(x.rc)
dt

(4.23)

V-L,-D
V

M ACC

yN
y55

V ^
M ACC

^ACC

(4.24)

Jss

4.1.4.4 Reboiler
For the reboiler the material balance equations are given as.
(4.25)
dt
^y-^REB'^

dt

REB /

A'jL,

yREB^

-^REB^

The energy balance equation is not needed for the reboiler since the vapor
flowrate is set independentiy by a controller. The Hnear model for the reboiler is

23

(4.26)

dM
-^^ = L,-V
dt

-B

'

(4.27)
\

^1

dt

-^REB

^REB
/

M REB

\
y REB

L.-H
Jss

yJ^REB JSS

^REB

M REB

VJss

r L,-V
"REB

y ^REB

Jss

\
yREB

Jss

(4.28)
4.1.5 Level Controllers
The C3 splitter has two level control loops. PI controllers are employed in both
loops. The level controllers are actuaUy treated as part of the distillation column model as
this would facilitate in isolating the impact of composition control. In addition the level
control loops perform much faster than composition control loops.
Industrially, a hot wastewater stream is used as a heating medium for C3 splitter.
The hot wastewater stream experiences frequent, intermittent organic liquid
contamination, which adversely affects the reboiHng heat transfer, even if the wastewater
flowrate remains constant. This results in signinficant fluctuations in the reboiler which
can be dampened through loose level control. Hence, loose level control is a requirement
for C3 splitter (Gokhale, 1994). The level control loops can be tuned onHne but this
would consume considerable computation time. To enable faster tuning, analytical
expressions derived by Marlin (1995) for level controller parameters are used
straightaway. These equations are,
K^ = -0.736

T =

AF,

(4.29)

ALmax

A*f-

(4.30)

where AFuix represents the maximum expected flowrate change in the streams
entering/leaving, ALj^ the resulting expected maximum level change, and ^ the
damping ratio for the desired level control response. The above tuning parameters are
applied separately to each level control loop by assuming they are independent. The
24

level control loops are tuned for overdamped behavior. Hence, ^ was chosen in the range
of 2.5-15.
The levels in the reboiler and accumulator are directly affected by the variations
in the vapor Hquid traffic throughout the column. The vapor Hquid traffic in a column
varies around its reflux flowrate. A distiUation column in general witnesses a disturbance
of around 5-10% in the vapor liquid traffic during normal operation. Hence, it can be
reasonably assumed that a flowrate disturbance of around 5-10% of reflux flowrate
affects the levels in the reboiler and accumulator. Hence, for both level control loops,
AFmax is arbitrarily chosen as 5% of the reflux flowrate.
For the two level control loops, different ALmax and t, are chosen for different
configurations by taking into consideration the degree of coupHng between the two loops
and the dynamics expected. As an example, consider level controller tuning for the [L,B]
configuration. In this configuration, the distiUate flowrate is used to control the
accumulator level and the vapor flowrate is used to control the reboiler level. The
bottoms loop is independent, whereas, the overhead loop is coupled with the bottoms
loop. Changing the vapor flowrate in the bottoms loop would directiy affect the
accumulator level but changing the distillate flowrate in the overhead loop would not
affect the reboiler level. Thus, we may tune the bottom loop independently but the
overhead loop needs to be tuned in such a way so as to counter the interactions
introduced by the bottoms loop. This may be achieved by making the overhead loop more
sluggish as compared to the bottoms loop, i.e., by choosing larger AL,nax and <^ for the
overhead loop. Similar heuristics is used in the other configurations to determine the level
controller tuning parameters. For the different steady state designs, the common ALmax
and (^ , chosen for reboiler and accumulator level control, are shown in Table 4.2. The
different values for ALmax and ^ were compared with the results obtained by Hurowitz
(1998). The controllers were executed every 15 seconds. The last configuration shown in
Table 4.2 represents the [D,B] configuration for which tight level control is employed.
The [D,B] configuration with tight level control was investigated because Finco et al.
(1989) reported that excellent control performance for C3 splitter could be obtained.
25

Table 4.2 A Lmax and ^ chosen for calculating the level controller tuning parameters for
C3 spHtter
Configuration
AL^ax
^
Level Controller AL
L,B

L,V

L,V/B

D,B

D,V

D,V/B

L/D,B

L/D,V

I7D,V/B

D,B Tight

Accumulator

13%
13%

Reboiler

7%

Accumulator

7%

Reboiler

7%

Accumulator

12%

Reboiler

7%

Accumulator

10%

Reboiler

10%

Accumulator

7%

Reboiler

13%

Accumulator

7%

Reboiler

10%

Accumulator

10%

Reboiler

7%

Accumulator

7%

Reboiler

12%

Accumulator

9%

Reboiler

9%

Accumulator

1%

2.5

Reboiler

1%

2.5

4.1.6 Dynamic Simulation Development


The linear differential equations comprising the C3 splitter model are integrated
using an Explicit Euler Integrator. The time interval for the integration was chosen as 0.3
seconds. The small integration step did not lead to numerical instabiHty and provided
accuracy. The computational time to execute on a 450 MHz PII was around 5 minutes of
CPU time corresponding to the simulation time of approximately 5000 minutes.
26

4.1.7 Linear Model Benchmarking


The Hnear model developed for the C3 spHtter is derived from a rigorous nonlinear model developed by Hurowitz (1998). He validated his C3 spHtter model using
steady state and dynamic data from industrial C3 spHtters, operating with the [L,B]
control configuration. Hence, the Hnear dynamic model for C3 spHtter developed in this
work is benchmarked against the non-Hnear simulator of Hurowitz (1998).
The open-loop responses of the linear model are compared with the open-loop
responses of the non-linear simulator using [L,B] configuration. The open-loop responses
considered are (a) 0.1% step increase in feed flowrate and (b) 0.1% step increase in feed
composition. Figures 4.5 (a) and 4.5(b) show the comparisons of the open-loop responses
of the linear model and the non-linear model. It can be seen that the linear model's
responses match well with that of the non-Hnear model. For the feed flowrate step
increase, the linear model and non-linear model show a little mismatch at steady state.

27

-^ 0.306
"o 0.305 H
.-^ 0.304
E 0.303 H

1^ 0.302 i
o
T3
CD
(U

0.301 0.3

>

0.299
0

1000

2000
Time

3000

4000

^miniitaQ^

Non-linear M o d e l

Linear M o d e l

2.02
o
E
3
Q.

1.98

1.96

Q.

1.94

O
^

CO

E
o
o

1.92 1.9
0

1000

2000

3000

4000

Time (minutes)
Non-linear M o d e l

^ ^ " " L i n e a r Model

Figure 4.5. Comparison of open-loop responses for C3 SpHtter. (a) For 0.1% step increase
in feed flowrate between linear and non-linear model for [L,B] configuration.

28

0.302

0.294
0

1000

3000

2000

4000

Time (minutes)
Non-linear M o d e l

Linear Model

1.96
0

1000

2000

3000

4000

Time (minutes)
Non-linear Model

" " " " " L i n e a r Model

Figure 4.5. Continued, (b) For 0.1% step increase in feed composition between Hnear and
non-Hnear model for [L,B] configuration.

29

4.2 Depropanizer - Multicomponent distillation column


4.2.1 ModeHng Assumptions
Depropanizer is also one of the widely used multicomponent distiUation columns
in the industry. The depropanizer distillation column is structurally identical to the C3
splitter. The assumptions used to develop a model of the depropanizer are:
1. The feed stream is comprised of six components: ethane, propane, /-butane, -butane,
pentane, and hexane.
2. The column is composed of a single-feed stream and two product streams, an
overhead and bottoms product
3. The column pressure is assumed to remain constant at each tray but varies Hnearly
throughout the column.
4. The vapor holdup is negligible.
5. At each tray of the column, the vapor is in equiHbrium with the liquid. The same is
assumed at the reboiler.
6. The vapor Hquid equiHbrium is represented by the Soave-Redlich-Kwong (SRK)
equation of state (Duvall, 1999).
7. Enthalpies are estimated using ideal enthalpy data and enthalpy departure functions
obtained from the SRK equation of state (Duvall, 1999).
8. Trays are assumed to be non-ideal and are modeled using a Murphree tray efficiency
(Murphree, 1925) for all trays except the reboiler. The reboiler is treated as an ideal
stage.
9. The column has a partial reboiler and a total condenser.
10. Tray liquid dynamics are treated using the hydraulic time constant approach
presented by Franks (1972) and Luyben (1990).
11. The heat transfer dynamics of the condenser are not considered. The reboiler heat
transfer dynamics are treated using a first-order lag.
12. Row rates are modeled with first-order lags to represent valve dynamics.
13. Holdup in the reboiler and condenser are designed for 5 minute residence time.
14. Tray temperatures are used to infer overhead and bottom compositions for control.

30

15. The heat transfer dynamics of the tray temperature measurements are modeled using
first-order lags.
16. Sensible heat change at each tray of the column is assumed to be small enough to be
neglected.
17. The liquid levels and compositions at the reboiler and accumulator are controUed
using PI controllers.

4.2.2 Vapor Liquid Equilibrium


The depropanizer separates a multicomponent mixture of ethane, propane, /butane, n-butane, pentane and hexane. The vapor Hquid equiHbrium between these
components can be represented as (Smith et al., 1987),
r>.j=f'i.j

(4.31)

where f^'. represents the fugacity of vapor of component j in the vapor stream leaving
tray i and / . j represents the fugacity of component j in the Hquid stream at tray i.
Expressing the fugacity in terms of a fugacity coefficient, the above equation becomes,
yi.jfi-j = xj'i.j

(4.32)

where 0/. represents the fugacity coefficient of component j in the vapor stream leaving
tray i, yij represents the corresponding vapor mole fraction, 0. ^ represents the fugacity
coefficient of component j in the Hquid stream at tray i and Xij represents the
corresponding liquid mole fraction.
The Hquid and vapor fugacity coefficients are estimated using SRK equation of
state. The expressions for fugacity coefficients obtained using an SRK equation of state
are provided by Walas (1985). Thus, Equation (4.32) can be convenientiy represented as,
yi.j = K^,x,^j

(4.33)

where Ki.j represents the vaporization equilibrium ratio and is defined as,
K..=^.

4.34)

31

For multicomponent systems, relative volatility is defined as.


^i.h
^iM,

(4.35)

f^.J

where a.^^j represents the relative volatihty of component h to component j at tray i.


Relative volatility is a measure of the degree of separation. For the depropanizer, the
relative volatility lies in the range of 1.5-2.0, indicating a fairiy moderate degree of
separation (Duvall, 1999).
The above VLE equations are solved on each tray. Hence, to obtain a Hnear
dynamic model, these equations are linearized at each tray around its steady state
operating condition. Since the SRK equation of state is non-linear and involves many
parameters, analytical Hnearization techniques result in large dimensional equations.
Hence, numerical techniques are used instead which provide linear equations and are
easily employed. The resulting equations can be written as.
/

yij

(^.;l^.;+I

dK 'i
"'^a^

^..+x

A-,. ,.

y55

9^M1
';
'

- J X:

'J

^y.

Jss

dK,, \
dT Jss

T
J

(4.36)
where the partial derivatives are calculated using numerical techniques. The above
resulting equations are linear but solving these equations for each tray requires
considerable amount of computational time (Duvall, 1999). To reduce the amount of
computational time, an inside-out algorithm is used (Boston and SulHvan, 1974). The
inside-out algorithm represents a stable efficient algorithm to solve VLE for
multicomponent distiUation columns. The inside-out algorithm is a modified version of
the KB method. An exponential temperature dependence is assumed for KB.
\n(K,^) = A.

(4.37)
B.i

Using the basic rules of vapor-liquid equiHbrium for a given tray j and i components, it
can be shown that KB is given by,
MK,,)

= J^y,j\n(K,j)

(4.38)

32

To estimate the parameters A and B, the linear vapor-Hquid equilibrium equations


are solved at two temperatures close to the tray temperature, T and T-H AT, AT being
small. For the depropanizer AT was chosen as 0.2PC. Thus, using the above two
equations, the parameters are estimated as,
S)'Mln(/f,;.r.Ar)-Iy,.;ln(^MT)
B,/ =
;
r^
I
1
T T + AT

(4.39)

4.40)

The KB model approximation is assumed to be valid for small changes in tray


temperature. Hence, if the tray temperatures do not vary much, the above equations may
be used to represent the VLE. It was found that for a change of at most 1C, the KB model
may be assumed accurate to represent the rigorous VLE equations. To keep the KB model
accurate, it is updated periodically every 10 seconds and when any of the tray
temperature rises above lC (Duvall, 1999).

4.2.3 Depropanizer Steady State Designs


Four different depropanizer steady state designs developed by Duvall (1999)
were studied. These designs differed from each other with respect to product purities
namely, low, asymmetric, base case, and high purity. The procedure followed to design
these columns can be summarized as:
1. The minimum number of trays, feed tray location and minimum reflux ratio were
obtained using Fenske-Underwood-GilHland procedure (Holland, 1991).
2. The reflux ratio was taken as 1.2 times the minimum reflux (Holland, 1991).
3. A rigorous column design is used to identify the column pressure drop, sub-cooHng
on reflux, Murphree efficiency (Murphree, 1925), and desired product purities.
4. The feed tray location and number of trays were adjusted to minimize the energy
requirements for the reboiler and accumulator.
5. The design parameters for the four different depropanizer columns are shown in
Table 4.3.
33

c
3

CO

x:
too

ON
^

-"^

ON
^

<N

r~; o
CO
ri

ON

CO

NO

m
o
^
CN

<N

ON

rf

ON

o ^ O^ ^
C^
d d o d o

NO

lO

"O

CO

in

in

CO

in

in

in
CO

in

in

< a.

ON
'^

Tl<N

<N

0,
-^

cs d

ON

NO

"

CO

(N

o
c^ o
d <N

CO

00
<N
(N

o
^

oo
n

,204

ON
-^

Tl-

00

613

ON
VD

CO

.409

4>

<N

u^

NO

m
H

in

p
r4

CO

in

p
<N

u
(U
on

C3

PQ

ON
^

'^

<N

ON
"^

m
d

n
d

.25

ID

(/3

<u
4>

(U

ea
OH

<D

top

>>

VM

41

00
Ui

(U

o
Vi

'c
OH

o
V-i

Q
CO

o-!
,*

cier

Vi

_o w
>>
*41

:z;
13
4->

>>
C3
TJ

a
0)>

tin

<D

o
PQ

4>

(u

t^

H
o

8
3

!3:.

X
3
!;:

PC

Qii

. - 1

^ ^

a.
o
PQ

1I

oo

PC
73
3
Vi

O
4-^

(D

00

4>

00

Vi
00

>

PQ

PQ

3^
H

00
Vi

C3

tin

.1.1

Vi

4>

a
Vi

<
u
Vi

o
<u

T3
(U

'o

00

<u

4->
4

41

cd
D

C3

OO
>^

4*

e3
O

U
H

c
o

11
41

.2

o
<u

*4->

00

Vi

too
C

<-i

Oi

o 11

Vi
4->
Vi

u,
O

O
0)

too
3
^ -

o
Vi
Vi

<D

PC
too
3

Vi

Vi

'S

13

<D
N
>^

O
N
>>

3
O

3
O

. " " I

' " ^

00

Vi

o
U

o
U

4.2.4 Linear Modeling


4.2.4.1 Interior Trays of distillation column
Franks (1972) has provided a description about modeling of multicomponent
columns. While a model of depropanizer is similar to the C3 spHtter, the equimolal
overflow assumption cannot be assumed.
The material balance on each tray is given as,
dM
^
= A>.+V;_,-L, -V,

(4.41)

flf(A-,.M,)

J^

= ^,>..;^,>I-^MA +K-,>'/-U

-Viyi.j-

4.42)

While the energy balance is,


^ ^ ! ^ = I,,,,L,,-h,L, +V,_,H,_, -V,H,.

4.43)

Expanding the derivative amounts to,


^ W , +h,^

= h.,,,L.-h,L, + V,_,//,_, -V,H,.

4.44)

The first term of this above equation on the left hand side represents the sensible
heat change on each tray. For the depropanizer, the sensible heat change at each tray is
negHgible and substituting Equation (4.41) into Equation (4.44) yields,
,,

L.., (ft,.,-/,) + l^,(g,-/.,-)


//.-ft,
/
/

Each tray is treated as a non-ideal tray with a Murphree tray efficiency (Murphree, 1925).
y>,=y>-^, + E(yZ-y:-^.,)

(4-46)

Each tray is assumed to be a CST with a hydraulic time constant, T.^, . Hence Equation
(4.41) gives,
^ ^ ^ - " ^ - ' - ^ - ^

(4.47)

The trays in the stripping and rectifying section have different hydraulic time
constants (Duvall, 1999). For the rectifying section, the hydrauHc time constant is 3.5
seconds whereas in the stripping section it is 5.25 seconds.
35

The liquid and vapor enthalpies are obtained from ideal enthalpy data using
departure functions (Duvall, 1999). These departure functions are estimated using SRK
equation of state (Walas, 1985). Thus the enthalpy equations for each component in
Hquid and vapor are given by

ft,,=ft,0+Aft,,.

(4.48)

//,,,=//,+AW,,

(4.49)

where h^ . is the ideal liquid enthalpy component j and A/z,.^ represents the departure of
liquid enthalpy of component j at tray i from an ideal solution. Similar statements can be
made about vapor enthalpy. It should be noted that the multicomponent liquid and vapor
streams are assumed to be ideal solutions. Hence the Hquid and vapor stream enthalpies,
respectively are given by

Hi=J,yi,H,..

4.51)

To obtain a Hnear model for the depropanizer, the non-linear equations at each
tray are linearized around the steady state operating conditions on each tray. The
linearization gives
dL.
dt
dx.i-j

_L,.,+V,_,-Li-V.
r.^^j
^x

dt

-x

(L

^,>.+

M,
^L

(4.52)

Jss

yi.j-^'
^i+1.;

\ ^ ' JSS

^v, ^

y^'jss

^Vi-.'

y>.i +

^MV ^ ' JSS

M,

K. +

M..

y55

V.;-l

Jss

yi-Lj

I Mi ,cc
V
' JSS

^H,_,-h^
Lv+i
V. =
^,-1 +
A>i +
H,-h
"^-^jss
v^'-^'y55
i Jss
r J
\
( v.. \ H._
h:H.-h
^H,-h,
Jss

^y. , . - A ^

i-j

(4.53)
(

K, +

v..

;-l

yli.-K

H i-\
Jss

Jss

(4.54)

36

y,i = y,-.,+E(jZ-y:-uy

4.55)

In the case of the feed tray the non-linear material and energy balance equations are.
(4.56)

dt

tray,I

d{x,jM,)
= XMjEM-^i,j^

dt

(4.57)

+K-l3^/-W - ^ / X . ; +^^F.y

A..(Kx -hi) + Vj.,(//,_. - / t , ) + F(h, -h^)

V: =

H^-k

yi.j = y,-ij +

4.58)

EMv(yLj-yi-iJ-

Correspondingly, linearizing these equations gives


JL..
dt

li^i+V

,-li-V,+J

1-1

(4.59)

tray

dl.

>-j

^i+\.i

^i.j

dt

(L

^,>.+
(

M..

^i+l.j

-x..\

v..-h

v..
y55

755

V. \
yi-Lj+i^F.j

yi.j +

V ^ ' JSS

755

M.

y^'^jss

Jss
^/.y

-^',.y)55^ + (^)55^^-

y'^'jss

(4.60)

v.. =^Kri^
H.-K

/^
-IV

/ / .1-1 , - /"zi ^

A>i +

//.-ft.

755

K/-I
H:-k
'
' JSS

^,-1

^,-, +

r h^ - h.

h:-

hi+\

K"'-''-JSS

//.+
y">-hips

yHi-hijss

"i+l

F+

yHi-h^jss

Jss

/i,

. tli - h: ,
V

'

'JSS

(4.61)
4.62)

>'/.; = > ' / - l , ; + ^ w ( K ! - > ' / - l . ; )

4.2.4.2 Accumulator
For the accumulator, the material balance equations are
dM

(4.63)

^'^^ =V -L -D
^ N
^R
^
dt

dix^cc^ACc)
dt

_
= VA^;, - (L,, + D)x^^^.
37

4.64)

The accumulator is assumed to be an ideal stage. Since the cooling water is not
considered a part of the distillation column, the energy balance equation is not
considered. The linear model for the accumulator is given by,
dM ACC

dt
d(^,cc)
dt

V,-L,-D
( y^N

(4.65)
(

.rr JSS
V M ^^^

yN-

M ACC

4.66)

'ACC

Jss

4.2.4.3 Reboiler
For the reboiler, the material and energy balance equations are
dM REB

dt

L,-V,-B

(4.67)

dJ^REB^REB)
^IM

dt
Q+
^0

yREB*^0

L,{h,-h^^)-M REB

(4.68)

-^REB"

dh,"REB
dt

(4.69)

-h

^^ REB

''^REB

The reboiler is assumed to be an ideal stage. The resulting Hnear model is given
by
dM REB

= L,-V,-B

dt

(4.70)
\

" \XREB

dt

-^1

^ L
L,M

XREB

M REB
M REB

755

y^REB

^
A,H
JSS

yREB

-^REB

M REB

KJss

"REB
y "^ REB

Jss

yREB

Jss

(4.71)

4.2.5 Depropanizer Level Controllers


The level controllers are addressed in the same way as the C3 spHtter. The level
control loops are tuned for overdamped behavior (MarHn, 1995). Equations (4.29) and
(4.30) are used for estimating the controller tuning parameters. AFmax is chosen as 5% of
38

the reflux flowrate. Different ALmax and ^ are chosen heuristically for different
configurations by taking into consideration the interactions between the two loops. Table
4.4 shows the values of A Lmax and ^ selected for the reboiler and accumulator level
control for different configurations in the four depropanizer designs. For the asymmetric
purity design A Lmax was chosen as twice that shown in the table. Since the depropanizer
is more non-linear as compared to C3 splitter, composition control is more difficult.
Hence larger values of ^ were chosen for depropanizer to dampen the level control loops
and reduce the interaction with composition control loops. The results obtained for
various values of ALmax and ^ are compared with the results obtained by Duvall (1999).

4.2.6 Inferential Composition Control


The analyzer sampling period for depropanizer was one minute. To improve
composition control, infkerential control was used, i.e., the composition is inferred using
a particular tray temperature. Riggs (1998) has provided a simple technique to identify
the best tray for estimating product composition. The different trays used for inferring
product compositions are provided in Table 4.3 for each design. The basic correlation
used for inferring compositions from tray temperature is
(4.72)

ln(A) = A +

where x is product composition estimated using tray temperature measurement T.


For implementing inferential composition control, on-line analyzers were used to
update the parameters used in Equation (4.72) based on previous analyzer readings and
temperature measurements. Parameter updates were carried out at each sampHng time.
Using these updated parameters, past analyzer measurement and past temperature
measurement, the composition is inferred as
/

Ax = X old

exp B

1
T

1
Told

(4.73)

>-l

(4.74)

x = x^+Ax

39

where Xoid and Tdd are analyzer and temperature measurements at recent past sampling
times.

Table 4.4 ALmax and ^ chosen for calculating the level controller tuning parameters for
C3 spHtter
Configuration
Level Controller A Lmax
^
L,B

Accumulator

25%
25%

15

Reboiler

10%

10

Accumulator

10%

10

Reboiler

10%

10

Accumulator

20%

15

Reboiler

10%

10

Accumulator

15%

15

Reboiler

15%

15

Accumulator

10%

10

Reboiler

25%

15

Accumulator

10%

15

Reboiler

15%

15

Accumulator

15%
15%

15

Reboiler

10%

15

L/D,V

Accumulator
Accumulator
Reboiler

10%
10%
20%

10
15

L/D,V/B

Accumulator

13%
13%

13

Reboiler

13%

13

L,V

L,V/B

D,B

D,V

D,V/B

L/D,B

4.2.7 Dynamic Simulation Development


The Hnear differential equations comprising the depropanizer model are
integrated using an expHcit Euler Integrator. The time interval for integration was chosen
as 0.5 seconds. A small integration step was chosen to achieve numerical stabiHty and
40

accuracy. The computational time on a 450MHz PII was approximately 30 seconds of


real time corresponding to a simulation time of 400 minutes.

4.2.8 Linear Model Benchmarking


Analogous to the C3 spHtter, the Hnear model for the depropanizer was adopted
from Duvall (1999). In that work, a rigorous non-linear model was developed for
depropanizer based on industrial depropanizers operating with an [L,B] configuration. He
benchmarked his non-Hnear model against the steady state and dynamic data of the
industrial columns by adjusting the efficiency of the trays and the hydraulic time constant
for the stripping and rectifying sections. To obtain an industrially benchmarked linear
model of the depropanizer, this Hnear dynamic model is benchmarked against this nonHnear model. This is accompHshed by comparing the open loop responses of the nonlinear and Hnear models. The open loop tests considered are (a) 0.1% step increase in
feed flowrate and (b) 0.1% step increase in the propane feed composition. These
responses are shown in Figure 4.6. From the responses it is observed that the dynamics
between the non-linear and Hnear model are identical. However there is some mismatch
between the Hnear and non-Hnear models at steady state.

41

0.58

0.48
300

400

500

Time (m inutes)
Non-linear Model

200

"

'Linear Model

300

500

Time (minutes)
Non-linear Model

" "

'Linear M o d e l

Figure 4.6. Comparison of open-loop responses for depropanizer. (a) For 0.1% step
increase in feed flowrate between Hnear and non-linear model for [L,B] configuration.

42

0.51

100

200

300

400

500

T i m a / m in ii t<sc \

Non-linear M o d e l

^ ^ ^ L i n e a r Model

0.49
0

100

200

300

400

500

Time ( m i n u t e s t
Non-linear M o d e l

^ ^ " " L i n e a r Model

Figure 4.6. Continued, (b) For 0.1% step increase in feed flowrate between Hnear and
non-Hnear model for [L,B] configuration.

