Air Separation Unit
Air Separation Unit
Air Separation Unit
TED JOHANSSON
by
Ted Johansson
I detta examensarbete presenteras en ny metod för att göra schemaläggningsbeslut inom processindustrin
och samtidigt ta hänsyn till det dynamiska beteendet hos processen. En modell av en luftseparerings-
anläggning som producerar kvävgas och utnyttjar ett rörligt elpris användes för att exemplifiera denna
metod. Modellen omfattade en kryogenisk destillationskolonn med en integrerad återloppskokare /kon-
densator, en multiströms värmeväxlare, en kompressor, två turbiner och en kondensator. Den innehöll
5079 ekvationer och 437 differentiella variabler. Dynamisk optimering användes för att approximera
det dynamiska beteendet hos processen vid skiftningar mellan olika driftpunkter. Den registrerade data
utnyttjades sedan för att identifiera en reducerad modell som fångade det transienta beteendet hos rele-
vanta processvariabler. Den reducerade modellen bestod av 525 ekvationer och 67 differentiella variabler.
Den identifierade modellen visade på god matchning mellan relevanta processvariabler i de simulerade
övergångarna och den reducerade modellen.
Den reducerade modellen användes för att optimera schemaläggningen av luftsepareringsprocessen så
att elkostnaden över en tredagars period minimerades. De optimala resultaten visade på en minskning av
kostnaden på 2.6 % jämfört med en konstant produktionstakt. Schemat implementerades och simulerades
i den fullt dynamiska modellen över de första 24 timmarna för att jämföra relevanta processvariabler med
den reducerade modellen. Resultaten visade på god matchning mellan de båda modellerna.
Detta examensarbete visar att en exakt reducerade modell kan användas för att snabbt hitta ett
optimalt schema över ett större processystem. Detta genom att kraftigt minska systemets storlek utan
att offra noggrannhet av det dynamiska beteendet.
Abstract
A novel framework for making plant scheduling decisions while considering the plant process dynamics
is presented in this thesis. A model of an air separation unit built to supply nitrogen gas and subject
to time-varying electricity prices was used to illustrate this framework. The model includes a cryogenic
distillation column with an integrated reboiler/condenser, a multi-stream heat exchanger, a compressor,
two turbines, and a liquefier. It consisted of 5079 equations and 437 differential variables. The dynamic
behavior of the process during operating point transitions was determined using dynamic optimization.
This data were used to establish a reduced order dynamic model of the system which captures the
transient behavior of relevant process variables. The reduced order model consisted of 525 equations and
67 differential variables. The identified model showed a good fit between the relevant process variables
in the simulated transitions and the reduced order model.
The air separation unit process schedule was optimized using the reduced order model to minimize
electricity cost over a three day time horizon. The optimal result showed a 2.6 % reduction in electricity
cost compared to a flat production rate. The optimal schedule was implemented and simulated in the
full dynamic model for the first 24 hours to compare the relevant process variables to the reduced model
predictions. The result displayed good match between the reduced model and the full dynamic model.
This thesis shows that an accurate reduced order dynamic model can be used for quickly finding
the optimal schedule of large process systems. This by greatly reducing the size and complexity of the
system without sacrificing accuracy of the dynamic behavior. Furthermore, it also shows the economic
benefits of the integrating scheduling and control to count for the dynamic behavior of the system.
Acknowledgments
I was fortunate enough to conduct my research for this thesis at the University of Texas at Austin in the
renowned McKetta Department of Chemical Engineering. I worked under the supervision and guidance
of Dr. Michael Baldea, who helped me achieve my goals in research and writing. This thesis would not
have been possible without his help. I would also like to extend my thanks and gratitude to his research
group, who were always welcoming and helpful. Richard Pattison and Cara Touretzky were especially
valuable resources, whose knowledge and experience were always readily provided to me. I am proud to
present this thesis, a product of a great team of researchers whose help and expertise must be recognized.