43

CHAPTER 5
DUAL-ENDED COMPOSITION CONTROL

Research (Luyben, 1975; Chiang and Luyben, 1985; Ryskamp, 1980; Stanley and
McAvoy, 1985) has shown that dual-ended composition control provides significant
reduction in energy consumption and better control as compared to single ended control.
Hence, for analyzing distillation column operation, dual-ended composition control is
considered. Closed-loop Bode plots for feed composition disturbance are generated for
the C3 splitter and the depropanizer. The Bode plots are obtained from simulations of the
linear dynamic models of the C3 spHtter and the depropanizer.
In section 1, different configurations for distillation control are discussed and the
issue of configuration selection is described. The utiHties of using an invariant structure
of a distillation column are presented in section 2. The tuning criteria used to tune the
composition controllers is discussed in section 3. Comparison of the results for dualended composition control between the Hnear model and non-Hnear model is carried out
in section 4. The closed-loop Bode plots for feed composition disturbance for various
configurations are also shown in section 4.

5.1 Configuration Selection


From a process control perspective, a distillation column essentially needs five
variables to be controlled to maintain continued operation. These variables are: pressure,
reboiler level, accumulator level, overhead composition, and botto
ms composition. These variables can be controlled by choosing any of the following
independent variables: reflux, distiUate, bottoms, and vapor boilup (reboiler duty) flow
rates and condenser duty. Thus, it makes the distillation problem a [5x5] control problem
(Duvall, 1999; Hurowitz, 1998). For C3 spHtter and depropanizer, the column pressure is
usually allowed to float, thereby cHminating one controlled variable and hence one
manipulated variable (condenser duty). This reduces the distillation problem to just two
composition control loops and two level control loops. In addition, reflux ratio and boilup
ratios can be used to replace any of the overhead and bottoms manipulated variables.
44

The procedure of pairing controlled variables to manipulated variables is called


loop pairing. Assuming that the top manipulated variables are not used to control bottom
controlled variables and vice versa, only nine configurations are left possible. These nine
configurations are Hsted in Table 5.1. The ratio configurations are implemented using
Ryskamp (1980) ratio control arrangement. Configuration selection is very crucial for
achieving good distillation control. Configuration selection problem inherently addresses
the problem of coupHng between the loops.

Table 5.1 Controlled and Manipulated Variable pairings for dual PI composition control

L,B

Overhead
Composition
L

Bottoms
Composition
B

Accumulator
Level
D

Reboiler
Level
V

L,V

L,V/B

V/B

V+B

D,B

D,V

D,V/B

V/B

V-HB

L/D,B

L/D

L-hD

L/D,V

L/D

L-hD

L/D,V/B

L/D

V/B

L-t-D

V-fB

Configuration

Several researchers (Shinsky, 1984; Skogestad and Morari, 1987; Skogestad et al.,
1990a) have proposed guidcHnes to faciHtate indentification of the best configuration. In
regards to C3 spHtter, Gokhale (1994), Finco et al. (1989), and Hurowitz (1998),
compared the different configurations and analyzed tiieir control performance. In the case
of the depropanizer, Duvall (1999) carried out an extensive analysis of its different
control configurations.
5.2 Invariant Structure of a DistiUation Column
One of the most important problems associated with distillation conU-ol is
configuration selection. It has been shown that choosing a reasonable but inferior
45

configuration can result in an order of magnitude higher variabihty in products for


distillation columns (Anderson, 1998; Hurowitz, 1998). The use of the relative gain array
and other configuration selection statistics has been shown not to correlate well with
configuration control performance, even if accurate Hnear models of the column are
available and they are not generally available (CarHng and Wood, 1988; Anderson, 1998;
Hurowitz, 1998; Duvall, 1999). It seems reasonable to use the method proposed in
Chapter 3 for calculating product variabihty to identify the optimum control
configuration.
Figure 4.3 shows an invariant structure for a distillation column that can facilitate
these calculations. In this structure, we assume that the distillation column is made up of
three separate entities namely, the accumulator, reboiler and the interior trays of the
distillation column (Yang et al., 1990). Note that regardless of which configuration is
used, the same model will apply. For example, the (L,V) configuration would use the
distillate and bottoms flowrates to control the accumulator and reboiler levels,
respectively, while the reflux and boilup would be used to control the product
compositions. For the (L/D, B) configuration, the sum of (L-HD) and vapor boilup would
be used to control the accumulator and reboiler levels respectively, while L/D and B
would be manipulated to control the product compositions. As a result, the same
Hnearized model could be used regardless of the configuration chosen which greatly
simpHfies the modcHng problem. This invariant structure was utilized in developing the
Hnear dynamic models of C3 spHtter and depropanizer.

5.3 Composition ControUer Tuning Criteria


The aim of tuning the controllers is to achieve stable conttol with minimum
variation from setpoint. A common tuning approach is developed for both the Q spHtter
and the depropanizer to compare results among different configurations. C3 spHtter and
depropanizer are slow response loops. Traditional open-loop tuning methods such as step
tests cannot be appHed on the depropanizer and C3 spHtter because tiiey have very long
response times. In addition, since tiiese step tests are lengthy, they may be significantiy
affected by unmeasured disturbances and asymmetric dynamics of the process. Hence,
46

an ATV test is used to identify the process parameters, which can be run online without
significantiy varying the process.
The initial tuning parameters for the composition control loops are obtained by
treating both loops to be independent of one another. First the ultimate gain and period of
both loops are obtained independently using ATV tests (Astrom and Hagglund, 1991)
Tyreus and Luyben (TL) settings are then used to derive initial tuning settings, which are
given as

^^

^''''~

0.45

^^ ^ ' ^ ^ "

0.45

^^^^

where the subscripts TOP and BOT correspond to the top and bottom composition loops,
superscript TL refers to TL settings and Ku and Pu correspond to ultimate gain and period.
Applying the above settings may result in suboptimal control because they do not
account for the coupling between the two loops. Hence, the tuning parameters need to be
detuned. This is achieved by using a common on-line detuning factor for both loops.
Thus the new tuning settings are given by

(Kc)roR =^^4^

^^OBOT

(h )TOP = (r'')TOpXFo

=^^4^

(h )BOT = (h'')BOT

^^O

^^'^'^^
(5-2.2)

where FD represents the detuning factor.


The detuning factor is determined on-Hne by repeated simulation in a similar
approach to the one used by Finco (1987). The detuning factor is tuned for minimum
integral of the absolute value of error (lAE) in controller response for overhead and
bottoms impurity setpoint changes.

47

5.4 Composition Control Results


5.4.1 Base Case C3 SpHtter
The base case design of the C3 spHtter was analyzed for dual PI composition
control for nine configurations listed in Table 5.1. The level controller tuning parameters
for the base case C3 spHtter were obtained using Equations (4.29) and (4.30) and Table
4.2. The TL tuning settings were obtained independently for overhead and bottoms
composition loops using an ATV test. These settings were then detuned using a common
detuning factor as given by Equations (4.29) and (4.30). The detuning factor was tuned
for minimum integral of the absolute value of error (LAE) in controller response for
setpoint changes. For the base case C3 spHtter, since the overhead product is more
important than the bottoms product, the detuning factor is tuned for minimum LAE in the
overhead impurity response for an overhead impurity setpoint change. The overhead
impurity setpoint change was carried out from 0.3% to 0.25% at t = 100 minutes and
from 0.25% to 0.35% at t = 1000 minutes and the simulation stopped at t = 2000 minutes.
The above procedure was repeated for the nine configurations. The TL settings and
detuning factors obtained for the different configurations in the base case design of C3
splitter are shown in Table 5.2
Previously dual PI composition control of the base case C3 spHtter has been
carried out by Hurowitz et al. (1998). He used a non-Hnear dynamic tray-to-tray model
for his analysis. He tuned the composition control loops for minimum lAE for setpoint
changes in a similar manner to the one adopted here. The TL settings obtained in Table
5.2 compared well with those obtained by Hurowitz (1998). However the detuning
factors did not match well with those generated by Hurowitz. The tuning factors obtained
here were aggressive as compared to those obtained by Hurowitz. The difference between
the tuning used here and that of Hurowitz, is that Hurowitz used larger setpoint changes
to tune the composition controllers. The Hnear dynamic model of C3 spHtter used here,
the range of setpoint changes did not affect the tuning factors. For higher setpoint
changes, the non-Hnear C3 spHtter model resulted in highly non-Hnear behavior, which
makes the Hnear model inaccurate. Hence there is a mismatch between the tuning results
obtained here and those of Hurowitz. Due to the highly non-linear behavior of the C3
48

splitter, Hurowitz obtained conservative settings to provide the best control performance.
The degree of mismatch between the linear and non-Hnear model can be seen from Figure
5.1 which shows the comparison between the linear and non-linear model for [D,V]
configuration for a setpoint change of overhead impurity from 0.3% to 0.15% at t=100
minutes and then to 0.45% at t=1000 minutes. For these results the tuning parameters
generated by Hurowitz are used for both linear and the non-linear model.

Table 5.2 Base case C3 splitter dual PI composition controller tuning parameters
Overhead Loop
Bottoms Loop
Configuration

Gain

L,B

Gain

-387.28

Integral
Time
5333.33

5.05

Integral
Time
16000

Detuning
Factor
0.8

L,V

-305.06

8000

-8.59

10666.67

1.1

L,V/B

-277.47

6666.67

-1660.14

9333.33

0.8

D,B

234.05

13333.33

9.75

13333.33

0.9

D,V

125.91

16000

-12.23

6666.67

0.5

D,V/B

204.45

13333.33

-2377.13

6666.67

0.7

L/D,B

-12667-4

5333.33

8.46

13333.33

L/D,V

-8598.05

8000

-12.54

6666.67

1.1

L/D,V/B

-9907

6666.67

-2419.2

6666.67

D,B Tight

335.24

8000

9.87

10666.67

49

200

400

600

800

1000

1200

1400

1600

1800

2000

1800

2000

Time (minutes)
Non-linear Model

'Linear Model

SP

_0

o
E

3
O.

E
o
c
ra
a
o
w
Q.

E
o
o
m
0

200

400

600

800

1000

1200

1400

1600

Time (minutes)
Non-linear Model

Linear Model

SP

Figure 5.1 Comparison between the closed-loop responses of the Hnear and non-Hnear
model for setpoint tracking of the [D,V] configuration of the Base case C3 Splitter
5.4.1.1 Setpoint Control Results
The lAE control performance measures for the nine configurations for setpoint
tracking are shown in Table 5.3. The setpoint changes are the same as the ones used to
50

select the detuning factor. The units of the lAE indices are [impurity mole fractionjx
[secj. It can be seen from Table 5.3 that [L/D,V] configuration provided the best
overhead loop lAE performance with an lAE of 4.58. [L,V] and [D,B1 performed worst
in the overhead loop with lAE values of 10.99 and 10.41, respectively. The overhead
loop performance of [L,B], [D,B] (with tight level control) and [L/D,B] configurations
was comparable to that of [L/D,V1. The overhead loop performance of [L,V/B] and
[L/D,V/B] configurations was also satisfactory with LAE values of 6.10 and 6.21,
respectively. For the bottoms loop, the performance of the [L,V/B] configuration with an
lAE of 55.56 was the best while the [L,V] performed the worst with an lAE of 460.63.
The next best control performance was provided by [D,B] (with tight level control) with
an lAE of 71.51. The [L,B],[L/D,V/B], [D,V/B] and [D,V] configurations provided
reasonable bottoms loop performance with lAE values around 100. [L/D,V] which
provided the best overhead loop performance fared poorly for bottoms loop performance
with an lAE of 245.88.

Table 5.3 Base case C3 Splitter LAE indices for overhead impurity setpoint control
Configuration
Bottoms Loop LAE
Overhead Loop LAE
_

L,B

106.81

L,V

460.63

10.99

L,V/B

55.56

6.21

D,B

227.09

10.41

D,V

125.05

7.36

D,V/B

115.69

8.24

L/D,B

162.84

5.61

L/D,V

245.88

4.58

L/D,V/B

106.80

6.10

D , B Tight

71.51

5.72

51

5.4.1.2 Closed-Loop Bode Plots


The closed-loop Bode plots of overhead and bottoms impurity for sinusoidal feed
composition disturbance were generated using the Hnear dynamic model of C3 spHtter
developed in Chapter 4. For generating closed-loop Bode plots, sinusoidal feed
composition disturbances at various frequencies between 0.1 and 0.001 rad/min are input
to the linear dynanuc model. From the resulting steady state responses, amplitude ratios
and phase angle shifts were calculated to obtain the closed-loop Bode plots. The
frequencies were chosen between 0.1 and 0.001 rad/min as the ampHtude ratios outside
this range were negHgible. The amplitude of each feed composition disturbance was
chosen as 0.1% of the propylene feed composition. The closed-loop Bode plots of [L,B]
and [L/D,V] configurations are shown in Figures 5.2 and 5.3, respectively.

\ I'

200

-i^

1
}

100

0.1 I

K-

CD

tr

0.01 {

' 1
=t

TJ

0.001 M*'

s:

-100 s.

^KHI,

7 - - ^

0^

^^%^

-200 w

0)

Vi

<

-300.2
f\ A A A I

0.0001 j

-400

^ 1

-500
0.01

0.001

0.1

Disturbance Frequency (rad/minute)


'Overhead AR

Bottoms AR

'Overhead Phase

Bottoms Phase

Figure 5.2 Base case C3 Splitter Closed-Loop Bode plot for feed.composition disturbance
rejection for [L,B] configuration

52

I I I

300

:::::::^:::^

200
0.1

100

V)
<D

0)
(C

a:
0.01

O)
0

-100
CO

<

--- ^
0.001

^^^ B 4 B ^ ^ ^ ^ ~

B
,.

~'

i i ^
"

-200

(/)

-300

.c
CL

(U

(C

^^^^p^^ I

-400

_ _ _ ^

-500

0.0001
0.001

0.01

0.1

Disturbance Frequency (rad/minute)


'Overhead AR

Bottoms AR

'Overhead Phase

' Bottoms Phase

Figure 5.3 Base case C3 Splitter Closed-Loop Bode plot for feed composition disturbance
rejection for [L/D,V] configuration
The overhead impurity amplitude ratio plots for feed composition disturbance
rejection for the nine configurations are shown in Figure 5.4. From the analysis of the
ampHtude ratio plots, it can be seen that the [D,B] (with tight level control) configuration
provided the best unmeasured disturbance rejection performance for overhead impurity.
The [L/D,V] configuration provided the next best control performance for the overhead
loop. [L/D,B], [L/D,V/D], and [D,V] configurations provided reasonable overhead
control performance with each of the configurations peaking at around 0.02 rad/min (314
min disturbance period). The [L/D,V] configuration, however, exhibited a blunt peak at
around 0.03 rad/min (209 min disturbance period). The [D,B] configuration preformed
poorly at low frequencies but performed reasonably well at high frequencies. In contrast,
the [L,V] configuration fared well for low frequencies but resulted in poor overhead
impurity control performance at high frequencies. The performance at low frequencies is
a measure of steady state behavior and the performance at high frequencies is a measure
of the initial response behavior. It means that the [D,B] configuration is better off
rejecting disturbances which have a period of a day or greater than the [L,V]
configuration. In addition [L,V] and [D,B] configurations exhibited another peak at high
53

frequencies. The overhead impurity control performance for disturbance rejection was
relatively poor for the [L,V/B] configuration as compared to the remaining
configurations. The [L,B] configuration performed better than the [L,V/B] configuration.
Both configurations exhibited analogous overhead impurity control performance and
exhibited peaks at around 0.011 rad/min (571 min disturbance time).
The bottoms impurity ampHtude ratio plots for feed composition disturbance
rejection are shown in Figure 5.5. From Figure 5.5, it can be seen that the [L,V/B]
configuration provided the best overall feed composition disturbance rejection for
bottoms impurity. It exhibited a blunt peak at around 0.02 rad/min (314 min disturbance
period). The [D,V], [D,V/B], and [L/D,V/B] configurations provided reasonable
disturbance rejection performance. Their ampHtude ratio plots were analogous and each
exhibited a large peak at around 0.02 rad/min (314 min disturbance period) followed by a
small peak at around 0.07 rad/min (89 min disturbance period). Bottoms impurity control
performance of [D,B] and [L/D,B] configurations was analogous, with both exhibiting a
peak at around 0.015 rad/min (418 min disturbance period). The [D,B] (with tight level
control) configuration performed better than these configurations. [L,V] and [L/D,V]
configurations provided extremely poor bottoms impurity control performance as
compared to all other configurations.
The ampHtude ratio plots of overhead and bottoms impurity for sinusoidal feed
composition disturbance rejection provided lesser ampHtude ratios than those obtained by
Hurowitz (1998) for most of the configurations. But the nature of the closed-loop Bode
plots was analogous to tiiose obtained by Hurowitz. He also found that the [D,B] (with
tight level control) and [L/D,V] configurations provided the best overhead impurity
control performance while the [L,V/B] configuration provided the best bottoms impurity
control.

54

0.012
(0

tr

0.01

0)
T3
3

. ^

t 0.008
E

^'

<

'' ' /

0.006

I.

a
- 0.004

(0
0)

>

0.002

1 g

^p^**^i^

:":-)

0 ^
0.001

0.01

0.1

Disturbance Frequency (rad/minute)


-

-[UB]

[UV]

[L,V/B]

[D.B]

[D,V]

(a) For [L,B],[L,V],[L,V/B],[D,B], and [D,V] configurations.


0.012

Disturbance Frequency (rad/min)


- - - [D,V/B]

tL/D,Bl

IL/D,V]

lUD.V/B]

[D.B] (Tight)

(b) For [D,V/B],[L,D/B],[L/D,V],[L/D,V/B] and [D,B] (with tight level control)


configurations.
Figure 5.4 Base case C3 SpHtter Overhead Impurity AmpHtude ratio plots for feed
composition disturbance rejection.

55

0.35
r

CO

CC

0.3

pl tud

0.25

0.2

<

' T

:/ /.>r-v A ''

>^

L.

Q.

0.15

E
(0

\ *

0.1

o
o 0.05

-0

'

'

'

x/V X

? J<

CO

0.001

0.01

0.1

Disturbance Frequency (rad/minute)

- IL.B1

[L,V/B]

"[UV]

[D,V]

[D,B]

0.35
0.3

0.25

a
E

0.2

<

,_-1

S 0.15 I
Q.

0.1

. -

. ^^^^

1 .' ^ ^

*^V
X.
v^
X
^ ^ ^
X

%,x^*^^

I 0.05

CQ

tf

J ^^^* m

-'

U^

s;.i^

0.01

0.001

T/l

Q:

(a) For [L,B],[L,V],[L,V/B],[D,B], and [D,V] configurations.

0.1

Disturbance Frequency (rad/min)


- -

-[D,V/B]

[L/D,B]

[L^D.Vl [LyD,V/B]

tD.B] (Tight)

(b) For [D,V/B],[L,D/B],[L/D,V],[L/D,V/B] and [D,B] (with tight level control)


configurations.
Figure 5.5 Base case C3 splitter Bottoms Impurity Amplitude ratio plots for feed
composition disturbance rejection.

56

5.4.2 High Purity C. Splitter


The high purity design of the C3 spHtter was analyzed for dual PI composition
control for all nine configurations listed in Table 5.1. The level and composition
controller tuning parameters were obtained in a manner similar to that used for the base
case design of the C3 spHtter. For the high purity C3 spHtter both overhead and bottoms
product are given equal priorities. Hence, setpoint changes in both overhead and bottoms
impurities were used to tune the detuning factor. The overhead impurity setpoint was
changed from 1% to 1.2% at t=100 minutes and then to 0.8% at t=1500 minutes. This
was followed by a setpoint change in bottoms impurity from 1% to 1.2% at t=3000
minutes and then to 0.8% at t=4500 minutes. The simulation was stopped at t= 6000
minutes. The tuning factor was tuned for minimum lAE in overhead and bottoms
impurities for the above set of setpoint changes. The TL settings and detuning factors
obtained for the different configuration for the high purity design of C3 splitter are shown
in Table 5.4. The TL settings obtained in Table 5.4 compared well with the results
obtained by Hurowitz (1998) using a non-Hnear dynamic model of the high purity C3
spHtter. However the detuning factors did not match well with those generated by
Hurowitz analogous to the base case C3 spHtter. The tuning factors obtained here were
aggressive as compared to those obtained by Hurowitz. The mismatch between the Hnear
and non-Hnear model for high setpoint changes resulted in the difference in the detuning
factors. The mismatch between the Hnear and non-Hnear model of the high purity design
of C3 splitter can be seen from Figure 5.6. Figure 5.6 shows the comparison between the
Hnear and non-linear model for [D,V] configuration. For these results the tuning
parameters generated by Hurowitz were used for both linear and non-linear model.

5.4.2.1 Setpoint Control Results


The LAE control performance indices for the nine configurations for setpoint
tracking are shown in Table 5.5. The setpoint changes are the same as the ones used for
tuning the detuning factor. The units of the lAE indices are [impurity mole fraction]x
[sec]. It can be seen from Table 5.5 that [L/D,V] and [D,B] (with tight level control)
configurations provided the best overhead loop LAE performance with lAE values of 2.30
57

and 2.38, respectively. The overhead impurity control performance of [L,B], [L/D,B] and
[L/D,V/B] configurations was also good with lAE values of 2.88, 2.70, and 2.86,
respectively. The [L,V] and [D,B] performed worst in the overhead loop lAE
performance with LAE values of 4.96 and 5.11, respectively.

Table 5.4 High Purity C3 spHtter dual PI composition controller tuning parameters
Overhead Loop
Bottoms Loop
Configuration

Gain

L,B

Gain

-766.62

Integral
Time
6666.67

L,V

-916.07

L,V/B

63.65

Integral
Time
18666.67

Detuning
Factor
0.5

8000

-146.56

10666.67

1.2

-775.21

6666.67

-29666.8

9333.33

0.9

D,B

647.89

13333.33

114.13

16000

D,V

372.75

16000

-197.53

6666.67

0.6

D,V/B

573.21

13333.33

-38016.5

6666.67

0.8

L/D,B

-26580.1

6666.67

135.14

13333.33

0.8

L/D,V

-34073.4

6666.67

-203.59

6666.67

L/D,V/B

-27099.7

6666.67

-39689.7

6666.67

1.1

D , B Tight

925.64

8000

158.52

10666.67

1.3

For the bottoms loop, the lAE performance of the [L,V/B] configuration gave the
best result with an LAE of 6.17- The [L/D,V] configuration which performed best for the
overhead loop performed worst with an LAE of 43.56. The [D,B] configuration provided
an lAE 7.07 comparable to that of [L,V/B]. The [L,B] [D,V/B] and [L/D,V/B]
configurations provided good bottoms impurity control performance with lAE values of
11.57, 10.31 and 9.64, respectively. The [L,V] configuration, which performed poorly for
overhead loop, also performed poorly for the bottoms loop with an LAE of 34.81. The
[D,B] (with tight level control) and [L,V/B] configuration which performed well for both
overhead and bottoms loop provided the best total lAE performance with lAE values of
9.45 and 9.67, respectively.

58

1000

2000

3000

4000

5000

6000

Time (minutes)
Non-linear Model

1000

2000

'Linear Model

3000

SP

4000

5000

6000

Time (minutes)
Non-linear Model

'Linear Model

SP

Figure 5.6 Comparison between the closed-loop responses of the Hnear and non-Hnear
model for setpoint tracking of tiie [D,V] configuration of tiie High Purity C3 Splitter

59

The total LAE performance was somewhat misleading because some of the
configurations performed well for one loop but poorly for the other. For example the
[L/D,V] configuration which performed best for the overhead loop but worst for the
bottoms loop resulted in the worst total LAE performance with an LAE of 45.86. On the
other hand the [L,V] configuration which performed poorly for both loops resulted in
poor total LAE performance with an LAE of 39.76, better than that of [L/D,V]
configuration. The [L/D,V/B] configuration performed reasonably well with an LAE of
12.5.

Table 5.5 High Purity C3 SpHtter LAE indices for overhead impurity setpoint control
Configuration Bottoms Loop LAE
Overhead Loop LAE Total LAE
L,B

11.57

2.88

i4!45

L,V

34.81

4.96

39.76

L,V/B

6.17

3.50

9.67

D,B

24.27

5.11

29.38

D,V

13.73

3.76

17.49

D,V/B

10.31

3.87

14.19

L/D,B

12.85

2.70

15.55

L/D,V

43.56

2.30

45.86

L/D,V/B

9.64

2.86

12.50

D , B Tight

7.07

2.38

9.45

5.4.2.2 Closed-Loop Bode Plots


The closed-loop Bode plots of overhead and bottoms impurity for sinusoidal feed
composition disturbance were generated in a manner similar to those generated for the
base case C3 spHtter. The closed-loop Bode plots of [L,V] and [L/D,V/B] configurations
are shown in Figures 5.7 and 5.8, respectively.
The overhead impurity ampHtude ratio plots for feed composition disturbance
rejection for all nine configurations are shown in Figure 5.9. Looking at the ampHtude
ratio plots it can be seen that the best overhead impurity control performance is provided
60

by [LyD,V] and [D,B] (with tight level control) configurations. The [L/D,V/B] and
[L/D,B] configurations provided next best overhead impurity control performance.
Bothse configuration exhibited peaks at around 0.01 rad/min (628 min disturbance
period). The overhead impurity control performance of [L,B] and [D,V] configurations
was equivalent over the whole frequency range. Both configurations exhibited peaks at
0.011 rad/min (571 min disturbance period). The [L,V] configuration provided poor
control performance at lower frequencies but provided good control performance at
higher frequencies. It exhibited two peaks, one each at 0.005 rad/min (1257 min
disturbance period) and 0.05 rad/min (126 min disturbance period). The [L,V/B]
configuration resulted in the worst overall overhead impurity control performance. It
however performed better than [L,V] configuration for frequencies below 0.05 rad/min.
The [D,V] configuration too resulted in poor overhead impurity control.
The bottoms impurity ampHtude ratio plots for feed composition disturbance
rejection are shown in Figure 5.10. From Figure 5.10 it can be seen that [L,V/B]
configuration, which provided the worst overhead impurity control, resulted in the best
bottoms impurity control performance for feed composition disturbance rejection. The
[L/D,V/B] and [D,V/B] configurations too provided excellent bottoms impurity control
performance comparable to that of [L,V/B] configuration. The [L,B] configuration
performed better than [D,B] configuration. The [L/D,V] configuration resulted in the
worst control performance for bottoms impurity. The control performance of [L,V]
configuration was analogous to that of [L/D,V] but it consistentiy provided lesser
ampHtude ratios than [L/D,V] configuration. Both of these configurations exhibited peaks
at around 0.05 rad/min (126 min disturbance period).
Again, the ampHtude ratio plots of overhead and bottoms impurity for sinusoidal
feed composition disturbance rejection showed lesser ampHtude ratios than those
obtained by Hurowitz (1998) for most of the configurations. But the nature of the closedloop Bode plots was analogous to those obtained by Hurowitz. He too found that [D,B]
(with tight level control) and P-,V/B] configurations provided the best control
performance for the overhead and bottoms impurity, respectively.

61

100

r""

0.1

0)
-1

0)

'

-100
CU

0.01

<u

O)
0)

0)
D

Q.

-200
CO
0)

0.001
-300

<

(C

.c

CL

0.0001

-400

0.00001

-500

0.001

0.01

0.1

Disturbance Frequency (rad/minute)


'Overhead AR

Bottoms AR

'Overhead Phase

Bottoms Phase

Figure 5.7 High Purity C3 Splitter Closed-Loop Bode plot for feed composition
disturbance rejection for [L,V] configuration.
0.1

300

..\

200
0.01

100 2
0

cr
0)
_3

0.001

-100
CO

"Q.

-200 <K

<

(D

0.0001

-300

__^

-400
-500

0.00001
0.01

0.001

0.1

Disturbance Frequency (rad/minute)


'Overhead AR

Bottoms AR

'Overhead Phase

Bottoms Phase

Figure 5.8 High Purity C3 SpHtter Closed-Loop Bode plot for feed composition
disturbance rejection for [L/D,V/B] configuration.

62

0.005

"T!

(D

">"

^Miw^^^^

T'

I*

turn

^^^^,

Cd

0) 0.004

"D
3

X ^ ^ ^

Q.

0.003

I 0.002

- j/^ : : ;y^'>rNr^K^.

- -

*^^^^^

^^!?^ytwA.

I 0.001
0)

>

0
0.01

0.001

0.1

Disturbance Frequency (rad/min)


- -tUB]

[UV] -

'[L,V/B]

[D,B]

[D.V]

(a) For [L,B],[L,V],[L,V/B],[D,B], [D,V], and [D,B] (with tight level control)
configurations.
0.005 r

0.01

0.001

0.1

Disturbance Frequency (rad/min)


- - - [D,V/B]

[LyD,B]

[LVD.V]

ItVD.V/Bl

[D.B] (Tight)

(b) For [D,V/B],[L,D/B],[L/D,V],[L/D,V/B] and [D,B] (with tight level control)


configurations.
Figure 5.9 High Purity C3 SpHtter Overhead Impurity Amplitude ratio plots for feed
composition disturbance rejection.

63

0.04 \
CD

CU
0)

o 0.03
3

"5.
E
<
> 0.02
3

a
E
(0

E
o
o

0.01

CD

0
0.001

0.01

0.1

Disturbance Frequency (rad/min)


-

-[UBl

[L.V]

[L,V/B]

[D.B]

[D.V]

(a) For [L,B],[L,V],[L,V/B],[D,B], ],[D,V] and [D,B] (with tight level control)
configurations.

Disturbance Frequency (rad/min)


-

- - [D,V/B]

[UD,B]

[UOy]

[UD.V/Bl * [ D , B ] (Tight)

(b) For [D,V/B],[L,D/B],[L/D,V],[L/D,V/B] and [D,B] (with tight level control)


configurations.
Figure 5.10 High Purity C3 Splitter Bottoms Impurity AmpHtude ratio plots for feed
composition disturbance rejection.
64

5.4.3 Low Purity C3 Splitter


The low purity design of C3 splitter was analyzed for dual PI composition control
for all nine configurations provided in Table 5.1. The level and composition controller
tuning parameters were obtained in a manner similar to the one used for the base case
design of C3 spHtter. For the low purity C3 spHtter both overhead and bottoms product are
accorded equal priorities. Hence the setpoint changes in both overhead and bottoms
product are used to tune the detuning factor. The overhead impurity setpoint was changed
from 2% to 2.5% at t=100 minutes and then to 1.5% at t=1500 minutes. This was
followed by a setpoint change in bottoms impurity from 2% to 2.5% at t=3000 minutes
and then to 1.5% at t=4500 minutes. The simulation was stopped at t= 6000 minutes. The
tuning factor was tuned for minimum LAE in overhead and bottoms impurities for the
above set of setpoint changes. The Tyreus and Luyben settings and detuning factors
obtained for the different configuration for the low purity design of C3 spHtter are shown
in Table 5.6. The TL settings obtained in Table 5.6 compared well with the results
obtained by Hurowitz (1998) using a non-Hnear dynamic model for low purity C3 spHtter.
However the detuning factors did not match well with those generated by Hurowitz as
was the case with the base case C3 spHtter. The tuning factors obtained here were
aggressive as compared to those obtained by Hurowitz. The mismatch between the Hnear
and non-Hnear model for high setpoint changes resulted in the difference in the detuning
factors. The mismatch between the Hnear and non-Hnear model of the low purity design
of C3 spHtter can be seenft-omFigure 5.11. Figure 5.11 shows the comparison between
the Hnear and non-Hnear model for [D,V] configuration. For these results the tuning
settings generated by Hurowitz (1998) were used for both Hnear and non-Hnear model.

5.4.3.1 Setpoint Control Results


The LAE control performance indices for the nine configurations for setpoint
tracking are shown in Table 5.7. The setpoint changes are the same as the ones used for
tuning the detuning factor. It can be seen from Table 5.7 that [L/D,V] configuration
provided the best overhead loop lAE performance with an lAE of 52.72. The overhead
impurity control performance of [L,B], and [L,V/B] configurations was also good with
65

L\E values of 73.93, and 75.11, respectively. The [L/D,V/B] configuration provided
reasonable control performance with an lAE of 87.97. The [L,V] and [D,B] (with tight
level control) configurations performed worst in overhead loop lAE performance with
LAE values of 176.15 and 164.78 respectively. For the bottoms loop lAE performance,
tiie [L,V/B] again stood best witii an L\E of 181.73. The [L/D,V/B] and [D,B] (with tight
level control) configuration provided the next best control performance with LAE values
of 295.53 and 274.25. The [L/D,V] configuration which performed best for the overhead
loop provided poor bottoms impurity control with an lAE of 828.74. The [L,V]
configuration, which performed worst for overhead loop, also performed worst for
bottoms loop with an lAE of 1257.66. Hence it exhibited the worst total lAE performance
with an LAE of 1433.81. On the other hand due to its excellent control of overhead and
bottoms the [L,V/B] configuration performed best witii a total LAE of 255.66. The
[L/D,V/B] configuration provided the next best total lAE performance with an lAE of
383.50.

Table 5.6 Low Purity C3 spHtter dual PI composition controller tuning parameters
Overhead Loop
Bottoms Loop
Configuration

Gain

Gain

-65.32

Integral
Time
5333.33

3.83

Integral
Time
18666.67

Detuning
Factor
0.8

L,B
L,V

-46.49

8000

-8.36

10666.67

1.2

L,V/B

-65.88

5333.33

-1574.25

9333.33

D,B

40.3

13333.33

7.79

16000

1.7

D,V

19.9

16000

-13.38

6666.67

0.5

D,V/B

34.34

13333.33

-2453.41

6666.67

1.4

L/D,B

-1852.88

5333.33

9.01

13333.33

1.7

L/D,V

-1132.79

8000

-13.69

6666.67

1.5

L/D,V/B

-1400.89

6666.67

-2462.62

6666.67

1.6

D,B Tight

36.51

10666.67

10.46

10666.67

2.8

66

^ 4.5

1000

2000

3000

4000

5000

6000

Time (minutes)
Non-linear Model

1000

2000

'Linear Model

3000

SP

4000

5000

6000

Time (minutes)
Non-linear Model """Linear Model

SP

Figure 5.11 Comparison between the closed-loop responses of the Hnear and non-linear
model for setpoint tracking of the [D,V] configuration of the Low Purity C3 Splitter.

67

Table 5.7 Low Purity C3 spHtter LAE indices for overhead impurity setpoint control
Configuration Bottoms Loop LAE
Overhead Loop LAE Total lAE
L,B

382.42

75.11

457.53

L,V

1257.66

176.15

1433.81

L,V/B

181.73

73.93

255.66

D,B

665.64

142.90

808.54

D,V

312.88

103.46

416.34

D,V/B

391.68

153.24

544.92

L/D,B

480.38

79.58

559.96

L/D,V

828.74

52.72

881.46

L/D,V/B

295.53

87.97

383.50

D,B Tight

274.25

164.78

439.03

5.4.3.2 Closed-Loop Bode Plots


The closed-loop Bode plots of overhead and bottoms impurity for sinusoidal feed
composition disturbance were generated in a manner similar to those generated for the
base case C3 SpHtter. The closed-loop Bode plots of [L,V] and [L/D,V/B] configurations
are shown in Figures 5.12 and 5.13, respectively.
The overhead impurity ampHtude ratio plots for sinusoidal feed composition
disturbance rejection are shown in Figure 5.14. From the ampHtude ratio plots it can be
seen that [L/D,B] and [L/D,V] configurations provided the best overall feed composition
disturbance rejection for overhead impurity. The overhead impurity control performance
of [L/D,V/B] configuration was comparable to these configurations except between 0.01
(628 min disturbance period) and 0.02 rad/min (314 min disturbance period). Between
these frequencies, the [L/D,V/B] configuration showed twice the ampHtude ratios. The
[L,B] and [L,V/B] configurations provided analogous feed composition disturbance
rejection performance for overhead impurity with [L,B] configuration being consistently
sHghtly lower than [L/D,B] configuration. The overhead impurity control performance of
[L,V] configuration was pretty reasonable with two peaks at 0.005 rad/min (1257 min
disturbance period) and 0.06 rad/min (105 min disturbance period). The [D,B]
68

configuration performed poorly for low frequencies but provided reasonable control
performance at higher frequencies. It showed a very low minima at around 0.011 rad/min
(571 min disturbance period). In contrast the [D,V] configuration performed well at low
frequencies but exhibited poor overhead impurity control performance at higher
frequencies especially around 0.02 rad/min (314 min disturbance period). The [D,B]
(with tight level control) and [D,V/B] configurations performed worst among all
configurations.
The bottoms impurity ampHtude ratio plots for sinusoidal feed composition
disturbance rejection are shown in Figure 5.15. From the ampHtude ratio plots, it can be
seen that [L,V/B] configuration provided the best bottoms impurity control performance
for feed composition disturbance rejection. The [D,V], [D,V/B] and [L/D,V/B]
configurations provided remarkably analogous control bottoms impurity control
performance which was comparable to that of [L,V/B]. These configurations exhibited
peaks at around 0.012 rad/min (524 min disturbance period). The bottoms impurity
control performance of [L/D,B] and [L,B] configuration was relatively poor with both
configurations exhibiting peaks at around 0.012 rad/min (524 min disturbance period).
The bottoms impurity control performance of [L,V] configuration was poor at low
frequencies but reasonable at high frequencies. Among all configurations, the [D,B] and
[D,B] (with tight level control) configurations exhibited the worst bottoms impurity
control performance.
Again the ampHtude ratio plots of overhead and bottoms impurity for sinusoidal
feed composition disturbance rejection provided lesser ampHtude ratios than those
obtained by Hurowitz (1998) for most of the configurations. But the nature of the closedloop Bode plots was analogous to those obtained by Hurowitz. For tiie overhead
impurity, he found that [L/D,V/B] configuration provided the best overall feed
composition disturbance rejection performance. For the bottoms impurity, he also found
[L,V/B] configuration to be the most optimal.

69

300
200
0.1

100

(0
0)
Q)
L-

(D

Cd
0)
D

0.01

2,

-100

jC

-200

CO

Q.

<

D)
<D

Z
(D

0.001

-300

CL

-400
0.0001

-500

0.001

0.01

0.1

Disturbance Frequency (rad/minute)


'Overhead AR

Bottoms AR

'Overhead Phase

Bottoms Phase

Figure 5.12 Low Purity C3 SpHtter Closed-Loop Bode plot for feed composition
disturbance rejection for [L,V] configuration.
1

100

0
0.1

-100

I-

CD

O)
CD

cr

a>

D
3
*->

-200 S

0.01

-300

5.

<

(0
0)
0)

0.001

Ic
CO
0)
(0
CD

-400 sz
!

-:;

0.

-500
-600
0.01

0.001

0.1

Disturbance Frequency (rad/minute)


'Overhead AR

Bottoms AR

'Overhead Phase

Bottoms Phase

Figure 5.13 Low Purity C3 SpHtter Closed-Loop Bode plot for feed composition
disturbance rejection for [L/D, V/B] configuration.

70

0.1
CD

Q:

....

0) 0.08

. . .

(.

ij

D
3

Q.

0.06

'

'

'

<

I 0.04

' ' ' . ' ' ' ' . ' , ' '

"D
CD

- [^^

0.02

0)

>
O
0

-^"^^^^1=*^
^H^^^t

^ii^?'*****
0.01

0.001

0.1

Disturbance Frequency (rad/min)


" - - [L.B]

[L.Vl

[L,V/B]

[D,B]

[D.V]

(a) For [L,B],[L,V],[L,V/B],[D,B], ],[D,V] and [D,B] (with tight level control)
configurations.
0.08

M^fWMMMIVIMiMMflMIMi^WIMMifMH

l^fMttlHHVMIWIHfaMiWMIMM|IMIl

0.01

0.001

0.1

Disturbance Frequency (rad/min)


- - - [D.V/B]

[L/D.B]

[L/D,V]

[UD,V/B]

[D,B] (Tight)

(b) For [D,V/B],[L,D/B],[L/D,V],[L/D,V/B] and [D,B] (with tight level control)


configurations.
Figure 5.14 Low Purity C3 Splitter Overhead Impurity Amplitude ratio plots for feed
composition disturbance rejection.

71

Disturbance Frequency (rad/min)


- - - [UB]

"[L.V]

[L,V/B]

[D.B]

[D.V]

(a) For [L,B],[L,V],[L,V/B],[D,B], ],[D,V] and [D,B] (with tight level control)
configurations.

Disturbance Frequency (rad/min)


- - - [D,V/B]

[UD,B]

[UD,V] [ U D . V / B ]

ID.B] (Tight)

(b) For [D,V/B],[L,D/B],[IVD,V],[L/D,V/B] and [D,B] (with tight level control)


configurations.
Figure 5.15 Low Purity C3 SpHtter Bottoms Impurity Amplitude ratio plots for feed
composition disturbance rejection.
72

5.4.4 Hiverted Purity C^ SpHtter


The inverted purity design of C3 splitter was analyzed for dual PI composition
control for all nine configurations provided in Table 5.1. The level and composition
controller tuning parameters were obtained in a manner similar to the one used for the
base case design of C3 spHtter. For the inverted purity C3 spHtter the bottoms product is of
higher priority than the overhead product. Hence a setpoint change in overhead impurity
was used to tune the detuning factor. The overhead impurity setpoint was changed from
0.3% to 0.15% at t=100 minutes and then to 0.45% at t=1500 minutes. The simulation
was stopped at t= 3000 minutes. The tuning factor was tuned for minimum LAE in
bottoms impurity for the above set of setpoint changes. The Tyreus and Luyben settings
and detuning factors obtained for the different configuration of the inverted purity design
of C3 spHtter are shown in Table 5.8. The TL settings and detuning factors obtained in
Table 5.8 compared well with the results obtained by Hurowitz (1998) using a non-Hnear
dynamic model for inverted purity C3 splitter. This is because the Hnear model matched
up well with the non-linear model. The comparison between the Hnear and non-Hnear
model for [D,V] configuration can be seen from Figure 5.16. For these results, the tuning
settings generated by Hurowitz were used for both Hnear and non-Hnear model. It can be
seen that the mismatch is much lesser as compared to that for the base case, low purity
and high purity designs of C3 splitter.

5.4.4.1 Setpoint Control Results


The LAE control performance indices for the nine configurations for setpoint
tracking are shown in Table 5.9. The setpoint changes are the same as the ones used for
tuning the detuning factor. The units of the LAE indices are [impurity mole fractionjx
[sec]. It can be seen from Table 5.9 that [L/D,V] configuration provided the best
overhead loop LAE performance with an LAE of 9.41, respectively. The [L/D,B] and
[L/D, V/B] configurations provided the next best overhead impurity control performance
with lAE values of 12.07 and 12.06, respectively. The overhead impurity control
performance of [L,B] configuration was also good with an lAE of 13.97. However the
[D,B] (with tight level control) configuration provided only reasonable control
73

performance with an LAE of 23.14. The [D,V/B] configuration perform.ance was


comparable to that of [D,B] (with tight level control) with an lAE of 26.43.The [L,V] and
[L,V/B] configurations performed worst in overhead loop lAE performance with LAE
values of 53.64 and 65.06. For the bottoms loop lAE performance, as expected [L,V/B]
stood best with an LAE of 19.65. The [L/D,V/B] and [D,V] configurations also provided
good bottoms loop control performance with LAE values of 20.12 and 23.27 respectively.
Surprisingly the [L/D,V] configuration provided reasonable control performance with an
LAE of 22.57. It performed better than [L,B] configuration which provided an LAE of
29.07. The performance of [D,B] (with tight level control) configuration was comparable
to that of [L,B].The [L,V] configuration, which performed poorly for overhead loop, also
performed poorly for bottoms loop with an lAE of 49.61.

Table 5.8 Inverted Purity C3 splitter dual PI composition controller tuning parameters
Bottoms Loop
Overhead Loop

21.11

Integral
Time
18666.67

Detuning
Factor
0.5

8000

-47.69

10666.67

0.9

-44.69

6666.67

-7229.65

10666.67

1.3

D,B

38.47

13333.33

42.16

16000

1.2

D,V

16.58

18666.67

-56.33

8000

0.7

D,V/B

33.01

13333.33

-13525.4

6666.67

2.1

L/D,B

-1666.11

5333.33

36.53

16000

L/D,V

-1103.5

8000

-75.32

6666.67

L/D,V/B

-1720.29

6666.67

-13846.5

6666.67

2.2

D , B Tight

52.91

8000

57.62

10666.67

1.4

Configuration

Gain

L,B

-62.31

Integral
Time
5333.33

L,V

-46.6

L,V/B

Gain

74

500

1000

1500

2000

2500

3000

Time (minutes)
Non-linear Model

'Linear Model

SP

0.5

500

1000

1500

2000

2500

3000

Time (minutes)
Non-linear Model

'Linear Model

SP

Figure 5.16 Comparison between the closed-loop responses of the linear and non-linear
model for setpoint tracking of the [D,V] configuration of tiie Low Purity C3 SpHtter.

75

Table 5.9 Inverted Purity C3 splitter LAE indices for overhead impurity setpoint control
Configuration
Bottoms Loop lAE
Overhead Loop lAE
L,B

29.07

13^97

L,V

49.61

53.64

L,V/B

19.65

65.06

D,B

35.42

29.12

D,V

20.12

17.77

D,V/B

24.38

26.43

L/D,B

33.13

12.07

L/D,V

22.57

9.41

L/D,V/B

23.27

12.06

D,B Tight

30.44

23.14

5.4.4.2 Closed-Loop Bode Plots


The closed-loop Bode plots of overhead and bottoms impurity for sinusoidal feed
composition disturbance were generated in a manner similar to those generated for the
base case C3 spHtter. The closed-loop Bode plots of [L,V] and [L/D,V/B] configurations
are shown in Figures 5.17 and 5.18, respectively.
The overhead impurity ainpHtude ratio plots for sinusoidal feed composition
disturbance rejection are shown in Figure 5.19. Among aU configurations, the [L/D,B]
and [L/D,V] configurations provided the best overall feed composition disturbance
rejection for overhead impurity. The [L/D,V/B] and [L,B] configurations provided good
overhead impurity control with both configurations exhibiting peaks at around 0.015
rad/min (419 min disturbance period). The [L,V] configuration provided reasonable
control performance and exhibited ampHtude ratios close to that of D-/D,V]
configuration. It exhibited a blunt peak at around 0.01 Irad/min (571 min disturbance
period) and a much sharper peak at 0.07 rad/min (628 min disturbance period). Around
0.07 rad/min, tiie control performance of [L,V] was the worst among all configurations.
The overhead impurity control performance of [D,V/B] and [D,V] configuration was the
worst among all configurations. The [D,B] configuration exhibited better control
76

performance than these configurations and shoed a peak at aroun 0.012 rad/min (524 min
disturbance period)
The bottoms impurity amplitude ratio plots for sinusoidal feed composition
disturbance rejection are shown in Figure 5.20. Looking at the ampHtude ratio plots, it
can be concluded that bottoms impurity control performance of [L,V/B] and [D,B] (with
tight level control) configurations were the best overall among all configurations.
However, the [D,B] configuration performed poorly for frequencies around 0.02 rad/min
(314 min disturbance period). The [L,V] configuration performed poorly for frequencies
below 0.01 rad/min (628 min disturbance period) but did fairly well for higher
frequencies. It exhibited two peaks one each at 0.01 rad/min and 0.07 rad/min. The
performance of [L/D,V/B] configuration was reasonable and comparable to that of
[L/D,V] configuration. The [D,B] and [L/D,B] configuration provided the worst control
performance among all configurations.
The ampHtude ratio plots of overhead and bottoms impurity for sinusoidal feed
composition disturbance rejection were comparable to those obtained by Hurowitz (1998)
for the inverted purity design of C3 splitter.

5.4.5 Base Case Depropanizer


The base case design of depropanizer was analyzed for dual PI composition
control for all nine configurations shown in Table 5.10. The level controller tuning
parameters for all configurations were obtained using Equations (4.29) and (4.30) and
Table 4.4. The Tyreus and Luyben (TL) tuning settings were obtained independentiy for
overhead and bottoms composition loops using an ATV test. These settings were then
detuned using a common detuning factor as given by Equation (5.2). The detuning factor
is tuned for minimum integral of the absolute value of error (lAE) in controller response
for setpoint changes. For depropanizer, both overhead product and bottoms are assigned
equal priorities, hence the detuning factor is tuned for minimum LAE in both overhead
and bottoms impurity for a sequence of overhead and impurity setpoint changes. The
setpoint changes were carried out as foUows:

77

100
0
0.1

Vi
Oi

-100

CD

0)

a:
CD

n3

k-

0.01

-200

"5.
E

2,
it
CO

-300

<

0.001

o
(0
CD
JC

CL

-400

Si

0.0001

-500

0.001

0.01

0.1

Disturbance Frequency (rad/minute)


'Overhead AR

Bottoms AR

'Overhead Phase

Bottoms Phase

Figure 5.17 Inverted Purity C3 SpHtter Closed-Loop Bode plot for feed composition
disturbance rejection for [L,V] configuration.
300
200

0.1

100

CD

CU
O
TJ
_3

0.01

1t

-100

CO

11

"Q.

(0
0)
CD
u.
O)

"

!c
CD

</)

<

-200

0.001

-300
1

-400

0.0001
0.01

0.001

0.1

Disturbance Frequency (rad/minute)


'Overhead AR

'Overhead Phase

Bottoms AR

Bottoms Phase

Figure 5.18 Inverted Purity C3 SpHtter Closed-Loop Bode plot for feed composition
disturbance rejection for [L/D,V/B] configuration.

78

CD
.C

CL

0.12

Disturbance Frequency (rad/min)

- -[UB]

[UV]

[L,V/B]

[D.B]

'[D.V]

(a) For [L,B],[L,V],[L,V/B],[D,B], and [D,V] configurations.


0.12

Disturbance Frequency (rad/min)


- -

-[D,V/B]

[lyD.B]

'[LVD.V]

[LyD,V/B]

[D.B] (Tight)

(b) For [D,V/B],[L,D/B],[L/D,V],[L/D,V/B] and [D,B] (with tight level control)


configurations.
Figure 5.19 Inverted Purity C3 Splitter Overhead Impurity Amplitude ratio plots for feed
composition disturbance rejection.

79

0.001

0.01

0.1

Disturbance Frequency (rad/min)


[UB]

[LV]

[L,V/B]

[D,B]

[D,V]

(a) For [L,B],[L,V],[L,V/B],[D,B], and [D,V] configurations.

0.001

0.01

0.1

Disturbance Frequency (racd/min)


-

- - [D,V/B]

[UD.B]

[UOy] [LVD, V/B]

[D.B] (Tight)

(b) For [D,V/B],[L,D/B],[L/D,V],[L/D,V/B] and [D,B] (with tight level control)


configurations.
Figure 5.20 Inverted Purity C3 SpHtter Bottoms Impurity AmpHtude ratio plots for feed
composition disturbance rejection.

80

1. The overhead impurity setpoint was increased to 5% below the nominal impurity at
t=100 minutes.
2. The overhead impurity setpoint was decreased to 5% below the nominal impurity at
t= 1800 minutes.
3. The bottoms impurity setpoint was increased to 5% below the nominal impurity at
t=3600 minutes.
4. The bottoms impurity setpoint was decreased to 5% below the nominal impurity at
t=5400 minutes.
5. The simulation was stopped at t=7200 minutes.
The above procedure was repeated for all nine configurations. The Tyreus and
Luyben settings and detuning factors obtained for the different configurations of the base
case design of depropanizer are shown in Table 5.1.
Previously dual PI composition control of the base case design of depropanizer has
been carried out by Duvall (1999). He used a non-Hnear dynamic tray-to-tray model for
his analysis. He tuned the composition control loops for minimum LAE for setpoint
changes in a similar manner to the one adopted here. However, he used 25% setpoint
changes in the overhead and bottoms impurity. Remarkably, the TL settings and detuning
factors obtained in Table 5.10 were comparable to those obtained by DuvaU (1999) which
indicated that the depropanizer model behaved linearly for the setpoint changes. This can
be seen from Figure 5.21. Figure 5.21 shows the comparison between the responses of
the Hnear model and the non-linear model for setpoint changes. The results are for
[L/D, V/B] configuration of the base case depropanizer and are obtained using the tuning
settings generated by DuvaU (1999).

5.4.5.1 Setpoint Control Results


The LAE control performance indices for the nine configurations for overhead and
bottoms impurity setpoint tracking are shown in Table 5.11. The setpoint changes are the
same as the ones used for tuning the detuning factor. The units of the lAE indices are
[impurity mole fractionjx [sec]. From Table 5.11, it can be seen tiiat [L/D,V/B] provided
the best overhead loop performance with an LAE of 1.63. The overhead loop performance
81

of [L/D,B] and [L/D,V] configurations was comparable to that of [L/D,V/B]. [L,V]


configuration performed worst among all configurations with an LAE of 16.11 and its
response resulted in high oscillations. The overhead loop performance of [L,B] and
[L,V/B] configuration was reasonable with an LAE of 4.33. For the bottoms loop LAE
performance, remarkably the [D,V/B] configuration provided the lowest LAE of 2.49.
[L/D,V/B] configuration also performed well with an LAE of 3.25 for the bottoms loop.
The [L,V] configuration again provided the worst LAE for bottoms impurity control due
to oscillatory behavior.

Table 5.10 Base case Depropanizer dual PI composition controller tuning parameters
Overhead Loop
Bottoms Loop

2.81

Integral
Time
6666.67

Detuning
Factor
0.9

8000

-8158.69

4666.67

0.7

-12.53

5333.33

-32.72

4266.67

0.8

D,B

6.75

14000

3.62

6666.67

0.7

D,V

2.43

16000

-11244.2

4000

0.6

D,V/B

3.89

16000

-34.73

4000

0.6

L/D,B

-432.94

5333.33

3.09

6666.67

1.0

L/D,V

-125.4

8000

-11157.2

4000

0.6

L/D,V/B

-318.17

5866.67

-35.57

4000

0.8

Configuration

Gain

Gain

-16.9

Integral
Time
4666.67

L,B
L,V

-3.78

L,V/B

The bottoms impurity control performance of [L/D,V] and [D,V] configurations


was relatively poor but better than [L,V] configuration. The overall control performance
for botii overhead and bottoms loop was the best for tiie [L/D,V/B] configuration and the
worst for [L,V] configuration. The [LyD,B] configuration provided the next best control
performance with an lAE of 8.82. The total lAE was somewhat misleading because some
of the configurations performed well for one loop but performed poorly for the other
loop. For example the [L/D,V] configuration performed better for overhead loop than

82

[D,V/B] configuration. However it performed much poorer than [D,V/B] configuration


for the bottoms loop. But both configurations resulted in total LAE values close to 8.8.

0.70

r
0.30

200

400

600

800

1000

1200

1400

1600

1400

1600

Time (minutes)
Non-linear Model

Linear Model

SP

0.70
o
E 0.60 3
a 0.50
E

k
'

A^

"o
3

0.40 -

Vi

E
o 0.30

O
CD

0.20
0

200

400

600

800

1000

1200

Time (minutes)
Non-linear Model

'Linear Model

SP

Figure 5.21 Comparison between the closed-loop responses of the linear and non-linear
model for setpoint tracking of [L/D,V/B] configuration of base case depropanizer.
83

Table 5.11 Base Case Depropanizer LAE indices for overhead impurity setpoint control
Configuration Bottoms Loop lAE
Overhead Loop LAE Total lAE
L,B

5.03

3.59

8.61

L,V

15.18

16.11

31.29

L,V/B

5.03

4.33

9.35

D,B

4.65

6.47

11.12

D,V

7.65

6.65

14.30

D,V/B

2.49

6.24

8.73

L/D,B

4.79

2.18

6.97

L/D,V

6.97

1.85

8.82

L/D,V/B

3.25

1.63

4.88

5.4.5.2 Closed-Loop Bode Plots


The closed-loop Bode plots of overhead and bottoms impurity for sinusoidal feed
composition disturbance rejection were generated for al the nine configurations shown in
Table 5.1. For generating the closed-loop Bode plots, sinusoidal feed composition
disturbances in the low key (LK) and heavy key (HK) are input to the Hnear dynamic
model of base case depropanizer under closed-loop conditions. From the resulting steady
state responses, ampHtude ratios and phase angle shifts were calculated to obtain closedloop Bode plots. The frequencies were chosen in the range of 0.5 to 0.001 rad/min, as the
ampHtude ratios outside this range were negHgible. The closed-loop Bode plots of [L,V]
and [L/D,V] configurations are shown in Figures 5.22 and 5.23, respectively. The
ampHtude ratio parts of the Bode plots of all nine configurations are shown in Figures
5.24 and 5.25, respectively.
The overhead impurity ampHtude ratio plots for sinusoidal feed composition
disturbance rejection are shown in Figure 5.24. From careful analysis of the ampHtude
ratio plots it can be seen that [L,V/B] configuration provided the best overall feed
composition disturbance rejection for overhead impurity. The overhead impurity control
performance of [L,V] and [L/D,V/B] configurations was comparable to that of [L,V/B]
configuration. The overhead impurity control performance of [D,V/B] and [D,V]
84

configuration was the worst among all configurations for frequencies below 0.08 rad/min.
But these configurations performed reasonably at higher frequencies. The performance of
[D,B] configuration was very poor between 0.04 (157 min disturbance period) and 0.1
rad/min (63 min disturbance period). AU configurations exhibited peaks at around 0.02
rad/min (314 min disturbance period) indicating that the overhead impurity of base case
depropanizer was most sensitive to feed composition disturbance around these
frequencies. All configurations except [D,V], [D,V/B] and [L/D,V/B] configurations
showed smaller peaks at high frequencies.
The bottoms impurity amplitude ratio plots for sinusoidal feed composition
disturbance rejection are shown in Figure 5.25. From the ampHtude ratio plots it can be
seen that [L/D,V/B] and [D,V/B] configurations provided the best feed composition
disturbance rejection for bottoms impurity. The disturbance rejection performance of
[L,V/B] configuration was comparable to that of [L/D,V/B] configuration. Both
configurations exhibited a peak at around 0.011 rad/min (571 min disturbance period).
The [L,V] configuration performed well for low frequencies but performed poorly for
frequencies in the range of 0.1 (63 min disturbance period) and 0.15 rad/min (42 min
disturbance period). The [D,B] configuration similarly performed well at low frequencies
but provided worst bottoms impurity control performance for frequencies between 0.03
(209 min disturbance period) and 0.1 rad/min (63 min disturbance period). In contrast to
the [L,V] and [D,B] configurations, [L,B] and [D,V/B] performed poorly for low
frequencies but delivered reasonable performance for high frequencies. UnUke the
overhead impurity ampHtude ratio plots, most of the configurations exhibited much closer
heighted peaks for bottoms impurity amplitude ratio. On the higher frequency side most
of the configurations exhibited peaks at around 0.02 rad/min (314 min disturbance
period) which indicated the most sensitive zone of bottoms impurity for feed composition
disturbance. On the lower frequency side, the [L,B], [D,V] and [D,V/B] configurations
exhibited peaks at around 0.01, 0.02 and 0.03 rad/min, respectively.
The ampHtude ratio plots of overhead and bottoms impurity for sinusoidal feed
composition disturbance rejection were analogous and comparable to those obtained by
DuvaU (1999). For the overhead impurity he too found [L,V/B] and [L,V] to provide the
85

best feed composition disturbance rejection performance. For the bottoms impurity, he
too found [D,V/B] configuration to provide the best control performance.
0.01

300
200
100

0.001

(0

0
CD

-100 ^

tr
D

CD
CD

0.0001

-200 ^

-300 !

E
<

(/>

-400 ^
a.

0.00001

-500
-600

0.000001
0.001

0.01

-700

0.1

Disturbance Frequency (rad/minute)


'Overhead AR

Bottoms AR

'Overhead Phase

Bottoms Phase

Figure 5.22 Base Case Depropanizer Closed-Loop Bode plot for feed composition
disturbance rejection for [L,V] configuration.
0.01

f=
1

0.001

CD

tr

1
1

""""^

CD
D

\\

0.0001

CD

-300 %

CD

-400 f
-500
-600
-700

^ ^ -

0.01

0.1

Disturbance Frequency (rad/minute)


'Overhead AR

o)

-100 I
-200 I

^^^

"Q.

0.00001 L
0.001

?
CD

400
300
200
100

'Overhead Phase

Bottoms AR

Bottoms Phase

Figure 5.23 Base Case Depropanizer Closed-Loop Bode plot for feed composition
disturbance rejection for [LTD,V/B] configuration.
86

0.24
'

<

CD

cr

0.2

<D
D

Q. 0.16
E
<
^ 0.12
3

a.
-E 0.08
T3
CD
CD

JZ
k-

CD

0.04

>

^^^^^^^^^^^^^^T^

V^^^^^^^^^^

1 t

^ ^ ^ ^ ^

^ \ ^

'

1 (

0
0.001

0.01

0.1

Disturbance Frequency (rad/min)

[UV]

-[UB]

[L,V/B]

[D.V]

[D,B]

(a) For [L,B],[L,V],[L,V/B],[D,B], and [D,V] configurations.


0.24

'

CD

0.2

,s

CD

T5
_3

t 0.16
E

<
^

- - *

'

- -

0.12

- %

k-

a3
-^ 0.08

T V

T3
CD
<D

0.04 ^

. -

>

0.01

0.001

0.1

Disturbance Frequency (rad/min)


- -

-ID.V/B]

[LyD.B]

[LVD.V]

[LyD,V/B]

(b) For [D,V/B],[L,D/B],[L/D,V] and [I7D,V/B] configurations.


Figure 5.24 Base Case Depropanizer Overhead Impurity AmpHtude ratio plots for feed
composition disturbance rejection.

87

0.42
CD

cr
<D
D
3
Q.

-.._.J

-,-~rA

0.35

.i
1 ll

]- -

0.28

<

If

0.21

a
E 0.14

.
^ -

i / ^ I* X

' ''A"

\\

.'

, .'

*^

E
o
o 0.07

.^

. . .

. - '

CD

iS^

0.001

0.01

\% 1
^ 1

0.1

Disturbance Frequency (rad/min)


- [UB]

ILV]

[L,V/B] [ D , B ]

[D,V]

(a) For [L,B],[L,V],[L,V/B],[D,B], and [D,V] configurations.

0.01

0.1

Disturbance Frequency (rad/min)


- - - [D,V/B]

[LyD.B]

[LiD.V]

[L/D, V/B]

(b) For [D,V/B],[L,D/B],[L/D,V] and [L/D,V/B] configurations.


Figure 5.25 Base case Depropanizer Bottoms Impurity AmpHtude ratio plots for feed
composition disturbance rejection.

88

5.4.6 High Purity Depropanizer


The high purity design of depropanizer was analyzed for dual PI composition
control for all nine configurations shown in Table 5.1. The level and composition
controller tuning parameters were obtained in a manner similar to the one used for the
base case design of depropanizer. The same sequence of setpoint changes was used to
tune the common detuning factor. The Tyreus and Luyben settings and detuning factors
obtained for the different configurations of the high purity design of depropanizer are
shown in Table 5.12. The tuning parameters obtained in Table 5.12 compared well with
the results obtained by Duvall (1999) using a non-linear dynamic model for high purity
depropanizer. This is because the Hnear model used here matched well with the nonHnear model used by Duvall (1999). The comparison between closed-loop responses of
the linear and non-Hnear model can be seen from Figure 5.26.

5.4.6.1 Setpoint Control Results


The LAE control performance indices for the nine configurations for overhead and
bottoms impurity setpoint tracking are shown in Table 5.13. From Table 5.13, it can be
seen tiiat [L,V/B], [L/D,V/B] and [L/D,V] configurations provided tiie best overhead
loop performance with LAE values of 0.64, 0.66 and 0.69, respectively. The [L/D,B] and
[L,B] configurations provided the next best overhead control performance with LAE
values of 0.71 and 0.86, respectively. Remarkably the [L,V] configuration performed
better than both [D,B] and [D,V] configurations in both overhead and bottoms impurity
control. The [D,B] and [D,V] configurations provided the worst bottoms and overhead
impurity control performance. The bottoms impurity control performance was best for
[L,V/B] and [LVD,V/B] configurations. The bottoms loop L\E values of aU remaining
control configurations were relatively close indicating analogous bottoms loop behavior.
Thus, the overall control performance of [L,V/B] and [L/D,V/B] configurations was the
best among all configurations. The [L,B] configuration too provided good total L\E
control performance with an lAE of 1.97. The [D,B] and [D,V] configurations resulted in
the worst total lAE performance due to their poor control performance for both loops.

89

0.14

I 0.13 i
^

0.12 -

3
CI

E 0.11
"o
3
D
O

T3

0.10
0.09 -

CD
(D

s:
<D

0.08 -

>

o 0.07
0

200

400

600

800

1000

1200

1400

1600

Time (minutes)
Non-linear Model

'Linear Model

SP

0.14

I 0.12
g. 0.10

o
-D 0.08
o
CL
CO

i 0.06 o
CD

0.04

-1

200

400

600

"T'

800

1000

1200

1400

1600

Time (minutes)
Non-linear Model

'Linear Model

SP

Figure 5.26 Comparison between tiie closed-loop responses of the Hnear and non-Hnear
model for setpoint tracking of [L/D,V/B] configuration of high purity depropanizer.

90

Table 5.12 High Purity Depropanizer dual PI composition controller tuning parameters
Overhead Loop
Bottoms Loop
Configuration

Gain

Integral
Time

Gain

Integral
Time

Detuning
Factor

L,B

-27.09

6266.67

10.54

9466.67

0.9

L,V

-9.71

10000

-21709.8

5333.33

0.7

L,V/B

-23.22

6933.33

-72.18

4533.33

0.8

D,B

28.28

19333.33

13.24

10000

1.4

D,V

9.67

20000

-24348.4

4666.67

0.6

D,V/B

18.38

17866.67

-72.86

4000

1.1

L/D,B

-710.9

6133.33

11.97

9333.33

1.2

L/D,V

-296.89

9066.67

-24153.3

5066.67

0.6

L/D,V/B

-593.96

6666.67

-72.26

4000

0.9

Table 5.13 High Purity Depropanizer LAE indices for overhead impurity setpoint control
Configuration Bottoms Loop LAE
Overhead Loop lAE Total LAE
L,B

1.26

0.71

1.97

L,V

1.44

1.60

3.04

L,V/B

0.86

0.64

1.49

D,B

1.57

2.16

3.73

D,V

1.77

1.99

3.75

D,V/B

1.33

1.88

3.21

L/D,B

1.25

0.86

2.10

L/D,V

1.41

0.69

2.09

L/D,V/B

0.98

0.66

1.65

5.4.6.2 Closed-Loop Bode Plots


The closed-loop Bode plots of overhead and bottoms impurity for sinusoidal feed
composition disturbance were generated in a manner similar to those generated for the

91

base case depropanizer. The closed-loop Bode plots for [L,V] and [L/D,V/B]
configurations are shown in Figures 5.27 and 5.28, respectively.
The overhead impurity amplitude ratio plots for sinusoidal feed composition
disturbance rejection are shown in Figure 5.29. From the ampHtude ratio plots it can be
seen that [L,V/B] and [L/D,V/B] configurations provided the best overall feed
composition disturbance rejection for overhead impurity and exhibited peaks at around
0.03 rad/min (209 min disturbance period). The overhead impurity control performance
of [L/D,V] configuration was only slightly worse than these configurations. The [L,B]
and [L/D,B] configurations provided reasonable and analogous feed composition
disturbance rejection performance. The [L,V] configuration provided excellent overhead
impurity control performance at low frequencies but performed poorly at higher
frequencies. It exhibited a peak at around 0.1 rad/min (63 min disturbance period). The
overhead control performance of [D,B], [D,V] and [D,V/B] configurations was worst
among all configurations. All configurations exhibited peaks at different frequencies
indicating different sensitive zones for different configurations.
The bottoms impurity amplitude ratio plots for sinusoidal feed composition
disturbance rejection are shown in Figure 5.30. From tiie ampHtude ratio plots, it can be
seen that [L/D,V/B] and [L,V/B] configurations which performed best for overhead
impurity control also provided the best feed composition disturbance rejection for
bottoms impurity. Both configurations exhibited peaks at around 0.011 (571 min
disturbance period) and 0.1 rad/min (63 min disturbance period). The [L,V] and [L/D,V]
configurations provided excellent bottoms impurity control performance for lower
frequencies but performed poorly for higher frequencies. They both exhibited sharp peaks
at around 0.1 rad/min (63 min disturbance period). The [D,B] configuration performed
worst amongst all configurations.

92

0.01

0.001

400

^=Z\ \ I
ija^plg^

a.

200

I ^ ' ' i",.^


~

[ " ^ J P L ^

0)
D)

0.0001

-200

CD

n
_3

I"

CO
<D

-400

0.00001

<

-600

CD

!c
CO
CD
CO
CD

0.000001
-800
0.0000001
0.001

-1000
0.01

0.1

Disturbance Frequency (rad/minute)


'Overhead AR

Bottoms AR

'Overhead Phase

Bottoms Phase

Figure 5.27 High Purity Depropanizer Closed-Loop Bode plot for feed composition
disturbance rejection for [L,V] configuration.

400

0.01

300
200

0.001

100
CD

cr

0.0001

-100

(D

"5.
E

(0
CD
CD

D)
<D
T3

-200 g
0.00001

-300 ^

<

CD

-400 ^

0.000001

-500
-600

0.0000001
0.001

-700
0.01

0.1

Disturbance Frequency (rad/minute)


'Overhead AR

Bottoms AR

'Overhead Phase

Bottoms Phase

Figure 5.28 High Purity Depropanizer Closed-Loop Bode plot for feed composition
disturbance rejection for [L/D,V/B] configuration.

93

0.08

'>>W

I I

lyil

|iiiMiiin.M

1^

,1.11,1

'

<

CD

<

Q:

CD
D

_3

'

'

I ^X"
^^^
'^^*

3
D.

CD

^^^H^^^

'

^^^^^

' _lJr^

^^_^^^/K^^
^

E
CD

III

I lliimi

0.06

"5.
E
<
0.04

T3

"<

\ '

<

^'

t\

t \

'

^k

'

l l

1
I

I
1

0.02

JC
k.

CD

>

0
0.001

0.01

0.1

Disturbance Frequency (rad/min)


- -

[L.V]

-[UB]

[L,V/B]

[D.B]

[D.V]

(a) For [L,B],[L,V],[L,V/B],[D,B], and [D,V] configurations.


0.08
CD

Q:
(D
D

_3

"

0.06

"Q.

E
<
0.04

CD
.C
u.
CD

>

a
E
T3
CD

0.02

- 0

V
^k
t

>

^ ^ ^ ^

>

0
0.01

0.001

0.1

Disturbance Frequency (rad/min)


- - - [D,V/B]

[L/D,B]

[L/D,V]

"IL/D,V/B]

(b) For [D,V/B],[L,D/B],[L/D,V] and [L/D,V/B] configurations.


Figure 5.29 High Purity Depropanizer Overhead Impurity AmpHtude ratio plots for feed
composition disturbance rejection.

94

0.001

0.01

0.1

Disturbance Frequency (rad/min)


-

[L.V]

-[UB]

'[L.V/B]

[D.B]

[D.V]

(a) For [L,B],[L,V],[L,V/B],[D,B], and [D,V] configurations.

0.01

0.1

Disturbance Frequency (rad/min)


[D,V/B]

[L/D.B]

'[uoy]

[LVD.V/B]

(b) For [D,V/B],[L,D/B],[L/D,V] and [L/D,V/B] configurations.


Figure 5.30 High Purity Depropanizer Bottoms Impurity Amplitude ratio plots for feed
composition disturbance rejection.
The ampHtude ratio plots of overhead and bottoms impurity for sinusoidal feed
composition disturbance rejection were consistent with those obtained by DuvaU (1999).
95

For sinusoidal feed composition disturbance rejection, he too found [L,V/B] and
[L/D,V/B] configurations to be the most optimal for overhead impurity and [L/D,V/B]
configuration to be the most optimal for bottoms impurity.
5.4.7 Low Purity Depropanizer
The low purity design of depropanizer was analyzed for dual PI composition
control for all nine configurations shown in Table 5.14. The level and composition
controller tuning parameters were obtained in a manner similar to the one used for the
base case design of depropanizer. The same sequence of setpoint changes was used to
tune the common detuning factor. The Tyreus and Luyben settings and detuning factors
obtained for the different configurations of the low purity design of depropanizer are
shown in Table 5.15. The tuning parameters obtained in Table 5.15 compared well with
the results obtained by Duvall (1999) using a non-Hnear dynamic model for low purity
depropanizer. This is because the Hnear model used here matched weU with the nonlinear model used by Duvall (1999). The comparison between closed-loop responses of
the Hnear and non-Hnear model can be seen from Figure 5.31.
Table 5.14 Low Purity Depropanizer dual PI composition controller tuning parameters
Overhead Loop
Bottoms Loop
Configuration

Gain

Gain

-5.32

Integral
Time
4133.33

1.28

Integral
Time
6266.67

Detuning
Factor
0.9

L,B
L,V

-2.76

5733.33

-4487.78

4000

0.7

L,V/B

-4.67

4666.67

-11.2

4000

0.8

D,B

3.48

12000

1.65

6400

1.0

D,V

1.68

12666.67

-5311.38

3733.33

0.6

D,V/B

5.09

7600

-12.82

3733.33

1.2

L/D,B

-137.62

4666.67

1.43

6133.33

l.O

L/D,V

-77.32

5866.67

-5030.83

3866.67

0.7

L/D,V/B

-115.58

5066.67

-13.83

3600

0.9

96

3.00

t-

2.50 -

a
E

t3 2.00
3
T5
O

"S 1.50CD
SZ
CD

>
O 1.00

200

400

600

800

1000

1200

1400

1600

1400

1600

Time (minutes)
Non-linear Model

Linear Model

SP

3.00

I 2.50
i. 2.00
E
o
D
o

4.

VU

1.50

0.
CO

E 1.00
o
CD

0.50

200

400

600

800

1000

1200

Time (minutes)
Non-linear Model

'Linear Model

SP

Figure 5.31 Comparison between the closed-loop responses of the linear and non-Hnear
model for setpoint tracking of [L/D,V/B] configuration of low purity depropanizer.

97

5.4.7.1 Setpoint Control Results


The lAE control performance indices for the nine configurations for overhead and
bottoms impurity setpoint tracking are shown in Table 5.15. The setpoint changes are the
same as the ones used for tuning the detuning factor. The units of the L\E indices are
[impurity mole fractionjx [sec]. From Table 5.15, it can be seen that [I7D,V/B] and
[L/D,V] configurations provided the best overhead loop performance with an lAE values
of 5.54 and 5.13, respectively. The [L/D,B] configuration provided the next best
overhead control performance with an lAE of 6.53. [L,B] and [L,V/B] configurations
provided reasonable control performance with lAE values of 10.16 and 10.20,
respectively. The [L,V] configuration performed pooriy for overhead impurity control but
its performance was better tiian that of [D,V] and [D,B] configurations. The [D,B]
configuration provided the worst control performance with an lAE of 22.78.
The bottoms impurity control performance was the best for [L/D,V/B]
configuration. The [L,B], [L,V/B] and [I7D,B] configurations provided good bottoms
impurity control witii lAE values of 13.69, 14.04 and 14.09 respectively. The [D,B]
configuration which performed worst for overhead impurity provided resaonable control
performance with an lAE of 15.21. The [L,V] configuration resulted in tiie worst control
performance due to its oscillatory behavior. Thus, the overaU control performance of
[L/D,B] and [L/D,V/B] configurations was the best among aU configurations. The [L,B]
and [L,V/B] configurations provided equivalent control performance for both loops and
hence resulted in close total lAE values of 23.85 and 24.24. The [D,B] configuration
offered the worst total LAE performance due to its poor control performance for both
loops.

5.4.7.2 Closed-Loop Bode Plots


The closed-loop Bode plots of overhead and bottoms impurity for sinusoidal feed
composition disturbance were generated in a manner similar to those generated for the
base case depropanizer. The closed-loop Bode plots of [L,V] and [17D,V/B]
configurations are shown in Figures 5.32 and 5.33, respectively. The ampHtude ratio parts

98

of the Bode plots of all nine configurations are shown in Figures 5.34 and 5.35,
respectively.

Table 5.15 Low Purity Depropanizer LAE indices for overhead impurity setpoint control
Configuration Bottoms Loop lAE
Overhead Loop lAE Total lAE
L,B

13.69

10.16

23.85

L,V

30.91

16.22

47.13

L,V/B

14.04

10.20

24.24

D,B

15.21

22.78

37.99

D,V

16.18

18.92

35.10

D,V/B

19.37

19.88

39.25

L/D,B

14.09

6.53

20.62

L/D,V

22.13

5.13

27.26

L/D,V/B

10.70

5.54

16.24

The overhead impurity amplitude ratio plots for sinusoidal feed composition
disturbance rejection are shown in Figure 5.34. Among all configurations it can be seen
that [L,V/B] configuration provided the best overall feed composition disturbance
rejection performance for overhead impurity. The [L,V] configuration provided
analogous behavior but exhibited higher ampHtude ratios between 0.1 (63 min
disturbance period) and 0.2 (31 min disturbance period) rad/min. The [L/D,V/B] and
[L/D,V] configurations provided good overhead impurity control performance. The
performance of [L/D,B] configuration was reasonable but worse than [L/D,V/B] and
[L/D,V] configurations. The overhead impurity control performance of [D,V/B]
configuration was the worst among all configurations. It exhibited a huge peak at 0.04
rad/min (16 min disturbance period). The [D,B] and [D,V] configurations performed
pooriy but provided better control than [D,V/B] configuration.

99

0.01

400
300
200
100

0.001

CD

^W

a:

^.M-^^l

^ v

CD

CD
<D
k.

O)
CD
Tl

-200 (/)
-300 CO

_3

"5.

-100

CO

0.0001

-400
-500

CD
SZ

a.

-600
0.00001
0.001

-700
0.01

0.1

Disturbance Frequency (rad/minute)


'Overhead AR

Bottoms AR

'Overhead Phase

Bottoms Phase

Figure 5.32 Low Purity Depropanizer Closed-Loop Bode plot for feed composition
disturbance rejection for [L,V] configuration.
0.01

400
-

300

200
100
^

0.001

CD

a:

CD
(D
u.
CJ)
CD

-100 ;o^

CD
D
_3

-200 (/)
-300 0CO

Q.

<

CO

0.0001

-400
-500
-600
-700

0.00001
0. 001

0.01

0.1

Disturbance Frequency (rad/minute)


'Overhead AR

Bottoms AR

'Overhead Phase

Bottoms Phase

Figure 5.33 Low Purity Depropanizer Closed-Loop Bode plot for feed composition
disturbance rejection for [L/D,V/B] configuration.

100

CD

CL

0.8
CD

a:

0.7

>

<

CD

T3
3

0.6

= - - -

- -

- -

- ^ f ^ \ /
#^

'

^\'

' - L A I .

' \ '

M\

"5.
E 0.5
<

0.4

* /^

"W '

a 0.3

E
D
CD
CD
.C
k.

CD

>

^^.-- "''' ""> * \ \

0.2
0.1

"Xr*"/\

^_^

jjg

^jj^

0
0.001

0.01

0.1

Disturbance Frequency (rad/min)

[L.V]

-[UB]

'[L,V/B]

[D.B]

[D.V]

(a) For [L,B],[L,V],[L,V/B],[D,B], and [D,V] configurations.


0.8

Q:

0.7

. . .

CD

J -

1 ' . 1 :

"'

<

1 . -

mpl

0.6

0.4

ty

J _

: ; : : : : : : :

0.5

r,.,

<
k-

I -.-ir
.
. 1

T r . - . - - . . . .

^ : : : v

'

J.

- -

. - . - -

E 0.3

CD
0
r.

0.2

L-

0.1
0

p"*

, .

\l

>

.^
'

"

<

'

0.01

0.001

0.1

Disturbance Frequency (rad/min)


- - - ID.V/B]

[LVD.B]

[L/D,V]

[LVD, V/B]

(b) For [D,V/B],[L,D/B],[L/D,V] and [L/D,V/B] configurations.


Figure 5.34 Low Purity Depropanizer Overhead Impurity AmpHtude ratio plots for feed
composition disturbance rejection.

101

0.8
CD

a:
0
D
3

-_i

. .

0.7

'

'

'

. .

. 1

. .

1.

11

1 . 1

. . . .

1 1
1 1

1 . . .

0.6

- -, -

. .
E 0.5
Q.

..

y . . . . \

^^
O
CD

u,-

I 0.3

r-

r-

^' 0.4

E 0.2
o

- - - -

<

(0

I I

*J

A //

\\

- -

0.1

m^

0
0.001

0.01

0.1

Disturbance Frequency (rad/min)


-

[UV]

-[UB]

[L,V/B]

[D.B]

[D.V]

(a) For [L,B],[L,V],[L,V/B],[D,B], and [D,V] configurations.

0.8
CD

Q:

mpl

1.

1 T

-.>l

0.7

0
3
.ti

0.6

1^

-I

r -

0.5

<
>, 0.4

"

^ - 1

L_

a 0.3
b
CO
0.2
E
o
0.1

ott

^ 1

<

m^^

'.
^

'- "

III

CD

- , . , .--

* '

0.01

0.001

0.1

Disturbance Frequency (rad/min)


-

-[D,V/B]

[lyD.B]

'[LyD.V]

[L/D,V/B]

(b) For [D,V/B],[L,D/B],[L/D,V] and [L/D,V/B] configurations.


Figure 5.35 Low Purity Depropanizer Bottoms Impurity AmpHtude ratio plots for feed
composition disturbance rejection.
The bottoms impurity amplitude ratio plots for sinusoidal feed composition
disturbance rejection are shown in Figure 5.35. From the plots, it can be seen that
102

[L,V/B] configuration which performed best for overhead impurity control also provided
the best overall bottoms impurity control for feed composition disturbance rejection. It
exhibited a sharp peak at around 0.15 rad/min (42 min disturbance period). The [L/D,V]
and [L/D,V/B] configurations also provided excellent bottoms impurity control
performance comparable to that of [L,V/B] configuration. The [L,V] configuration
provided excellent control performance at low frequencies but resulted in high ampHtude
ratios at higher frequencies especially around 0.15 rad/min (42 min disturbance period).
The bottoms impurity control performance of [D,B] configuration was relatively worst
among all configurations. Most of the configurations exhibited dual peaks except for the
[L,V] configuration.
Again the ampHtude ratio plots of overhead and bottoms impurity for sinusoidal
feed composition disturbance rejection were consistent with those obtained by Duvall
(1999). For sinusoidal feed composition disturbance rejection, he too found [L,V/B] and
[L, V] configurations to be the most optimal for both overhead impurity and bottoms
impurity.
5.4.8 Asymmetric Purity Depropanizer
The asymmetric purity design of Depropanizer was analyzed for dual PI
composition control for all nine configurations shown in Table 5.1. The level and
composition controller tuning parameters were obtained in a manner similar to the one
used for the base case design of depropanizer. The same sequence of setpoint changes
was used to tune the common detuning factor. The Tyreus and Luyben settings and
detuning factors obtained for the different configurations of the high purity design of
depropanizer are shown in Table 5.16. The tuning parameters obtained in Table 5.16
compared well with the results obtained by DuvaU (1999) using a non-Hnear dynamic
model for the asymmetric purity depropanizer. This is because the linear model used here
matched well with the non-Hnear model used by DuvaU. The comparison between closedloop responses of the linear and non-linear model for setpoint changes can be seen in
Figure 5.36.

103

Table 5.16 Asymmetric Purity Depropanizer dual PI composition controller tuning


parameters
Overhead Loop
Bottoms Loop
Configuration

Gain

L,B

Gain

-58.52

Integral
Time
5866.67

L,V

-18.84

L,V/B

1.11

Integral
Time
6266.67

Detuning
Factor
1

9333.33

-3994.32

4133.33

0.6

-45.38

6666.67

-10.75

4000

0.7

D,B

25.44

16000

1.42

6533.33

0.8

D,V

10.65

16800

-4733.92

3866.67

0.6

D,V/B

15.57

16666.67

-13.65

3600

0.6

I7D,B

-1423.78

6666.67

1.27

6000

L/D,V

-572.31

9333.33

-4415.73

3733.33

0.6

L/D,V/B

-1083.13

7333.33

-12.84

3733.33

0.7

5.4.8.1 Setpoint Control Results


The LAE control performance indices for the nine configurations for overhead and
bottoms impurity setpoint tracking are shown in Table 5.17. The setpoint changes are the
same as the ones used for tuning the detuning factor.
The units of the LAE indices are [impurity mole fractionjx [sec]. From Table 5.17
it can be seen that [L/D,B], [LTD,V/B] and [L/D,V] configurations provided the best
overhead loop performance with LAE values of 0.60, 0.52 and 0.47, respectively. The
[L,V/B] configuration provided the next best overhead control performance with an lAE
value of 1.19. The [L,V] configuration again provided the worst lAE control performance
due to its oscillatory response. The overhead loop performance of [L,B] configuration
was reasonable with an lAE of 1.66. The [D,B] and [D,V] configurations performed
poorly for overhead impurity control but performed better that [L,V] configuration. The
bottoms impurity control performance was best for [D,V/B] and [L/D,V/B]
configurations. The [L,V] configuration as expected again resulted in the highest lAE
value of 25.34 due to its oscillatory response. Only the [D,B] and [L/D,V] configuration
came close to the performance of [L,V] configuration with lAE values of 22.61 and
104

19.52, respectively. The [D,B] configuration which performed worst for overhead
impurity control performed remarkably well for bottoms impurity control with an LAE of
11.34. The bottoms impurity control performance of [L/D,B] was comparable to that of
[D,B].
_ 0.15
^ 0.14
o

I; 0.13 i
10.12 ^

t3 0.11 -1
3

2 0.10
g 0.09 0

0 0.08
>

0.07 0

200

400

|i

600

800

1000

11

1200

1400

1600

1400

1600

Time (minutes)
Non-linear Model

Linear Model

SP

2.8
a> 2.6
o
E
^ 2.4

I 2.2
o 2.0 3

T5
O

D.
CO

E
o
5o
CD

1.81.61.4

200

400

600

800

1000

1200

Time (minutes)
Non-linear Model

'Linear Model

SP

Figure 5.36 Comparison between tiie closed-loop responses of the linear and non-Hnear
model for setpoint tracking of [L/D,V/B] configuration of high purity depropanizer.
105

The total LAE control performance was the best for [L/D,V/B] configuration with
an LAE of 8.60. As expected, it was the worst for [L,V] configuration with an LAE of
28.29. The [D,V/B] configuration, which performed well for bottoms loop and fared poor
for overhead loop, provided good total lAE control performance with an lAE of 13.54. In
contrast the [L/D,V] configuration provided poor total lAE control performance with an
LAE of 19.99. The overaU control performance of [D,V/B] and [I7D,B] configurations
was good with an LAE of 10.11 and 11.34, respectively.

Table 5.17 Asymmetric Purity Depropanizer LAE indices for overhead impurity setpoint
control
Configuration Bottoms Loop LAE
Overhead Loop LAE Total lAE
L,B

12.66

1.66

14.32

L,V

25.34

2.94

28.29

L,V/B

12.17

1.19

13.36

D,B

11.34

2.19

13.54

D,V

22.61

2.22

24.84

D,V/B

8.13

1.98

10.11

L/D,B

10.74

0.60

11.34

L/D,V

19.52

0.47

19.99

L/D,V/B

8.07

0.52

8.60

5.4.8.2 Closed-Loop Bode Plots


The closed-loop Bode plots of overhead and bottoms impurity for sinusoidal feed
composition disturbance were generated in a manner similar to those generated for the
base case depropanizer. The closed-loop Bode plots for [L,V] and [L/D,V/B]
configurations are shown in Figures 5.37 and 5.38, respectively. The amplitude ratio parts
of the Bode plots of all nine configurations are shown in Figures 5.39 and 5.40,
respectively.
The overhead impurity ampHtude ratio plots for sinusoidal feed composition
disturbance rejection are shown in Figure 5.39. Analyzing the amplitude ratio plots, it
can be seen that [L,V] and [L,V/B] configurations provided the best feed composition
106

disturbance rejection performance for overhead impurity. Both configurations provided


analogous control performance with peaks at 0.04 rad/min (157 min disturbance period).
The [L,V] configuration however showed another peak at around 0.12 rad/min (52 min
disturbance period). The [L/D,V] configuration provided good overhead impurity control
performance with a peak at 0.04 rad/min (157 min disturbance period). The overhead
impurity control performance of [L/D,V/B] configuration was similar to the [L/D,V]
configuration but it consistently exhibited sHghtly smaller ampHtude ratios than [L/D,V]
configuration. The [D,V] configuration provided the worst control performance among
all configurations and exhibited the largest peak at 0.025 rad/min (251 min disturbance
period). The overhead impurity control performance of [D,V/B] configuration was
relatively poor but it was better than [D,V] configuration. The [L/D,V] and [L/D,B]
configuration provided better control performance that [D,B] configuration.
The bottoms impurity ampHtude ratio plots for sinusoidal feed composition
disturbance rejection are shown in Figure 5.40. From a careful analysis of the plots, it can
be observed that [L/D,V] and [L/D,V/B] configurations provided the best overall feed
composition disturbance rejection performance for bottoms impurity. The two
configurations exhibited analogous control performance except around 0.1 rad/min (63
min disturbance period) where the [L/D,V] configuration exhibited much lower
ampHtude ratios than those of [L/D,V/B] configuration. The [L,V] configuration which
provided the best control performance for overhead impurity also performed well for
bottoms impurity except for frequencies between 0.1 (63 min disturbance period) and 0.2
rad/min (31 min disturbance period). The [D,V/B] configuration also provided good
control performance comparable to that of [L/D,V] configuration. The P,V/B] and [L,B]
configurations provided the worst bottoms impurity control performance among all
configurations. Their control performance was remarkably analogous and both exhibited
large peaks at 0.1 rad/min. The bottoms impurity control performance of [D,B]
configuration too was relatively poor and comparable to tiiat of [L,B] configuration.
The ampHtude ratio plots of overhead and bottoms impurity for sinusoidal feed
composition disturbance rejection were consistent with those obtained by Duvall (1999).
For sinusoidal feed composition disturbance rejection, he too found [L,V/B] and [L,V]
107

configurations to be the most optimal for overhead impurity. He als o found [L/D,V/B]
configuration to provide good control performance for the overhead impurity.

0.01

400
300
200

0.001
CD

tr

0.0001

CO
0
0

O)

-100 I

"5.
E

100

-200 I

0.00001

-300 S

<

CD

-400 ^

0.000001

-500
-600

0.0000001

-700
0.01

0.001

0.1

Disturbance Frequency (rad/minute)


'Overhead AR

Bottoms AR

'Overhead Phase

Bottoms Phase

Figure 5.37 Asymmetric Purity Depropanizer Closed-Loop Bode plot for feed
composition disturbance rejection for [L,V] configuration.

g1

0.01

400
300
200

0.001

100 0?

p=-^ 1 ^
CD

Q:

- J

11

_3
"Q.

0
^
-100 g

0.0001

-200 I

0.00001

-300 $

<

CD

-400
0.000001

-500
-600

0.0000001
0.001

-700
0.01

0.1

Disturbance Frequency (rad/minute)


'Overhead AR

Bottoms AR

'Overhead Phase

Bottoms Phase

Figure 5.38 Asymmetric Purity Depropanizer Closed-Loop Bode plot for feed
composition disturbance rejection for [L/D,V/B] configuration.
108

0.05
CD

0
D
3

0.04

I.

a:

"Q.

^^

1 1

^X

E 0.03
<>.
^.^

^ ^

3
Q.

*i-

0.02

E
D

x:

0.01

^ ^ - .

>

>

>

'

^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ ^ " ^ ^ ^ ^ ^ ^ ^ ^

0
0.001

^^ ^

>^

i.....

CD
0

0.01

0.1

Disturbance Frequency (rad/min)


'[L,V/B]

[UV]

- - - [L.B]

"[D.B]

[D.V]

(a) For [L,B],[L,V],[L,V/B],[D,B], and [D,V] configurations.


0.05
CD

CL

1
L

0 0.04
T3

"Q.

J<

0.03

...

,;

>.

*^

'L-

1 0.02

> ;;

T3

CD

0.01

. *.
0

P^

>

0
0.001

0.01

0.1

Disturbance Frequency (rad/min)


- -

-ID.V/B]

[LVD.B]

'[lyD.V]

[UD.V/B]

(b) For [D,V/B],[L,D/B],[I7D,V] and [L/D,V/B] configurations.


Figure 5.39 Asymmetric Purity Depropanizer Overhead Impurity AmpHtude ratio plots
for feed composition disturbance rejection.

109

0.7

T' 1

J, . A

.9

- --

1 0.6
0
T3

B 0.5

1 Xi

'1 *

] A, *
\ *
ip - |L

'

A
1 1

: / ; : :r\\ \

Q.

1 0.4
% 0.3
a
E
0.2
E
^ 0.1

, X,
y

- .* -

ma- - - y -

. . .

^ *

J *, ^j r '

JSrjfC

A J - -1 -

/ . .. yf\

.--

....

-^^^^

1 "

^^^^^^^^

CD

0.001

0.01

0.1

Disturbance Frequency (rad/min)

[UV]

- - -[UB]

[L.V/B]

[D.B]

[D.V]

(a) For [L,B],[L,V],[L,V/B],[D,B], and [D,V] configurations.

, 1 1 ,

if

l l

- 1

o o o o o o

Bottoms Impurity Amplitude Ratio

.1

-\

i-

#1

-1

0
1

0.01

0.001

0.1

Disturbance Frequency (rad/min)


-

- - [D,V/B]

[lyD.B]

[L7D,V]

[UD,V/B]

(b) For [D,V/B],[L,D/B],[L/D,V] and [L/D,V/B] configurations.


Figure 5.40 Asymmetric Purity Depropanizer Bottoms Impurity Amplitude ratio plots for
feed composition disturbance rejection.

110

CHAPTER 6
SIGNAL PROCESSING TECHNIQUES

Signal processing, in general, is essentially concerned with the mathematical


representation of a signal and the algorithmic operation carried out on it to extract
information present in the signal. The signal under scrutiny here is an industrial feed
composition signal and the information needed is the frequency components contained in
the signal. The aim is to represent the signal as a sum of different sinusoidal signals using
signal processing techniques. In an industrial environment, measured signals are
corrupted by sensor noise, which can either be due to electrical interference, mechanical
vibration etc. Signal processing techniques can help in cleaning a signal and removing
any unwarranted noise present in it. Hurowitz (1998) has provided a detailed description
of signal processing techniques appHed to discrete signals to obtain their frequency
information. The techniques employed here for signal processing are adopted from
Hurowitz (1998).
Thus, a feed composition disturbance entering a distillation column can be
represented as a series of sine waves entering a linear model of the column and yielding a
series of output sine waves with the same frequencies. The schematic for this is shown in
Figure 6.1, where a signal with frequency components, O},, is shown. Since a distillation
column behaves linearly for smaH sinusoidal changes around steady state conditions, a
linear model can be used to represent its behavior accurately. The resulting signal would
have the same frequency components (0 as the other signal since the process is linear.
For small ampHtude signals, the C3 spHtter and the depropanizer resulted in output
signals, which had the same frequency as the original signal, thereby supporting the
assumption of treating the distillation columns as Hnear systems. The only variation in the
output sine waves would be in their ampHtudes, A' and phase angles, 0,'. The changes in
the ampHtudes and phase angles can be obtained from a closed-loop Bode plot of the
linear model for the feed composition disturbance. Hence, the product variability
resulting from a feed composition disturbance can be obtained by multiplying each output
111

sine wave with its frequency associated amplitude ratio from the Bode plot and summing
up the series.
Therefore the objective of this approach is to characterize the ampHtude and
frequency components of a feed composition disturbance by a series of sine waves at
various frequencies.
XAsin(wr-H0)

Y^Ksiniwj+ (!);,)

Figure 6.1 Signal representation of feed disturbance entering a linear model


6.1 Discrete Fourier Transforms
Our aim is to obtain ampHtude and phase spectra of a discrete signal using
discrete Fourier Transforms. Let's consider a signal h(k) with A^ sample points and
sampHng interval of A T. The discrete Fourier Transform for this signal is defined as
k=N-l

H = Y,h{k)e^-"^"' .

(6.1)

k=Q

Expanding the exponential term we get


k=N-\

..=

k=0

k=N-l
(2Kkn
27tkn
h(k)
sin
h(k)COS
+7-1
A^
N
k=0
V

(6.2)

which can be represented as


//=/? + ;/

(6.3)

where Rn represents the real part and / represents the imaginary part of //, the n"
frequency component of the discrete Fourier transform. These are given in terms of the
signal h(k) by
k=N-\

K=h(0)+

X h(k) cos
A=l

k=N-l

'..= 1
k=0

h(k)sm

^ 2Kkn ^
\

(6.4)

y-i

^ 2Kkn
V N

(6.5)
112

The Fourier transform can be represented in polar form as


J^n

H=\H.

(6.6)

where |//| is the ampHtude of // and Z // is tiie phase angle of //. These can be
represented in terms of tiie real and imaginary parts of // as
(6.7)

tan"' /

when

180-tan"'

when

ZH = i
180 +tan"'
360-ta n-' '

when
when

R>0
R<0
R>0

(6.8)

R<0

/ >o

The angular frequency of the n^ component is given by


( 2K ^
[NAT J

(6.9)

Here, // represents the discrete frequency spectrum of the signal with n varying
from 0 to A^-1. Similarly |//| and Z //. respectively represent the ampHtude and phase
spectra. These spectra represent the relative magnitudes and phases of the infinitesimal
signal making up a continuum of complex sinusoids representing the original signal.
The total power contained in the time discrete signal can be obtained using, Parseval's
Theorem as
1

k=N-l

Total P o w e r s - J^h'(k)=

N tA

r;=.V-l

J^H;

2;: S

(6.10)

For spectral analysis of a signal it is necessary to study the variation of power of a signal
with respect to frequency. This can be achieved by analyzing the one-sided power
spectral density {PSD) of the signal (Press et al., 1992)
(6.11)

(PSD)_,^,,^H;+H_:-

113

iPS^n)one-sided^K'^l'-

(6-12)

Thus by analyzing the power spectral density of a signal, one can learn the
various frequency components that make up the signal and their relative strengths. By
looking at the PSD, one can determine the effective bandwidth of a signal where most of
the signal power lies.
Thus, the frequency components of any signal can be obtained by analyzing its
power, ampHtude and phase spectra, which can be generated using the discrete Fourier
transformation as shown above. The standard discrete Fourier transform algorithms
consume lot of computational time. If a signal has A^ data samples, then it would require
(A^-1) calculations using these algorithms. As the number A'^ grows larger, (A^-l)^ amounts
to heavy computational burdon. Hence, a decimation-in-time, radix 2 fast Fourier
transform is used which reduces the required number of calculations to just ViNlogiN.
The only point to be considered here is that radix 2 fast Fourier transform requires A^ to
be an integral power of 2, A^=2''. This can be easily achieved by adding some zeros at the
end(WaDcer, 1991).
The decimation in time radix 2 fast Fourier transform essentially takes advantage
of the symmetry and periodicity involved in the discrete Fourier transforms computation.
It essentially divides the N sample points into two halves of N/2 points. Each of these
subdivisions is further divided until a single sample point is hit. This form of Fourier
transform is caUed decimation in time as the calculations are carried out in the time
domain with the number of sample points decreasing i.e., reducing by half in this case
(Hurowitz, 1998)

6.2 The SampHng Theorem and Signal AHasing


The sampling theorem states that a continuous bandlimited function h(t) can be
reconstructed from its discretely sampled data (WaUcer, 1991),
hit) = f

(6.13)

2/J
provided that f is bandlimited by -fc < f < fc , where/(- is called the Nyquist critical
frequency, defined as.
114

f . - ^

(6.14)

where AT is the sampHng interval of h{t). A time Hmited signal cannot be bandlimited
(Kamen, 1990). Hence its power lying outside of the frequency range -fc <f<fc is
folded over or "aHased" into that range as a result of discrete sampHng. This is called
signal aHasing and results in a distorted version of the original signal.
Signal aliasing can be observed by generating a discrete Fourier Transform from a
discrete time Hmited signal and checking the signal power at the Nyquist critical
frequency, fc. If the power near the Nyquist critical frequency is approaching zero then
signal aliasing is minimal. If on the contrary the power is approaching a finite value then
chances are that frequency components outside of the range have been folded back on to
the critical range. The only way to get rid of signal aliasing is by passing the signal
through a lowpass filter which removes all high frequency components present in the
signal. The cutoff frequency for the lowpass filter is chosen depending upon the
sensitivity of the system. Most of the measured industrial signals contain high frequency
noise, which could be due to electrical or vibrational interference in the sensors. Hence, a
lowpass filter would faciHtate in cleaning the signal by removing any noise present in it.
For the industrial feed composition signal, a bandpass filter is used to get rid of signal
aliasing and noise present in it (Hurowitz, 1998).

6.3 Digital Filtering in Time Domain


Digital filtering can be done in both time and frequency domain. Since the current
discussion deals with real time appHcations, time domain was chosen as a better
alternative. Moreover, the filtered data can be monitored and compared with original data
onHne. If filtering is done in tiie frequency domain, to study the filtered data tiie inverse
Fourier transformation would have to be carried out. Detailed descriptions about Digital
Filtering can be obtained from Proakis et al. (1996) and Press et al. (1992).
For analyzing the industrial feed composition signal, a five-coefficient infinite
impulse response (HR) bandpass filter proposed by Press et al. (1992) is used. The
response function for the filter is given by
115

Mf) =

b
+

_.
z '

{i + a){\^b)
(l-Hfl)(l-h^)
(l-l-a)(l-^)-h(l-fl)(l-HZ7) _ iX-a)i\-b)
_.
{\-\-a){\ + b)
z + (l-Ha)(l-f-^)
z

(6.15)

where a is the lower cutoff frequency and b is the upper cutoff frequency and they are
chosen based on the analysis of the phase spectra of the signal. The basic idea behind
filtering is to eliminate signal aliasing by removing high frequency noise. Recursive
filters are used because they offer much more stable performance, which is superior to
nonrecursive filters with the same number of coefficients (Hurowitz, 1998).

6.4 Treatment of End Effects by Zero Padding


Another phenomenon that happens as a result of convolution of discretely
sampled signals using the Fast Fourier Transform is the wrap around effect. This happens
because the discretely sampled signals are assumed to be periodic. Secondly it assumes
that the duration of the response signal is the same as the period of the data. This results
in false corruption of the first output channel with some wrapped around data from the far
end of the signal, hn-i,hn.2,hn-3 etc. The wrap around effect can be eUminated by adding a
buffer of zeros at the end of the signal to make the corruption zero (Press et al., 1992).
Complete details about the signal processing technique are provided by Hurowitz
et al. (1998). The signal processing procedure described above is shown in a nutshell in
Figure 6.2. The difference between the procedure here and the procedure followed by
Hurowitz et al. (1998) is that the procedure of detrending, i.e., the Hne joining the start
point and the end point of the data, has not been included in Figure 6.2. This is because
detrending the data results in loss of appreciable information contained in the data, which
is not really desired. Presence of trend in the data leads to some edge effects. To achieve
the best results using signal processing, it is recommended to use signals or select part of
the signal such that the trend in it is insignificant.

116

Step 1. Remove the mean the feed composition data samples to enhance the variation of
the signal around the mean
Step 2. Apply a lowpass HR recursive filter to eliminate the possibility of signal aHasing.
Step 3. Since the radix 2 fast Fourier transform requires N to be an integral power of 2,
i.e., N=2'', and to get around the wrap around problem, apply zero-padding to the
data.
Step 4. Apply a fast Fourier transform to obtain the following,
1. Power Spectral Density
2. Amplitude Spectrum
3. Phase Spectrum
Figure 6.2. Signal Processing Procedure
6.5 Industrial Feed Composition Signal
Figure 6.3 represents an industrial feed composition signal for C3 spHtter. It
corresponds to the propylene composition in the feed. This signal was used for analyzing
the C3 spHtter using approach proposed in Chapter 1. The feed composition sample
contains 5760 data measurements with a sampling interval of 1 min. The feed
composition signal value varies from 0.66% to 0.82%. The mean of the signal is 72.9%
whereas the steady state value of the C3 spHtter model is 70%. The sharp vertical Hues in
the signal correspond to analyzer failure or caHbration periods. Before signal processing
techniques were applied, these points were removed and replaced by the last
measurement/analyzer reading.
The signal used for the C3 spHtter was also used for the depropanizer with some
modifications. Firstiy, the steady state value of the signal is moved to the composition of
propane in depropanizer feed. Then the deviation of the signal around the steady state
value is reduced by a factor of three. Since the depropanizer is faster than the C3 spHtter,
the sampHng period for this signal was taken as 0.2 seconds. The resulting signal used
for depropanizer is shown in Figure 6.4.

117

1000

2000

3000

4000

5000

6000

Time (minutes)

Figure 6.3 C3 SpHtter Feed Composition Signal

200

400

600

800

1000

Time (minutes)

Figure 6.4 Depropanizer Feed Composition Signal


6.5 Results of Signal Processing Analysis
The feed composition signals of C3 spHtter and depropanizer are subjected to
signal processing techniques as shown in Figure 6.2 to obtain their power spectral
118

density, amplitude and phase spectrum. These results can be seen in Figures 6.5 and 6.6
corresponding to C3 splitter and depropanizer, respectively. It can be noticed that power
spectral density, ampHtude and phase spectra of the depropanizer are similar to those of
the C3 spHtter since the same C3 splitter feed composition signal with some modifications
was used for depropanizer. For both C3 sHtter and depropanizer, the lower cutoff
frequency for the bandpass filter was set to zero. The upper cutoff frequency was chosen
as O.l for C3 splitter because, for frequencies above 0.1 rad/min there was negHgible
effect on its closed-loop product variabiHties. For similar reasons, the upper cutoff
frequency was chosen as 0.5 rad/min for depropanizer. Since the feed composition
signals had 5760 data samples, 2432 zeros were added at the end of the data to make the
total number of data samples, N=8192=2 , required for radix 2 fast Fourier transform
computation.

119

Vi

c
0)

o
<u

Q.

(0
o
D.

0.01

0.02

0.03

0.04

0.05

0.06

0.07

0.08

0.09

0.1

Angular Frequency (rad/minute)

(a) Power Spectral density Spectrum.


12000
%

10000 -

%. 8000
E
<

6000 -

4000 -

V)

<

2000

0.01

0.02

0.03

004

0.05

0.06

0.07

0.08

0.09

0.1

0.08

0.09

0.1

Angular Frequency (rad/min)

(b) AmpHtude Spectrum.


360

0.01

0.02

0.03

0.04

0.05

0.06

Angular Frequency (rad/min)

(c) Frequency Spectrum.


Figure 6.5 Signal Processing of C3 SpHtter Signal.
120

0.07

1.E+08
V)

c
Q
"TO
k.
^

0)

Q.

w
o
CL

0.05

0.1

0.15

0.25

0.2

0.3

0.35

0.4

0.45

0.5

Angular Frequency (rad/minute)

(a) Power Spectral density.

0.05

0.1

0.15

0.25

0.2

0.3

0.35

0.4

0.45

0.5

0.35

0.4

0.45

0.5

Angular Frequency (rad/min)

(b) AmpHtude Spectrum.


360

I 300 i
1-

% 240
_2
c
<
V)
CD

SZ

CL

0.05

0.1

0.15

02

0.25

0.3

Angular Frequency (rad/min)

(c) Frequency Spectrum.


Figure 6.6 Signal Processing of Depropanizer Signal.
121

CHAPTER 7
PRODUCT VARIABILITY PREDICTION

For many processes in the chemical processing industries, the primary objective is
to reduce product variabihty. The approach shown in Figure 3.1 provides a simpHstic
approach to predict variabilities in distillation products. In chapter 4, linear industrially
benchmarked dynamic models were developed for C3 spHtter and depropanizer. In
Chapter 5, closed-loop Bode plots for feed composition disturbance were generated for
these distillation columns. The industrial feed composition signals of these distillation
columns were analyzed and their frequency content was extracted using signal processing
techniques in Chapter 6. Using the frequency content information and the closed-loop
Bode plots for disturbance, product variabihty can be obtained as proposed in Figure 3.1.

7.1 Prediction Technique


Consider the equations involved in combining the frequency information and the
closed-loop Bode plot. Let's consider a signal h(k) with a sampling interval of A 7 and N
sample points. Let H represent the Fourier transform coefficients obtained by using the
signal processing techniques. From Equation (6.1) it is given by,
H='%h{k)e^"^"'

(7.1)

k=0

The Fourier transform coefficients can be represented in terms of amplitude and phase
using Equation (6.6)
H^=:\H,y^"

{12)

where ampHtude |//| is given by Equation (6.7) and Z // is given by Equation (6.8). The
original signal can be reconstructed using the inverse discrete Fourier transform as,
h{k)=YHe-^''^"

(7.3)

The above signal is regenerated in discrete form and can be extended to obtain a
continuous signal as.

122

(7.4)

where, w is the frequency corresponding to the n''' Fourier transform coefficient and is
given by Equation (6.9),
0)=n

( 2K \
NAT)

(7.5)

The discrete Fourier transform and inverse Fourier transform are symmetric over positive
and negative frequencies (Kamen, 1990). Hence we get
tin = Hw-n

(7.6)

Expanding the Equation (7.4) results in


n=N/.i

h{t) =

(7.7)

77

^=0

n=N/2

From Equations (6.4), (6.5) and (6.8), we get


ZH\

=0

(7.8)

"\n^.N/2

Using the above relation. Equation (7.7) expands to


"l(n=0)

n=V.-l

.(.)4^

\H \{cos{cot -ZHJ-

^, [+ \H\{cos{cot-ZHJ +
H.

(n=N/2)

j sin{coj - ZH))"
jsin{(oj-ZHj)

(7.9)

which can be written as


n=%-l

w = ^A^^ / / U + [^^J^^^K^-^^")]+^l(.=./.)
n=l

(7.10)

The above equation can be modified to


n=y,-i

h{t) = ^'J(n=0) +
N

//..sin (OJ-ZH
w=l

+ H." l ( ; i = A ' / 2 )

(7.11)

J-i

The closed-loop Bode plot gives us information about ampHtude ratios and phase shifts
caused to the distillation products due to feed composition disturbance at various

123

frequencies. This information can be represented in the closed-loop Bode plot using the
Amplitude ratios as AR(w) and the phase shifts as 0{w) corresponding to frequency w.
Combining the closed-loop Bode plot and frequency information from signal
processing, we get
AR{(0)H\
^

var(r) = <
N +

"^

^ +AR{co )H I
''l(n=0)

"''

"\{n=N/2)

(7.12)

2A/?((yj|//|sin o)t-ZH +e{co) + ^


n=l

/JJ

which represents tiie product variability prediction.


It can be seen that Equation (7.12) involves N/2-i-l frequency components of
signal obtained using signal processing techniques. For most of the practical purposes, all
of the frequency components are never required. Only those frequency components
across which most of the signal power Hes would be required to obtain an accurate
prediction of the product variabiHties. These frequency components can be found by
inspecting the Power Spectral Density (PSD) of the signal, which gives the distribution of
power in the signal across its frequency components. Also only those frequency
components would be needed for which the ampHtude ratios of the system are not
negHgible. In fact this is achieved by passing the signal through a bandpass filter which
removes the frequency components of the signal for which the system ampHtude ratios
are negHgible.

7.2 Closed-Loop Product Variability Prediction


The approach shown in Figure 3.1 is demonstrated for product variabihty
prediction for the different designs of C3 splitter and depropanizer. The potential of the
approach as a quantitative tool for distillation configuration selection is also evaluated.
The ampHtude and phase spectra obtained using signal processing techniques in Chapter
5 are combined with the closed-loop Bode plot for feed composition disturbance
developed in Chapter 4 to predict product variability according to the Equation (7.12). In
order to verify our proposed approach for predicting distillation product variability, the
results are compared with the closed-loop results from a nonlinear tray to tray simulator
of C3 SpHtter (Gokhale et al., 1995) and depropanizer (DuvaU, 1999). The same tuning
124

parameters, used for the linear models, are appHed to tiie non-Hnear tray-to-tray
simulators. Essentially the product variabihty is predicted using the results of signal
processing combined with the closed-loop Bode plot obtained from tiie Hnear dynamic
model. These results are then compared with the results obtained from a nonHnear trayto-tray simulator by inputting the feed composition profile (Figures 6.2 and 6.3) to the
nonHnear simulator under closed-loop conditions. This procedure was done for all nine
configurations for the different designs of C3 splitter and depropanizer. For both C3
splitter and depropanizer, tiie results were generated using only 131 of the available
frequency components from the feed composition signals. In essence, only the frequency
components in the range, from 0 to 0.1 rad/min for C3 spHtter, and from 0 to 0.5 rad/min
for depropanizer, were used in the output prediction. The remaining frequency
components contained little or no signal power and hence, were neglected.

7.2.1 Base Case C^ SpHtter


7.2.1.1 Results
The results of the product variabihty prediction technique for [L,B], [D,B] (with
tight level control) and [L/D,B] configurations of the base case design of C3 spHtter are
shown in Figure 7.1. These results were obtained by combining the frequency
information shown in Figure 6.4 with the closed-loop Bode plot information for feed
composition disturbance shown in Figures 5.4 and 5.5 according to the Equation (7.12).
From the Figure 7.1, it can be seen that the product variabiHties were accurately
predicted for the base case design of C3 splitter. It should be noted that there is an
appreciable mismatch between the results from the nonlinear tray to tray simulator and
the results from the prediction technique during the initial minutes. This was observed for
aU nine configurations. From Figure 7.1, it can be seen that mismatch is observed for the
first 200 minutes for [L,B] configuration, 300 minutes for [L,V] configuration and 400
minutes for [L/D,B] configuration. The initial mismatch was observed because the mean
of the propylene feed composition signal is 72.9%, whereas the steady state value of the
propylene feed composition for the nonlinear tray to tray simulator is 70%. The C3
splitter nonHnear simulator requires the initial time period to suppress this deviation. It
125

should also be realized that the ampHtude and phase spectra in the closed-loop Bode plot
correspond to the steady state responses of the process at different frequencies and do not
account for their initial transients. From Figure 7.1, it can be seen that the variation of the
bottoms impurity (40%) is much higher than that of the overhead impurity (10%).
However, the overhead mismatch for the overhead and bottoms impurity was
comparable. On the whole, it can be said that an excellent prediction of the closed-loop
product variabilities was obtained by the proposed technique due to the relatively small
deviation of the feed composition from the steady state value.
The results obtained here for [L,B] configuration of C3 splitter were in fact better
than the results obtained by Hurowitz et al. (1998). The primary reason being that he
obtained closed-loop Bode plot using a non-Hnear model, which was inaccurate due to
non-linearity of the model. In addition, during the signal processing of the feed
composition signal, he removed the trend (the line joining the initial and the final point)
in the signal thereby compromising the information content of the signal.
The overhead propane impurity lAE and bottoms propylene impurity lAE values
obtained from the results using the prediction technique for all nine configurations of the
base case design of C3 spHtter are shown in Table 7.1. It should be noted that the LAE
values are a measure of the average variabihty in the products. From Table 7.1, it can be
seen that that [L/D,V] configuration provided tiie lowest overhead product variabihty for
C3 spHtter with an LAE of 7.59. The overhead loop performance of [L/D,B] and
[L/D,V/B] configurations was comparable to that of [L/D,V] configuration. [D,V] and
[D,V/B] provided reasonable overhead loop LAE performance with an LAE of 20.79 and
23.51, respectively. [L,V] and [L,V/B] configurations performed worst in overhead loop
lAE performance with an lAE of 49.34 and 46.51, respectively. Surprisingly [L,B]
configuration performed poorly with an lAE of 32.32. Also overhead loop LAE
performance of [D,B] configuration was comparable to that of [L,B]. For the bottoms
loop lAE performance, [D,V/B] stood best with an LAE of 317.91 and [L,V] performed
worst with an lAE of 2484.25. [L/D,V] which provided the best overhead loop
performance fared poorly for bottoms loop performance with an LAE of 1914.84 close to
that of [L,V]. The bottoms loop L\E performance of [L/D,V/B] and [D,V/B]
126

configurations was comparable to that of [L,V/B]. [L/D,B] and [L,B] configurations


provided reasonable bottoms loop LAE performance with LAE values around 900.

o
E

0.33
0.32

3
O.

0.31

0)

c
a.
o

(D

0.29
D

(0
0)

.c

0.28 -

0)

>
O

0.27
0

1000

2000

3000

4000

5000

Time (minutes)
Simulator Response

1000

2000

'Predicted Response

4000

3000

5000

Time (minutes)
'Predicted Response

Simulator Response

Figure 7.1 Closed-Loop Rejection of disturbance of the Base case C3 Splitter, (a) For the
[L,B] Configuration.

127

0.31

E
0.305

a
E
C

>p

CD o ^

Q.
O

T3

0.295 -

0)

>
O

1000

2000

3000

4000

5000

Time (minutes)
Simulator Response ""^Predicted Response

1000

2000

3000

4000

5000

Time (minutes)
"Simulator Response

Predicted Response

Figure 7.1 Continued, (b) For the [D,B] (with tight level control) configuration.

128

1000

2000

3000

4000

5000

Time (minutes)
'Simulator Response

1000

'Predicted Response

2000

3000

4000

Time (minutes)
Simulator Response

'Predicted Response

Figure 7.1 Continued, (c) For the [L/D,B] configuration.

129

5000

Table 7.1 Base case C3 SpHtter LAE indices for Product Variability Prediction
Configuration
Bottoms Loop LAE
Overhead Loop lAE
-

985.07

32.32

L,V

2484.25

49.34

L,V/B

356.60

46.52

D,B

884.45

32.39

D,V

638.85

20.79

D,V/B

317.91

23.52

L/D,B

891.55

10.53

L/D,V

1914.84

7.59

L/D,V/B

423.48

14.14

D,B Tight

449.66

10.17

7.2.1.2 Discussion
Based on the results of dual PI composition control of base case C3 spHtter it can
be concluded that [L/D,V], [L/D,V/B] configurations and [D,B] configuration (witii tight
level control) provided the best overall control. Using the approach shown in Figure 3.1,
the [L/D, V] configuration provided the least variabihty for the overhead product, which
is of higher priority than the bottoms. The [UDy/B] configuration and [D,B]
configuration employing tight level control also provided low variability in the overhead
product. However for bottoms product, the variability in [L/D,V/B] configuration and
[D,B] configuration (with tight level control) was much lower than tiiat of [L/D,V]
configuration. For setpoint tracking too, the [L/D,V] configuration provided the best
overhead impurity control performance but at the expense of the bottoms impurity. The
setpoint tracking performance of [L/D,V/B] and [D,B] configurations was good for botii
overhead and bottoms impurity. The performance of [L/D,V/B] configuration was
comparable to that of [D,B] configuration with tight level control. A similar trend was
observed for disturbance rejection of sinusoidal feed composition.
These results did not exactiy match with the results obtained by Hurowitz (1998)
and Finco et al.(l989) based on their analysis of the C3 spHtter using a non-Hnear tray-to130

tray model. Hurowitz found [L,B], [L,V/B] and [D,B] (with tight level control)
configurations to provide the best overall control performance for the base case C3
splitter. Finco et al. also found [D,B] with tight level control to be best for tiie C3 spHtter.
The [D,B] configuration with tight level control is however not recommended due to its
lack of integrity as pointed out by Finco et al. (1989). The results obtained here seem to
agree with the results of Hurowitz (1998) and Finco et al. (1989) only on tiie
performance of [D,B] configuration (with tight level control). The performance of the
remaining configurations did not match at aU. The reason for mismatch between the
results Hes in the tuning approach. Hurowitz used a non-linear model to tune his
composition controllers, whereas here a linear model was used to tune the composition
controllers. The composition controllers were tuned for setpoint changes. There is a high
degree of mismatch between the Hnear and non-linear model for setpoint changes as
shown in Figure 5.1, which resulted in the difference in tuning results. This mismatch in
the tuning results contributed to the inconsistency in the results obtained here and those
of Hurowitz (1998). In fact, for [L/D,V] and [L/D,V/B] configurations, Hurowitz (1998)
obtained much higher detuning factors which explains the poor performance he observed
for these configurations.

7.2.2 High Purity C^ SpHtter


7.2.2.1 Results
The results of the product variabihty prediction technique for [L,B], [L,V] and
[L/D,B] configurations of the high purity design of C3 spHtter are shown in Figure 7.2.
These results were obtained by combining the frequency information shown in Figure 6.4
with the closed-loop Bode plot information for feed composition dismrbance shown in
Figures 5.9 and 5.10, according to the Equation (7.12).

131

0.108

E
3
Q.

0.104 -

E
C NO
CD o^
Q.
O
T3
(C

0.1

0.096 -

>

0.092
0

1000

2000

3000

4000

5000

Time (minutes)
Simulator Response

Predicted Response

m 0.04 -^

1000

2000

3000

4000

5000

Time (minutes)
Simulator Response "Predicted Response

Figure 7.2 Closed-Loop Rejection of disturbance of the High Purity C3 SpHtter. (a) For
the [L,B] configuration.

132

^
o^
0)

0.104

0.102

>N

+.

Imp

0.1

0)

c
CL
O
IQ.

0.098

0.096
SZ
L0)

>

0 0.094
1000

2000

3000

4000

5000

Tinne (minutes)
"Simulator Response

Predicted Response

0.14
O

>, 0.12 3
Q.

E
0)

0.1

Oi

>>
a.
o

Q:

0.08 -

(0

E
o
m 0.06
0

1000

2000

3000

4000

5000

Time (minutes)
'Predicted Response

Simulator Response

Figure 7.2 Continued, (b) For the [D,B] (with tight level control) configuration.

133

Time (minutes)
"Simulator Response

o^
0)

Predicted Response

0.14

>^
> H
^

0.12

u.

E
0)

c
0)
>>
a.
o

Q.

0.1

0.08

(A

o
*.o

0.06
0

1000

2000

3000

4000

5000

Time (minutes)
"Simulator Response ~ ^ P r e d i c t e d Response

Figure 7.2 Continued, (c) For the [L/D,B] configuration.

134

From the Figure 7.2, it can be seen that the product variabiHties were accurately
predicted for the high purity design of C3 spHtter. For the remaining configurations too,
the product variabiHties were accurately predicted. Analogous to C3 spHtter, an initial
mismatch between the results from the nonHnear tray to tray simulator and the results
from the prediction technique was observed during the initial minutes for aU
configurations. The reason for the initial mismatch was identical to tiiat for the base case
C3 spHtter. The variation in the bottoms and overhead product was much smaller than that
for the base case C3 splitter. However, tiie overall mismatch between the results from tiie
nonlinear tray to tray simulator and the results from the prediction technique for the high
purity C3 splitter was comparable to that for the base case C3 spHtter. But on the whole it
can be concluded that an excellent prediction of the closed-loop product variabiHties was
obtained using the proposed technique.
The overhead impurity and bottoms impurity LAE values obtained from the
product variabihty prediction for the different configurations of the high purity design of
C3 spHtter are shown in Table 7.2. From Table 7.2, it can be seen that [L/D,V]
configuration provided tiie best overhead loop LAE performance with an lAE of 3.89. The
[L/D,B], [D,B] (with tight level control) and [L/D,V/B] configurations also provided
excellent overhead impurity control with LAE values of 4.45, 5.13 and 5.62, respectively.
The overhead loop LAE performance of [D,V/B] configuration was comparable to that of
[D,V]. These configurations performed better than [L,B]. The worst control performance
resulted for [L,V] and [L,V/B] configurations. Both configurations exhibited LAE values
around 22. For the bottoms loop performance, the [D,V/B] configuration resulted in the
best controller performance with an LAE of 24.84. Surprisingly the [D,V/B] configuration
provided better performance than [L,V/B] configuration which showed an LAE of 24.84.
The [L/D, V/B] configuration also provided good control performance with an LAE of
29.38. The [D,B] (with tight level control) and [L/D,B] configuration provided
reasonable bottoms loop performance with lAE values of 40.85 and 44.01. The bottoms
loop performance of [L/D,B] configuration was better than that of [L,B] and [L/D,B]
configurations. The [L,V] configuration performed poorly for bottoms loop too. However
the performance of the [L,V] configuration was better than that of [L/D,V] configuration
135

which performed worst among all configurations. The total LAE performance was
minimum for [D,V/B] configuration indicating that it offered the most optimal control for
C3 splitter. The [L/D,V/B] configuration also provided excellent control for both
overhead and bottoms loop and showed total lAE of 35.01.

Table 7.2 High Purity C3 SpHtter lAE indices for Product Variability Prediction
Configuration Bottoms Loop LAE
Overhead Loop LAE Total LAE
L,B

50.82

9.73

60.55

L,V

189.32

21.88

211.20

L,V/B

24.84

22.30

47.14

D,B

85.53

13.01

98.54

D,V

57.01

8.85

65.86

D,V/B

22.14

9.00

31.14

L/D,B

44.01

4.45

48.46

L/D,V

265.71

3.89

269.60

L/D,V/B

29.38

5.62

35.01

D , B Tight

40.85

5.13

45.99

7.2.2.2 Discussion
Based on the results of dual PI composition control of the high purity C3 spHtter,
it can be concluded that [D, B] (with tight level control), [D,V/B] and [L/D,V/B]
configurations provided the best control performance among aU configurations. Using the
approach shown in Figure 3.1, the [L/D,V/B] configuration resulted in low variabiHty for
both overhead product and bottoms product. It should be noted that for the high purity C3
splitter both overhead and bottoms product are of equal priorities. The [D,V/B]
configuration provided lower variabihty in bottoms product tiian [D,B] and [L/D,V/B]
configurations. However for overhead product the [L/D,V/B] configuration performed
much better than [D,B] (with tight level control) and [D,V/B] configurations. For setpoint
tracking the [D,B] (with tight level control) provided tiie best performance for both
overhead and bottoms impurity. For disturbance rejection of sinusoidal feed composition
136

The [D,V/B] configuration performed better for overhead impurity and the [L/D.V/B]
configuration performed better for bottoms impurity.
These results are not consistent with the results obtained by Hurowitz (1998)
based on his analysis of the high purity C3 splitter using a non-Hnear tray-to-tray model.
Hurowitz found [L,B], [L,V/B] and [D,B] (with tight level control) configurations to
provide the most optimal control performance for the high purity C3 spHtter. The [D,B]
configuration with tight level control is however not recommended due to its lack of
integrity as pointed out by Finco et al. (1989). The reason for mismatch is identical to that
for the base case C3 splitter. The mismatch between the linear model used here and the
non-Hnear model used by Hurowitz (1998) accounted for the inconsistency in the results.

7.2.3 Low Purity C3 SpHtter


7.2.3.1 Results
The results of the product variabiHty prediction technique for [L,B], [L,V] and
[L/D,B] configurations of the low purity design of C3 spHtter are shown in Figure 7.3.
These results were obtained by combining the frequency information shown in Figure 6.4
with the closed-loop Bode plot information for feed composition disturbance shown in
Figures 5.14 and 5.15, according to the Equation (7.12).
From the Figure 7.3, it can be seen that tiie product variabiHties were accurately
predicted for the low purity design of C3 spHtter technique for [L,B], [L,V] and [L/D,B]
configurations. For the remaining configurations too, tiie product variabilities were
accurately predicted. Analogous to the C3 spHtter an initial mismatch between the results
from the nonHnear tray to tray simulator and the results from the prediction technique
was observed during the initial minutes for aU configurations. The reason for the initial
mismatch is identical to tiiat for the base case C3 spHtter. However the overall mismatch
between the results from the nonlinear tray to tray simulator and the results from the
prediction technique for the low purity C3 spHtter was much higher than that for the high
purity C3 splitter and the base case C3 spHtter. Especially the mismatch was higher for tiie
final minutes. It was due to the much higher variation of the bottoms (50%) and overhead

137

impurity (20%). But on the whole it can be concluded that a good prediction of the
closed-loop product variabiHties was obtained using the proposed technique.
2.12

1.88-f
0

1000

2000

3000

4000

5000

Time (minutes)
"Simulator Response """"Predicted Response

1000

2000

3000

4000

5000

Time (minutes)
"Simulator Response "~"Predicted Response

Figure 7.3 Closed-Loop Rejection of disturbance of the Low Purity C3 Splitter, (a) For
the [D,V] configuration.

138

o
E
3

Q.

E
C vP
(D o^

ao

Q.

TD
CD
(1>

x:
0)

>
O
0

1000

2000

3000

4000

5000

Time (minutes)
Simulator Response

1000

2000

Predicted Response

3000

4000

5000

Time (minutes)
Simulator Response

Predicted Response

Figure 7.3 Continued, (b) For the [D,B] (with tight level control) configuration.

139

o
E

2.06
2.04 -

Q.

2.02 -

C
^
CD o^
CL
O

1.98
(0
0)
SZ

1.96

<1>

>

1.94
0

1000

2000

3000

4000

5000

Time (minutes)
Simulator Response

1000

2000

3000

Predicted Response

4000

Time (minutes)
Simulator Response " ^ P r e d i c t e d Response

Figure 7.3 Continued, (c) For the [IVD,V] configuration.

140

5000

The overhead impurity and bottoms impurity LAE values obtained from the
product variabiHty prediction for the different configurations of the low purity design of
C3 SpHtter are shown in Table 7.3. From Table 7.3 it can be seen that the best overhead
loop LAE performance is offered by [L/D,B] configuration which exhibited an LAE of
50.44. The [L/D,V/B] and [L/D,V] configurations also provided good control
performance with lAE values of 84.44 and 81.91. The performance of [L,B]
configuration too was good with an LAE of 88.02. Among all configurations, the worst
control performance was exhibited by [D,B] (with tight level control) configuration with
an lAE of 457.68. On the other hand for the bottoms loop [L,V/B] and [D,V]
configurations provided the best control performance with LAE values of 446.28 and
434.04. The [L/D,V/B] configuration provided the next best control performance with an
lAE of 574.91. The [D,V/B] configuration also provided reasonable control performance
witii an LAE of 602.51. The [L/D,V] configuration provided poor bottoms loop control
performance but performed better than [D,B] and [D,B] (with tight level control)
configurations which resulted in the worst bottoms impurity control performance. For the
total lAE performance, [L,V/B] and [D,V] configurations resulted in the best control
performance with LAE values of 560.94 and 549.56. The total lAE performance of
[L/D,V/B] configuration too was good with an lAE of 656.83. The [D,B] and [D,B] (with
tight level control) configurations due their poor performance for both overhead and
bottoms loop resulted in the largest total lAE values of 1879.50 and 2005.20,
respectively.

7.2.3.2 Discussion
Based on the results of dual PI composition control of the low purity C3 spHtter it
can be concluded that [L, V/B], [D,V] and [L/D,V/B] configurations provided the best
control performance among all configurations. Using the approach shown in Figure 3.1,
the [L/D,V/B] configuration resulted in lowest variabihty for the overhead product and
the [D,V] configuration resulted in the lowest variabihty in the bottoms product. The
[L,V/B] configuration provided the next lowest variability in the bottoms and overhead
product. For setpoint tracking the [L,V/B] configuration provided the best performance
141

for both overhead and bottoms impurity. The [L/D,V/B] configuration provided better
control of overhead impurity than [D,V] configuration. For disturbance rejection of
sinusoidal feed composition the [L/D,V/B] configuration provided better performance
than [L,V/B] and [D,V] configuration. However the bottoms loop performance of
[L,V/B], [D,V] and [L/D,V/B] configurations was equivalent.

Table 7.3 Low Purity C3 SpHtter LAE indices for Product Variability Prediction
Configuration Bottoms Loop LAE
Overhead Loop LAE Total LAE
L,B

1173.67

88.02

1261.68

L,V

1360.81

137.38

1498.19

L,V/B

446.28

114.66

560.94

D,B

1661.50

218.00

1879.50

D,V

434.04

114.52

548.56

D,V/B

602.51

244.57

847.07

L/D,B

1223.88

50.44

1274.32

L/D,V

1464.66

84.44

1549.10

L7D,V/B

574.91

81.91

656.83

D , B Tight

1547.52

457.68

2005.20

These results are not fully consistent with the results obtained by Hurowitz (1998)
based on his analysis of the low purity C3 spHtter using a non-linear tray-to-tray model.
Hurowitz found [L,V/B] and [L/D,V/B] configurations to provide the most optimal
control performance for the low purity C3 spHtter which is consistent witii the results
obtained here. However he found the performance of [D,V] configuration to be poor as
compared to [L,V/B] and [L/D,V/B] configurations. The reason for overall mismatch
between the results is identical to that for the base case C3 spHtter. The mismatch between
the Hnear model used here and the non-Hnear model used by Hurowitz (1998) accounted
for the inconsistency in the results.

142

7.2.4 Inverted Purity CT. SpHtter


7.2.4.1 Results
The results of the product variabiHty prediction technique for [L,B], [L,V] and
[L/D,B] configurations of the inverted purity design of C3 spHtter are shown in Figure
7.4. These results were obtained by combining the frequency information shown in
Figure 6.4 with the closed-loop Bode plot information for feed composition disturbance
shown in Figures 5.19 and 5.20, according to the Equation (7.12).
From the Figure 7.4, it can be seen that the product variabiHties were accurately
predicted for the inverted purity design of C3 splitter for [L,B], [L,V] and [L/D,B]
configurations. For the remaining configurations too the product variabilities were
accurately predicted. An initial mismatch between the results from the nonHnear tray to
tray simulator and the results from the prediction technique was observed during the
initial minutes for aU configurations. The reason for the initial mismatch is identical to
that for the base case C3 spHtter. But on the whole it can be concluded that an exceUent
prediction of the closed-loop product variabiHty predictions was obtained using the
proposed technique.
The overhead impurity and bottoms impurity LAE values obtained from the
product variabiHty prediction for the different configurations of the inverted purity design
of C3 SpHtter are shown in Table 7.4. From Table 7.4, it can be seen that [L/D,B]
configuration provided the best overhead loop LAE performance with an LAE of 44.36.
The [L/D,V] configuration too provided good control performance with an LAE of 63.42.
The performance of [L,B] configuration too was good with an LAE of 67.39. The [L,V/B]
configuration provided poor control performance with an LAE of 227.54. The overhead
loop performance of [D,V/B] configuration was worst with an lAE of 448.38. On the
other hand for tiie bottoms loop the [D,B] (witii tight level control) configuration
provided the best control performance with an LAE of 125.46. The [L,V/B] configuration
provided the next best control performance with an lAE of 144.15. The [L/D,V/B] and
[L,B] configurations provided good control performance with LAE values of 155.92 and
151.59, respectively. The [L/D,V] configuration performed reasonably witii an lAE of
168.24 but its performance was better than botii [D,B] and [L,V] configurations. The
143

[L,V] configuration resulted in poor performance with an LAE of 199.96. Its performance
was only better than [D,B] which showed an lAE of 199.96, the worst among all
configurations. For the inverted purity C3 splitter it should be reaHzed that the bottoms
product is of higher priority than the overhead product.
2.06

1.94
0

1000

2000

3000

4000

5000

Time (minutes)
Simulator Response

Predicted Response

0.45

m 0.15
0

1000

2000

3000

4000

5000

Time (minutes)
Simulator Response

Predicted Response

Figure 7.4 Closed-Loop Rejection of disturbance of the Inverted Purity C3 Splitter (a) For
the [L,B] configuration.
144

,0)

2.15

o
E
3

a
E
C >o

to o^
Q.
O

T3
TO
0)
.C
0)
>

1000

2000

3000

4000

5000

Time (minutes)
Simulator Response ~"~Predicted Response

<5^
0)

'

0.45
0.4

>s
^ H *

~i

o.

0.35

E
0)

0.3

c
0)
>N

o 0.25
^
CL
(0

0.2

o
*^
o
m 0.15

1000

2000

3000

4000

5000

Time (minutes)
Simulator Response

'Predicted Response

Figure 7.4 Continued, (b) For the [D,B] (with tight level control) configuration.

145

2.04

1000

2000

3000

4000

5000

Time (minutes)
Sinnulator Response ""Predicted Response

0.45
(1>

o
N

E
^

>%
*.*
3

0.4
0.35

Q.

E
0)

0.3

c
<D
>s

Q.

0.25

k..

Q.
SUJ

0.2

O
*j

o 0.15

CD

1000

2000

3000

4000

5000

Time (minutes)
Simulator Response

Predicted Response

Figure 7.4 Continued, (c) For the [L/D,B] configuration.

146

Table 7.4 Inverted Purity C3 splitter lAE indices for Product Variability Prediction
Configuration
Bottoms Loop lAE
Overhead Loop L\E
UB

151.59

67.39

L,V

199.96

119.58

L,V/B

144.15

227.54

D,B

226.16

189.49

D,V

204.21

283.54

D,V/B

178.52

448.38

L/D,B

188.72

44.36

L/D,V

168.24

63.42

UDy/B

155.92

115.42

D,B Tight

125.46

129.90

7.2.4.2 Discussion
Based on the results of dual PI composition control of the low purity C3 spHtter it
can be concluded tiiat [L, B], [L,V/B] and [D,B] (witii tight level contt-ol) configurations
provided the best control performance among all configurations. Using the approach
shown in Figure 3.1, the [D, B] configuration (with tight level control) resulted in lowest
variabiHty for the bottoms product which is of higher priority than the overhead product.
The next best variabiHty for the bottoms product was offered by [L,V/B] configuration.
However the [L,B] configuration performed much better for the overehad product than
[D,B] (with tight level control)and [L,V/B] configurations. For setpoint tracking, the
[L,V/B] configuration provided the best performance for the bottoms impurity but at the
expense of overhead impurity. The [L,B] configuration again provided better control of
overhead impurity than [D,B] (with tight level control) and [L,V/B] configurations.
Similar trend was observed for disturbance rejection of sinusoidal feed composition.
These results are fully consistent with the results obtained by Hurowitz (1998)
based on his analysis of the inverted purity C3 splitter using a non-linear tray-to-tray
model. Hurowitz found [L,B] and [L,V/B] configurations to provide the most optimal
control performance for the high purity C3 spHtter. He also found the performance of
147

[D,B] (with tight level control) configuration to be good but did not recommend it due to
its lack of integrity as pointed out by Finco et al. (1989). Thus it can be seen that the
product variability predicted using the approach shown in Figure 3.1 acts as a direct
measure of tiie controller performance of any distillation configuration. Hence the
approach can be used as the basis for identifying the best configuration suitable for
control of a distillation column.

7.2.5 Base Case Depropanizer


7.2.5.1 Results
The results of the product variabiHty prediction technique for [L,B], [L,V] and
[L/D,V] configurations of the base case design of depropanizer are shown in Figure 7.5.
These results were obtained by combining the frequency information shown in Figure 6.5
with the closed-loop Bode plot information for feed composition disturbance shown in
Figures 7.24 and 7.25, according to the Equation (7.12).
From the Figure 7.5, it can be seen that the product variabilities were accurately
predicted for the base case design of depropanizer for [L,B], [L,V], and [L/D,V]
configurations. For the remaining configurations too the product variabiHties were
accurately predicted. There was a relatively small degree of mismatch in the prediction,
which was due to the process non-linearities. For each configuration the mismatch
between the overhead and bottoms impurity was comparable. It should be reaHzed that
analogous to the C3 splitter an appreciable initial mismatch was observed between the
results from the nonlinear tray to tray simulator and the results from the prediction
technique. This was observed for all nine configurations. From Figure 7.5, it can be seen
that there is a mismatch for the first 200 minutes for [L,B], [L,V] and [D,V]
configurations. The initial mismatch is observed because the mean of the propane feed
composition signal is 32.51% whereas the steady state value of the propane feed
composition for the nonlinear tray to tray simulator is 31.54%. The depropanizer
nonHnear simulator takes the initial minutes to suppress this deviation. It should also be
realized that the amplitude and phase spectra in the closed-loop Bode plot correspond to
the steady state responses of the process at different frequencies and do not account for
148

their initial transients. From Figure 7.5, it can be seen that among the tiiree
configurations, the overall mismatch was the highest for [L,V] configuration and the
lowest for the [L,B] configuration. The overall mismatch of [L/D,V] configuration was
somewhat midway between [L.B] and [L,V] configuration. This is due to the relatively
high non-Hnear behavior of the [L,V] configuration as compared to the [L,B]
configuration and [L/D, V] configuration. But on the whole, an excellent prediction of the
closed-loop product variabiHties was obtained by the proposed technique due to the
relatively small deviation of the feed composition from the steady state value.
The overhead impurity and bottoms impurity LAE values obtained from the
product variability prediction for the different configurations of the base case design of
depropanizer are shown in Table 7.5. The LAE values are a measure of the average
variabiHty in the products. From Table 7.5, it can be seen that the overhead loop LAE
performance is best for [L/D,V/B] and [L,V/B] configurations indicating that the
variabihty in the overhead product is the lowest for these configurations. The [L/D,V]
configuration too provided good overhead impurity control performance with an LAE of
40.92. The performance of [L,V] and [L/D,V] configurations was comparable to that of
[L,V].The [D,B] and [D,V/B] configurations provided poor overhead impurity control
witii lAE values close to 80. However they performed better than [D,V] which performed
worst among aU configurations with an LAE of 107.15. Looking at the bottoms loop LAE
values, it can be concluded that [L/D,V/B] configuration again provided the best
performance with an lAE of 32.61. [L,V/B] and [L/D,V] configurations also provided
good control with lAE values of 36.58 and 37.75 close to that of [UDy/B]. The
performance of [D,V/B] was also good and comparable to that of [L/D,V]. The [D,V]
configuration, which performed worst for overhead impurity, performed reasonably witii
an lAE of 64.92. The worst performance for the bottoms loop was that of [L,B] and
[L/D,B] configurations which provided lAE values of 92.51 and 94.43, respectively. The
[L/D,V/B] configuration which performed best for both overhead and bottoms provided
the least total LAE of 61.95. The [L,V/B] configuration too provided good lAE contt-ol
performance with an lAE of 66.42. The [D,V] configuration resulted in the worst total

149

lAE performance with an lAE of 172.07. The [D,B] configuration which performed
poorly for both loops too resulted in poor total LAE performance with an LAE of 146.88.
0.65

0.40
0

200

400

600

800

1000

Time (minutes)
" Simulator Response " " " " Predicted Response

.2
o
E
3

Q.

E
"o
3
TJ
O
V)

E
o
o

CO

200

400

800

600

1000

Time (minutes)
"Simulator Response

'Predicted Response

Figure 7.5 Closed-Loop Rejection of disturbance of the Basecase Depropanizer. (a) For
the [L,B] configuration.

150

0.65

200

400

800

600

1000

Time (minutes)
"Simulator Response

200

400

Predicted Response

800

600
Time (minutes)

Simulator Response

'Predicted Response

Figure 7.5 Continued, (b) For the [L,V] configuration.

151

1000

0.65

200

400

600

800

1000

Time (minutes)
Simulator Response

Predicted Response

0.65

0.30
0

200

400

600

800

Time (minutes)
Simulator Response

'Predicted Response

Figure 7.5 Continued, (c) For the [L/D,V] configuration.

152

1000

Table 7.5 Base case Depropanizer LAE indices for Product Variability Prediction
Configuration Bottoms Loop LAE
Overhead Loop lAE Total lAE
L,B

92.51

45.52

138.02

L,V

53.74

43.26

97.01

L,V/B

36.58

29.84

66.42

D,B

71.31

75.57

146.88

D,V

64.92

107.15

172.07

D,V/B

39.22

80.18

119.40

L7D,B

94.63

42.43

137.06

L/D,V

37.75

40.92

78.67

UDy/B

32.61

29.34

61.95

7.2.5.2 Discussion
Based on the results of dual PI composition conttol of base case depropanizer, it
can be concluded that [L/D,V/B] and [L,V/B] configurations provided the best overall
control performance among aU configurations. Using the approach shown in Figure 3.1,
the [L/D,V/B] and [L,V/B] configurations resulted in the least variabiHty in the
distillation products. For setpoint tracking, the [L/D,V/B] configuration however
performed better that [L,V/B] configuration. For sinusoidal feed composition disturbance
rejection, [L/D,V/B] and [L,V/B] configurations both provided the best conttoUer
performance for both overhead and bottoms impurity. The [L,V] configuration performed
poorly for setpoint ttacking but resulted in good controller performance for the
disturbance rejection of the industrial feed composition signal. Among the configurations
not using reflux or boilup ratios, [L,V] configuration provided the lowest product
variability.
These results are in perfect agreement with the results obtained by Duvall (1999)
for the base case design of depropanizer. He also found [L/D,V/B] and [L,V/B]
configurations to be the best overaU for conttol of the base case depropanizer. He also
highhghted the improved controUer performance of the [L,V] configuration for feed
composition disturbance rejection in conttast to setpoint tracking. Thus it can be seen that
153

the product variability predicted using the approach shown in Figure 3.1 acts as a direct
measure of the conttoller performance of any distillation configuration. Hence the
approach can be used as the basis for identifying the best configuration suitable for
conttol of a distillation column.

7.2.6 High Purity Depropanizer


7.2.6.1 Results
The results of the product variabiHty prediction technique for [L,B], [L,V] and
[L/D,V] configurations of the high purity design of depropanizer are shown in Figure 7.6.
These results were obtained by combining tiie frequency information shown in Figure 6.5
witii the closed-loop Bode plot information for feed composition dismrbance shown in
Figures 5.29 and 5.30, according to the Equation (7.12).
From the Figure 7.6, it can be seen that the product variabiHties were accurately
predicted for the high purity design of depropanizer for P^,B], [L,V] and [L/D,V]
configurations. For the remaining configurations too the product variabiHties were
accurately predicted. Due to process non-Hnearities, there was a relatively small degree of
overall mismatch. Analogous to the base case depropanizer, an initial mismatch between
the results from the nonHnear ttay to ttay simulator and the results from the prediction
technique was observed for all configurations. The initial mismatch was observed for the
first 200 minutes for [L,B] and [L,V] configurations and for the first 300 minutes for the
[L/D,V] configuration. The reason for the initial mismatch was the same as that for the
base case design of the depropanizer. The overall mismatch between the results from the
nonHnear ttay to ttay simulator and the results from the prediction for the high purity
depropanizer was comparable to that for the base case depropanizer. Analogous to the
base case depropanizer, the overall mismatch was higher for [L,V] configuration as
compared to the [L,B] and [L/D,V] configurations. But on the whole it can be concluded
that an excellent prediction of the closed-loop product variabiHties was obtained using the
proposed technique.
The overhead impurity and bottoms impurity LAE values obtained from the
product variabihty prediction for the different configurations of the high purity design of
154

depropanizer are shown in Table 7.6. It should be noted that the lAE values are a measure
of the average variability in the products. From Table 7.6, it can be observed that the best
overhead loop lAE performance was for [L,V/B] configuration with an L\E of 7.24.
[UDy/B] configuration provided the next best performance with an lAE of 9.03.The
performance of [L,B], [L,V] and [L/D,V] configuration was close to that of [L/D,V/B]
which showed lAE values of 10.6, 9.91 and 10.8, respectively. The [D,B], [D,V] and
[D,V/B] configurations performed worst with L\E values around 30. The bottoms loop
performance of [L/D,V] configuration was the best among all configurations with an lAE
of 16.88. The [L,V/B] configuration provided close enough performance with an LAE of
17.60. The performance of [L,V] and [UDy/B] configuration was also good with an
LAE of 19.96 and 19.63, respectively. The worst performance for bottoms loop was
exhibited by [D,B]configuration. Since the [L,V/B] configuration performed best for both
bottoms and overhead loop, it provided the best total lAE performance with an lAE of
24.84. The [L,V], [L/D,V] and [UDy/B] configurations too provided good total lAE
performance with an LAE values around 28. The [D,B] configuration resulted in the worst
total LAE performance due to its poor performance for both bottoms and overhead.

7.2.6.2 Discussion
Based on the results of dual PI composition conttol of high purity depropanizer, it
can be concluded that [L/D,V/B], [L/D,V] and [L,V/B] configurations provided the best
control performance among all configurations. Using the approach shown in Figure 3.1,
these configurations resulted in the least variabiHty in the distillation products. The
variability in the overhead product was slightly better for [L,V/B] than that of [L/D,V]
and [UDy/B] configurations. However, the variabiHty in the overhead product was
slightly better for [L/D,V] configuration as compared to that of [L,V/B] and [L/D,V/B]
configurations. Similar trend was observed for setpoint ttacking. For sinusoidal feed
composition disturbance rejection, [UDy/B] and [L,V/B] configurations both provided
the better conttol performance than. [L/D,V] configuration. Except for setpoint ttacking
the performance of [L,V] configuration for the industrial feed composition signal and

155

sinusoidal feed composition disturbance rejection was comparable to any of these


configurations.
0.15

0.05
200

400

600

800

1000

Time (minutes)
"Simulator Response

'Predicted Response

0.20

0.00
0

200

400

800

600

1000

Time (minutes)
Simulator Response

'Predicted Response

Figure 7.6 Closed-Loop Rejection of disturbance of the High Purity Depropanizer. (a)
For the [L,B] configuration.
156

0.15

o
E 0.13

I 0.1 M
o

0.09

T3

0.07

(1>

>

0.05
0

200

400

800

600

1000

Time (minutes)
"Simulator Response

'Predicted Response

0.20

0.00
0

200

400

800

600
Time (minutes)

Simulator Response

'Predicted Response

Figure 7.6 Continued, (b) For the [L,V] configuration.

157

1000

0.13

600
Time (minutes)
"Simulator Response

'Predicted Response

0.16

0.04
0

200

400

600

800

Time (minutes)
Simulator Response ^ " " P r e d i c t e d Response

Figure 7.6 Continued, (c) For the [L/D,V] configuration.

158

1000

Table 7.6 High Purity Depropanizer lAE indices for Product VariabiHty Prediction
Configuration Bottoms Loop LAE
Overhead Loop lAE Total lAE
L,B

30.43

10.60

41.03

L,V

19.96

9.91

29.87

L,V/B

17.60

7.24

24.84

D,B

49.56

30.27

79.83

D,V

29.05

30.44

59.49

D,V/B

35.30

30.47

65.77

L/D,B

38.98

13.19

52.17

L/D,V

16.88

10.80

27.69

UDy/B

19.63

9.03

28.66

These results are also consistent with the results obtained by DuvaU (1999) for the
high purity design of depropanizer. He too found [L/D, V/B] and [L/D,V] configurations
to be the best overall control of the high purity depropanizer. He also pointed out the
excellent conttoller performance of the [L,V] configuration for feed composition
disturbance rejection. However he found the conttol performance of [L,V/B]
configuration to be poor. On the whole, it can be said that the results obtained by DuvaU
matched up well with the results obtained here. Thus it can be seen that the product
variability predicted using the approach shown in Figure 3.1 acts as a direct measure of
the conttoller performance of any distillation configuration. Hence the approach can be
used as the basis for identifying the best configuration suitable for control of a distillation
column.

7.2.7 Low Purity Depropanizer


7.2.7.1 Results
The results of the product variabihty prediction technique for [L,B], [L,V] and
[L/D,V] configurations of the low purity design of depropanizer are shown in Figure 7.7.
These results were obtained by combining the frequency information shown in Figure 6.5

159

with the closed-loop Bode plot information for feed composition disturbance shown in
Figures 5.34 and 5.35, according to the Equation (7.12).

2.40

1.60
0

200

400

600

800

1000

Time (minutes)
'Predicted Response

"Simulator Response

2.50

1.50
0

200

400

800

600

1000

Time (minutes)
"Simulator Response

'Predicted Response

Figure 7.7 Closed-Loop Rejection of disttirbance of the Low Purity Depropanizer. (a) For
the [L,B] configuration.
160

2.20

1.80
0

200

400

600

800

1000

Time (minutes)
Simulator Response

'Predicted Response

2.30

1.50
0

200

400

800

600
Time (minutes)

Simulator Response

'Predicted Response

Figure 7.7 Continued, (b) For the [L,V] configuration.

161

1000

2.20 -r-

600

800

Time (minutes)
"Simulator Response

Predicted Response

2.30

1.50
0

200

400

800

600

1000

Time (minutes)
"Simulator Response

'Predicted Response

Figure 7.7 Continued, (c) For the [L/D,V] configuration.


From the Figure 7.7, it can be seen that the product variabilities were accurately
predicted for the low purity design of depropanizer for [L,B], [L,V] and [L/D,V]
configurations. For the remaining configurations too the product variabiHties were
162

accurately predicted. Similar to the base case depropanizer, an initial mismatch between
the results from the nonlinear tray to ttay simulator and the results from the prediction
technique was observed for all configurations. For the [L,B] and [L,V] configurations, the
mismatch was observed for the first 200 minutes whereas for the [L/D,V] configuration
tiie mismatch was observed for the first 100 minutes. The reason for the initial mismatch
is same as tiiat for the base case design of the depropanizer. The overaU mismatch
between the results from the nonlinear ttay to ttay simulator and the results from the
prediction technique for the low purity depropanizer was comparable to that for both high
purity depropanizer and the base case depropanizer. Analogous to the base case
depropanizer the [L,V] configuration resulted in the worst mismatch among the three
configurations due to its high non-Hnear behavior. For the three configurations, the
mismatch in the overhead and bottoms product were comparable. But on the whole, it can
be concluded that even in presence of non-Hnearities, an excellent prediction of the
closed-loop product variabiHties was obtained using the proposed technique.
The overhead impurity and bottoms impurity LAE values obtained from the
product variabiHty prediction for the different configurations of the low purity design of
depropanizer are shown in Table 7.7. It should be noted that the LAE values are a measure
of the average variability in the products. From Table 7.7, it can be seen that the best
overhead loop LAE performance was offered by [L,V] and [L,V/B] configurations with
LAE values of 66.70, and 66.81, respectively. The [L,B] and [L/D,V/B] configurations
provided the next best performance with LAE values of 72.08 and 72.86. The [L,B]
configuration provided good conttol performance with an lAE of 96.91. The [D,B], and
[D,V] configurations performed worst with LAE values around 200. The bottoms loop
performance of [L/D,V] configuration was the best among aU configurations with an LAE
of 70.07. The [L/D,V/B] configuration provided close enough performance with an LAE
of 74.55. The performance of [L,V] and [L,V/B] configuration was also good with LAE
values of 85.24 and 83.12, respectively. The worst performance for bottoms loop was
exhibited by [D,B]configuration. Since the [L/D,V] configuration performed good for
both bottoms and overhead loop, it provided the best total LAE performance witii an LAE
of 142.15. The [L,V], [L,V/B] and [UDy/B] configurations too provided good total lAE
163

performance with lAE values around 150. The [D,B] configuration resulted in the worst
total LAE performance due to its poor performance for both bottoms and overhead loop.

Table 7.7 Low Purity Depropanizer LAE indices for Product Variability Prediction
Configuration Bottoms Loop LAE
Overhead Loop LAE Total L\E
L,B

195.53

96.91

292.44

L,V

85.24

66.70

151.95

L,V/B

83.12

66.81

149.92

D,B

180.82

212.65

393.46

D,V

104.52

202.10

306.62

D,V/B

123.78

177.85

301.63

L/D,B

200.31

87.80

288.11

L/D,V

70.07

72.08

142.15

UDy/B

74.55

72.86

147.41

7.2.7.2 Discussion
Based on the results of dual PI composition conttol of low purity depropanizer it
can be concluded that [UDy/B] and [L/D,V] configurations resulted in the best conttol
performance. Except for setpoint ttacking, the [L,V] configuration too provided excellent
control performance. Using the approach shown in Figure 3.1, the [L,V], [L/D,V/B] and
[L/D,V] configuration resulted in the least variabiHty in the distiUation products. The
[L,V] configuration however provided lower variabihty in overhead product than
[UDy/B] and [L/D,V] configuration. For the bottoms product the [L/D,V] configuration
provided lower variability than [L,V] and [UDy/B] configurations. For setpoint ttacking
however the [L,V] configuration resulted in the worst conttol performance. The
[UDy/B] configuration performed better than [L/D,V] configuration. For sinusoidal
feed composition disturbance rejection, [L,V] configuration performed better than
[L/D,V/B and [L7D,V] configuration for overhead impurity conttol. But for bottoms
impurity conttol the [L,V] configuration performed worse than [L/D,V/B] and [L/D,V]
configuration.
164

These results are analogous to the results obtained by Duvall (1999) for the low
purity design of depropanizer. He too found [L/D, V/B] and [L/D, V] configurations to be
the best for the conttol of the low purity depropanizer. He also observed the conttol
performance of [L,V] configuration to be good for feed composition disturbance rejection
but exttemely poor for setpoint ttacking. Thus it can be seen that the product variabihty
predicted using the approach shown in Figure 3.1 acts as a direct measure of tiie
conttoller performance of any distillation configuration. Hence the approach can be used
as tiie basis for identifying the best configuration suitable for conttol of a distiUation
column.

7.2.8 Asymmetric Purity Depropanizer


7.2.8.1 Results
The results of the product variabiHty prediction technique for [L,B], [L,V] and
[L/D, V] configurations of the asymmetric purity design of depropanizer are shown in
Figure 7.8. These results were obtained by combining the frequency information shown
in Figure 6.5 with the closed-loop Bode plot information for feed composition
disturbance shown in Figures 5.39 and 5.40, according to the Equation (7.12).
From the Figure 7.8, it can be seen that the product variabiHties were accurately
predicted for the asymmetric purity design of depropanizer for [L,B], [L,V] and [L/D,V]
configurations. For the remaining configurations too the product variabiHties were
accurately predicted. An initial mismatch between the results from the nonlinear ttay to
ttay simulator and the results from the prediction technique was observed for all
configurations. For the [L,B] and [L,V] configurations the mismatch was observed for the
first 200 minutes whereas for the [L/D,V] configuration the mismatch was observed for
the first 100 minutes. The reason for the initial mismatch is identical to that for the base
case design of the depropanizer. The overall mismatch between the results from the
nonHnear ttay to ttay simulator and the results from the prediction technique was
comparable to that of the base case depropanizer. Analogous to the base case
depropanizer the [L,V] configuration resulted in the worst mismatch among the three
configurations due to its high non-linear behavior. For each configuration the mismatch
165

between the overhead and bottoms impurity was comparable. But on the whole it can be
concluded that, even in the presence of non-linearities, an excellent prediction of the
closed-loop product variabiHties was obtained using the proposed technique.
The overhead impurity and bottoms impurity lAE values obtained from the
product variabiHty prediction for tiie different configurations of tiie asymmettic purity
design of depropanizer are shown in Table 7.8. From Table 7.8, it can be seen that
[L,V/B] and [L,V] configurations provided the best overhead loop lAE performance with
lAE values of 3.57 and 3.68, respectively. The [L/D,V/B] configuration too provided
good control with an lAE of 4.92. Surprisingly, the [L/D,V] configuration provided only
reasonable performance witii an LAE of 7.43. The worst overhead impurity conttol
performance was exhibited by [D,V] configuration. For the bottoms loop the [L,V]
configuration provided the best conttol performance with an lAE of 85.22. The [L/D,V]
and [UDy/B] configurations provided comparable performance with LAE values of
96.49 and 92.34, respectively. The performance of [L,V/B] and [D,V/B] configurations
was also good witii lAE values of 102.87 and 103.29. The [D,V] configuration provided
poor overhead impurity conttol performance but did better than [D,B] configuration
which performed worst among all configurations.

7.2.8.2 Discussion
Based on the results of dual PI composition control of asymmetric purity
depropanizer, it can be concluded that [L/D,V/B] and [L/D,V] configurations provided
the best control performance among all configurations. The [L,V] configuration too
provided excellent conttol performance except for setpoint ttacking. Using the approach
shown in Figure 3.1, the [L,V], [UDy/B] and [L/D,V] configuration resulted in the least
variability in the distillation products. The [L,V] configuration however provided slightly
lower variability than [L/D,V/B] configuration which in turn provided sHghtly lower
variability than [L/D,V] configuration. For setpoint ttacking however, the [L,V]
configuration resulted in the worst conttol performance. The [L/D,V/B] configuration
performed

better than [L/D,V] configuration. For sinusoidal feed composition

disturbance rejection, [L,V] configurations performed better than [UDy/B and [L/D,V]
166

configuration for overhead impurity conttol. For bottoms impurity control [L/D,V]
configuration performed worse than [L,V] and [IVD,V] configuration.
These results are in perfect agreement with the results obtained by DuvaU (1999)
for tiie asymmettic design of depropanizer. He also found [L/D,V/B] and [L/D,V]
configurations to be the best for conttol of the asymmetric depropanizer. Except for
setpoint ttacking he found the conttol performance of [L,V] configuration to be better
than [UDy/B]

and [L/D,V] configurations. Thus it can be seen that the product

variability predicted using the approach shown in Figure 3.1 acts as a direct measure of
the conttoller performance of any distillation configuration. Hence the approach can be
used as the basis for identifying the best configuration suitable for conttol of a distillation
column.

Table 7.8 Asymmetric Purity Depropanizer LAE indices for Product Variability
Prediction
Overhead Loop LAE Total lAE
Configuration Bottoms Loop LAE
L,B

245.69

8.89

254.58

L,V

85.22

3.68

88.90

L,V/B

102.87

3.57

106.44

D,B

197.61

14.58

212.19

D,V

165.42

22.25

187.67

D,V/B

103.29

16.80

120.08

L/D,B

221.29

8.28

229.57

L/D,V

96.49

7.43

103.92

UDy/B

92.34

4.92

97.26

167

0.13
o
E
3

0.11 -

a
E
o

3
"D
O

QI

0.09 -

(0

a>

>
O

0.07
0

200

400

600

800

1000

Time (minutes)
"Simulator Response ^""Predicted Response

3.00

1.00
0

200

400

600

800

1000

Time (minutes)
"Simulator Response " " " P r e d i c t e d Response

Figure 7.8 Closed-Loop Rejection of disturbance of the Asymmetric Purity Depropanizer.


(a) For the [L,B] configuration.

168

0.11

0.08
0

200

400

600

800

1000

Time (minutes)
'Predicted Response

"Simulator Response

2.50

1.50
0

200

400

800

600
Time (minutes)

"Simulator Response

'Predicted Response

Figure 7.8 Continued, (b) For the [L,V] configuration.

169

1000

0.13
o
E
3
D.

0.11 -

E
"o
3
D
O

QI

0.09 -

T3
(C
0
SZ

>
O

0.07
0

200

400

800

600

1000

Time (minutes)
"Simulator Response

'Predicted Response

2.50

1.50
0

200

400

600

800

Time (minutes)
"Simulator Response ""^Predicted Response

Figure 7.8 Continued, (c) For the [L/D,V] configuration.

170

1000

7.3 Summary
The prediction technique shown in Figure 3.1 was demonstrated for two
distillation columns, C3 spHtter and depropanizer. Four different designs of these columns
were analyzed. Using the closed-loop Bode plots generated in Chapter 5 and frequency
information generated from the industtial feed composition signals in Chapter 6, closedloop product variabilities were obtained for the different configurations of the two
distillation columns. For each column the prediction technique was verified by comparing
the predictions with the results obtained by inputting the industrial feed composition
signal to a non-linear ttay-to-ttay simulator (Gokhale et al., 1994; Duvall, 1999) of that
column. The product variabilities were accurately predicted for both columns. An initial
mismatch was however observed between the predictions and the product variabiHties
that resulted from the industrial feed composition signal. The overall mismatch was
found to be higher for higher variation in product purities. For depropanizer however the
mismatch was higher than C3 spHtter due to the higher non-Hnearity of depropanizer for
feed composition disturbances as compared to the C3 splitter.
The potential of the approach as a tool for configuration selection was also
evaluated. For depropanizer, the product variabiHty prediction results of the different
configurations were consistent with tiie relative conttol performance obtained by Duvall
(1999) for tiie different configurations. The configurations recommended by Duvall
(1999) for tiie different designs resulted in the lowest product variabiHties. Thus showing
that product variabiHties obtained using the prediction technique directly correspond to
the conttol performance of each configuration. Thus cHciting the utility of the approach
as a quantitative tool for configuration selection. For the C3 spHtter, however, only the
inverted purity design provided results, which corresponded to those obtained by
Hurowitz (1998) for tiie different configurations. For the base case, high purity and low
purity designs of C3 splitter the results did not match well with tiiose obtained by
Hurowitz (1998) due to the high non-Hnearity associated with these designs. In fact the
results obtained here were accurate for smaller deviations of the columns. However,
Hurowitz (1998) tuned the designs for higher deviations around which the linear model is
no longer accurate. These results exposed tiie shortcoming associated with the approach.
171

Since the tuning of the composition loops is carried out using a Hnear model of the
column, the accuracy of the results is contingent upon the accuracy of the linear model
for setpoint changes. On the whole it can be said that the proposed technique offered a
quantitative tool for evaluating the performance of different configurations. Thus it can
used to identify the best configuration for a distillation column.

172

CHAPTER 8
CONCLUSIONS AND RECOMMENDATIONS

8.1 Conclusions
A novel technique was developed which uses industrial feed composition signals
and Hnear models to predict distillation column closed-loop product variabiHties. This .
technique was demonstrated for a binary distillation column, a C3 spHtter and a
multicomponent distillation column, a depropanizer. Four different designs of C3 spHtter
namely. Base Case, High Purity, Low Purity and Inverted Purity and four different
designs of depropanizer, namely Base Case, High Purity, Low Purity and Asymmetric
Purity were considered for this purpose. The potential of the approach as a quantitative
tool for configuration selection was also evaluated for these columns.
The technique essentially predicts distillation product variabihty resulting from an
industrial feed composition signal. For this purpose, a linear model was first developed
using the steady state and dynamic characteristics of the distillation column. The
conttoller tuning parameters were then derived using a preset conttoller tuning criteria.
Using these conttoller tuning parameters a closed-loop Bode plot for the feed
composition disturbance was generated. Meantime, an industrial feed composition signal
was subjected to signal processing techniques to exttact its frequency information. The
technique essentially combines this frequency information with the closed-loop Bode plot
information for the feed composition disturbance to predict closed-loop product
variabiHty. The approach was verified by comparing it with the results from a non-Hnear
ttay-to-ttay simulator.
For both C3 spHtter and depropanizer, Hnear dynamic ttay-to-ttay models under
dual PI composition conttol were developed by Hnearizing the non-Hnear mass and
energy balance equations and VLE equations at each ttay. An invariant structure of the
distillation column was employed in both models. The level conttoller tuning parameters
were then obtained using MarHn (1995) settings for each configuration of the two
columns. A consistent tuning approach based on minimum lAE was developed for both
columns in order to be able to compare results between different control configurations.
173

The composition conttoller tuning parameters were tuned for minimum lAE in product
impurities for product impurity setpoint changes depending on the priority of the
overhead and bottoms product. Using these controller tuning parameters, closed-loop
Bode plots for feed composition disturbance were developed. Standard signal processing
techniques were appHed to the industtial feed composition signals of C3 spUtter and
depropanizer to exttact their ampHtude and phase spectta. These spectta were then
combined with the closed-loop Bode plot information to predict product variabiHties
according to the proposed technique. The prediction results were compared with the
results obtained by obtained by inputting the feed composition signal to the non-Hnear
tray-to-ttay simulators of C3 spHtter and depropanizer (Gokhale et al., 1995; DuvaU,
1999).
The product variabilities were accurately predicted for both columns. However an
appreciable mismatch between the results from the non-linear ttay-to-ttay simulator and
the results from the prediction technique was observed during the initial minutes. This
was due to the mismatch in the mean of the feed composition signal and the steady state
value of the feed composition for the non-Hnear ttay to ttay simulator. Also there was a
sHght overall mismatch between the results due to the process non-Hnearities. The
mismatch tended to increase for higher variation in the product purities. For
depropanizer, the product variabilities were accurately predicted for all four designs due
to the relatively small deviation of the product purities from setpoint. For the C3 spHtter
however, the low purity design resulted in large changes in the overhead and bottoms
impurity and hence amounted to the largest mismatch between the results from the nonHnear ttay-to-ttay simulator and the results from the prediction technique as compared to
the Base Case, High Purity and Inverted Purity designs. For the Base Case, High Purity
and Inverted Purity designs the overall mismatch was comparable. For depropanizer
however the mismatch was higher than C3 spHtter due to the higher non-Hnearity of
depropanizer for feed composition disturbances as compared to the C3 splitter.
The potential of the approach as a tool for configuration selection was also
evaluated for both columns. For depropanizer, the product variabihty prediction results of
the different configurations were consistent with the relative control performance
174

obtained by Duvall (1999) for the different configurations. For the four designs of
depropanizer the [UDy/B] and [L,V] configurations provided the least product
variabiHty as compared to the other configurations. These configurations were also
recommended by Duvall (1999) for the four different designs. Thus showing that product
variabilities obtained using the prediction technique directly correspond to the conttol
performance of each configuration. Thus ehciting the utiHty of the approach as a
quantitative tool for configuration selection. For the C3 spHtter, however, only the
inverted purity design provided results, which corresponded to those obtained by
Hurowitz (1998) for the different configurations. For the base case, high purity and low
purity designs of C3 splitter the results did not match well with those obtained by
Hurowitz (1998) due to the higher non-Hnearity associated with these designs for setpoint
changes. In fact the results obtained here were accurate for small deviations in the
process variables. However Hurowitz (1998) tuned the designs using larger magnitudes
of setpoint changes, around which the Hnear model is no longer accurate. These results
exposed the shortcoming associated with the approach. Since the tuning of the
composition loops is carried out using a Hnear model of the column, the accuracy of the
results is contingent upon the accuracy of the Hnear model.
Thus, the demonsttation of the concept for C3 spHtter and depropanizer
highUghted the two factors, which influence the accuracy of the approach: the nonHnearity of the process and the variation of the process variables due to the disturbance.
The higher the variation in the process variables (i.e., product purities) of the column
higher the inaccuracy of the predicted results. This was observed for the low purity C3
splitter for which the variation in the product purities was relatively high as compared to
the other designs. Lligher the non-Hnearity greater the inaccuracy of the approach. This
was seen for the depropanizer, which showed higher mismatch than C3 spHtter due to its
higher non-linearity. Also the utiHty of the approach as a tool for configuration selection
has been shown to be contingent upon the non-linearity of the process for setpoint
changes. This was observed for the Base Case, High Purity and Low Purity designs of the
C3 splitter which did not provide results consistent with Hurowitz (1998) due to the
higher non-linearity associated with these designs for setpoint changes. Since a Hnear
175

model is used to tune the loops, the approach seems to provide the least product
variability for configurations that can provide the best conttol performance for small
deviations in the process.

8.2 Recommendations
The following recommendations are proposed for further exploring the utiHty of
the product variability perdiction approach proposed here.
1. The product variabiHty prediction approach can be further analyzed for various
distillation columns such as vacuum distillation columns, reactive distillation columns
and azeotropic distillation columns.
2. The product variabiHty prediction approach is truly general and its utility can be
further explored for processes other than distillation.
3. The potential of the approach for configuration selection has been studied using
analytical simulations. It can be further verified by implementing the approach in the
chemical processing industries.
4. The approach seems to have potential for identifying distiUation columns working
below their potential and for identifying the sources of product variabiHty. These
avenues can be explored for the approach using analytical simulations and industrial
implementations.

176

LITERATURE CITED
Anderson, J.A., Vacuum Distillation Control, Ph.D. Dissertation, Texas Tech University,
Lubbock, TX, 1999.
Asttom, K.J., Hagglund, "Automatic Tuning of Simple Regulators with Specifications on
Phase and AmpHtude Margins," Automatica, Vol. 20, No. 2, 1984, 645-651.
Boston, J.F., SulHvan, S.L., "A New Class of Solution Methods for Multicomponent
Multistage Separation Processes," Canadian Journal ofChem. Engr., No. 52,
1974, 52-63.
Box, G.E., Jenkins, G.M., Reinsel, G.C., Time Series Analysis: Forecasting and Control,
3'^ ed.. Prentice Hall, Englewood CHffs, NJ, 1994.
Carling, G.A., Wood, R.K., "The Dynamics and Conttol of a Depropanizer," Dynamics
and Control of Chemical Reactors and Distillation, No. 4, 1988, 167-173.
Chiang, T., Luyben, W.L., "Incentives for Dual Composition Conttol in Single and Heat
Integrated Binary Distillation Columns," Ind. Eng. Chem. Fundamentals, Vol. 24,
No 3, 1985, 352-359.
Cwiklinski, R.R., Brosilow, C.B., "Inferential Conttol of an Industrial Depropanizer,"
1977 Joint Automatic Control Conference, San Francisco, CA, June 1977, 15301536.
Degoyler & McNaughton, 20th Century Petroleum Statistics, Dallas, 1989.
Duvall, P.M., On Control of High Relative Volatility Distillation Columns, Ph.D.
Dissertation, Texas Tech University, Lubbock, TX, 1999.
Eduljee, H.E., "Equations Replace GiUilands' Plot," Hydrocarbon Proc, 1975, 54(9),
120.
Fenske, M.R., Ind. Eng. Chem. Res., 1932, 24, 482.
Finco, M.V., The Modeling and Control of Low Relative Volatility Systems, M.S. Thesis,
Lehigh University, Bethlehem, PA, 1987.
Finco, M.V., Luben, W.L., PoUeck, R.E., "Conttol of Distillation Columns with Low
Relative VolatiHties," Ifid. Eng. Chem. Res., 1989, 28(1), 75-83.
Franks, R.G.E., Modeling and Simulation in Chemical Engineering, John Wiley & Sons,
New York, 1972.
177

Freitas, M.S., Campos M.C.M.M., Lima, E.L., "Dual Composition Conttol of a


Debutanizer Column," ISA Trans., Vol. 33, No. 1, 1994, 19-25.
Fruehauf P.S., Mahoney D.P., "Lnprove Distillation - Column Control Design," Chem.
Engr. Progress, Vol. 90, No. 3, 1994,75-83.
Fuentes, C , Luyben, W.L., "Control of High Purity Distillation Columns," Ind Eng.
Chem. Process. Des. Dev., Vol. 22, No. 3, 1983, 361-366
GiUiland, E.R., Ind Eng. Chem., 1940, 32, 1220.
Gokhale, V.B., Control of a Propylene/Propane Splitter, M.S. Thesis, Texas Tech
University, Lubbock, TX, 1994.
Gokhale, V.B., Hurowitz, S., Riggs, J.B., "A Comparison of Advanced Distillation
Control Techniques for a Propylene/Propane SpHtter," Ind. Eng. Chem. Res.,
1995, 34(12), 4413-4419.
Holland, CD., Fundamentals of Multicomponent Distillation, McGraw Hill, New York,
1981.
Hovd, M., Skogestad, S., "Simple Frequency Dependent Tools for Conttol System
Analysis, Structure Selection and Design,'' Automatica, Vol. 28, No. 5, 1992,
989-996.
Humphrey, J.L., Seibert A.F., Koort, R.A., "Separation Technologies - Advances and
Priorities," U.S Department of Energy Final Report, Conttact No. DE-AC0790DD12920, 1991.
Hurowitz, S.E., Superfractionator Process Control, Ph.D. Dissertation, Texas Tech
Univeristy, Lubbock, TX, 1998.
Joseph, B., Brosilow, C.B., "Inferential Conttol of Processes: Parts I and II," AIChE
Journal, Vol. 24, No. 3, 1978, 485-500.
Kamen, E.W., Introduction to Signals and Systems, 2"^^ ed., Macmillan PubHshing
Company, New York, 1990.
Kister, H.Z., Distillation Operation, McGraw HiU, New York, 1990.
Kister, H.Z., Distillation Design, McGraw HiU, New York, 1990.

178

Luyben W.L. "Steady State Energy Conservation Aspects of Distillation Column


Control System Design," Ind Eng. Chem. Fundamentals, Vol. 14, No. 4, 1975,

Luyben, W.L., Process Modeling, Simulation and Control for Chemical Engineers, 2"**
ed., McGraw Hill, New York, 1990.
Luyben, W.L., Practical Distillation Control, Van Nosttand Reinhold, New York, 1992.
Marlin, T.E., Process Control: Designing Processes and Control Systems for Dynamic
Performance, McGraw Hill, New York, 1995.
McAvoy T.J., Ye, N., Gang, C , "NonHnear Inferential Parallel Cascade Conttol," Ind
Eng. Chem. Res., Vol. 35, No. 1, 1996, 130-137.
Morari, M., "Design of Resihent Processing Plants-IE: A General Framework for tiie
Assessment of Dynamic Resilience," Chem. Eng. Set, Vol. 38, No. 11, 1983
1991-1891.
NiedeHnski, A., "A Heuristic Approach to the Design of Linear Multicariable Conttol
Systems,"" Autoinatica, No. 7, 1971, 691-695.
Rademaker, O.J., Rijnsdorp, A., Maarleveld, "Dynamics and Control of Continuous
Distillation \5mX.s," AIChE Journal, VoL 27, No. 11, 1991, 1634-1644.
Reid, R.C. Prausnitz, J.M., PoHng, B.E., The Properties of Gases and Liquids, 4* ed.,
McGraw Hill, New York, 1987.
Riggs, J.B., "Improve Distillation Column Conttol," Chemical Engineering Progress,
VoL 94, No. 10, 1998,31-47.
Shinskey, F. G., Distillation Control for Productivity and Energy Conservation, 2"^ ed.,
McGraw Hill, New York, 1984.
Skogestad, S., Morari, M. "Conttol Configuration Selection for Distillation Columns,"
AIChE Journal, Vol. 33, No. 10, 1987, 1620-1635.
Skogestad, S., Lundstrom, P, Jacobsen, E.W., "Selecting the Best Distillation Control
Configuration," A/C/zE/ownifl/, Vol. 36, No. 5, 1990, 753-764.
Smith, J.M., Van Ness, H.C., Introduction to Chemical Engineering Thermodynamics,
4th ed., McGraw HiU, New York, 1987.
Stanley, G.T., McAvoy, T.J., "Dynamic Energy Conservation Aspects of Distillation
Control," Ind. Eng. Chem. Fundamentals, Vol. 24, No. 4, 1985, 439-443.
179

Tseng, J.L., Cluett, W.R., Bialkowski, W.L., "New Analysis and Design Tool for
Achieving Low VariabiHty Process Designs," AIChE Journal, Vol. 45, 1999,
2188-2202.
Tyreus, B.D, Luyben, W.L., "Tuning PI ConttoUers for Integrator/Deadtime Processes,"
Ind. Eng. Chem. Res., Vol. 31, No. 11, 1992, 2525-2628.
Walas, S.M., Phase Equilibria in Chemical Engineering, Butterworth, Boston, 1985.
WaUcer, J.S., Fast Fourier Transforms, CRC Press, Boca Raton, FL, 1991.
Wolf, E.A., Skogestad. S., "Temperature Cascade Control of Distillation Columns," Ind.
Eng. Chem. Res., Vol. 35, No. 2, 1996, 475-484.
Yang, D.R., Waller K.V., Kurt V., Seborg D.E., MeUichamp, D.E. (1990). "Dynamic
Structural Transformations for Distillation Conttol Configuration," AIChE
Journal, 36, 1391-1402.
Zheng, A., Mahajanam, R.V., (1999). "A Quantitative ConttoUabiHty Index," Ind. Eng.
Chem. Res., 38, 999-1006
Ziegler, J.G, Nichols, N.B., "Optimum Settings for Automatic ConttoUers," Transactions
oftheASME, Vol. 64, 1942, 759-768.

180

PERMISSION TO COPY

In presenting tiiis tiiesis in partial fiilfiUment of tiie requirements for a master's


degree at Texas Tech University or Texas Tech University Health Sciences Center, I
agree tiiat tiie Library and my major department shaU make itfreelyavailable for
research purposes. Permission to copy this thesis for scholarly purposes may be
granted by tiie Director of the Library or my major professor. It is understood tiiat
any copying or publication of this thesis for financial gain shaU not be aUowed
v^thout my further written permission and that any user may be liable for copyright
infringement.

Agree (Permission is granted.)

Student Signature

Date

Disagree (Permission is not granted.)

Student Signature

Date

Anda mungkin juga menyukai