I would like to also acknowledge the strong foundation in chemical engineering that I built while at
KTH. Each of my professors imparted upon me the tools and knowledge I utilized in formulating and
carrying out my thesis work. Dr. Per Alvfors, as my examiner, was an invaluable link to my home
university and always available for counseling or advice. While I conducted my research abroad, it was
made possible by the amazing university that I can now call my alma mater.
Contents
1 Introduction 1
2 Background 3
2.1 Air Separation Unit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1.1 Electricity Pricing Options for ASUs . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1.2 Energy and Cost Reduction in ASUs . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1.3 Dynamic Operation and Scheduling of ASUs Subject to Time-Varying Electricity
Prices . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.1.4 Control Strategies of ASUs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
2.1.5 Dynamic models of ASUs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
A Concepts 50
A.1 Differential Alegbraic Equation System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
A.2 Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
A.2.1 Optimization Problem Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . 51
A.2.2 Dynamic Optimization and Discretization Approaches . . . . . . . . . . . . . . . . 51
A.2.3 Solution Algorithms . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
A.3 Reduced Order Modeling for Scheduling Applications . . . . . . . . . . . . . . . . . . . . . 52
A.3.1 Scheduling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
A.3.2 Reduced Order Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
i
A.3.3 Model Identification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
B Results 55
B.1 Steady state optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
B.2 Outputs From the Scheduling Using the Reduced Order Model . . . . . . . . . . . . . . . 55
ii
List of Figures
5.1 Step changes in production set point considered during the dynamic optimization . . . . . 36
5.2 Data and transfer function models for step change responses. . . . . . . . . . . . . . . . . 37
5.3 Dew pressure data compared to transfer function model and non-linear model . . . . . . . 38
5.4 Schedule for production rate set point . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.5 Hold up in storage unit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
5.6 Expected constraint and variable profiles from the dynamic scheduling using the reduced
order model for 72 hours . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
5.7 Production of the plant using the dynamic and reduced order model for the first 24 hours
of the schedule found using the reduced order model. NRMSE between the reduced order
model and the full dynamic model is also given . . . . . . . . . . . . . . . . . . . . . . . . 41
5.8 Hold up in storage during 24 hours using the dynamic and reduced model . . . . . . . . . 41
5.9 Variables and constraints using the full dynamic model and the reduced order model. . . . 43
B.1 Optimal control variables for different production rates with polynomial approximations . 55
B.2 Expected variable and constraint outputs from the scheduling using the reduced order
model for 72 hours . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
iii
List of Tables
iv
Chapter 1
Introduction
Profitability of a process is dependent on minimizing the cost of raw materials and capital expenditures.
This is achieved through process integration, where material and energy streams are recycled to exploit
the efficient use of raw materials and capital equipment [1]. To further improve profitability, chemical
processes should be able to react to (potentially rapid) changes in market conditions by switching prod-
ucts, product purities, product grades, or production rates. These market conditions can include changes
in product demand and costs of raw materials (e.g., chemicals, utilities) [2].
A plant schedule identifies the order of products to be produced in a plant, including the purity
grades and production rates. The schedule is optimized to ensure that the product demands are met,
while minimizing the inventory cost [3]. On the other hand, process control is used to ensure that the
set points of the control variables (e.g. temperature, pressure, or concentration) are held in the presence
of disturbances (e.g. inlet purities, pressures, or temperatures) to minimize the production of products
that does not meet the specified requirements (off-spec products). Process control often uses a model
to describe the dynamic behavior of variables of interest in the process and ensure that operational
constraints are met. A control law is determined which gives a relationship between the manipulated
variables (e.g. valve position or heat removal rate) and the controlled variables. When a plant transitions
between operating states based on the schedule, the setpoints of the control variables are changed. The
controller adjusts the manipulated variables to track the setpoints and move the system to the new
steady state without violating process constraints [4, 5].
In the conventional chemical plant operating paradigm, scheduling and control are performed sepa-
rated over different time scales, see Figure 1.1. Scheduling decisions are typically made days in advance
and control moves take place over the course of minutes or hours [4]. When a chemical plant is rapidly
changing the product, product quality, or production rate to respond to market fluctuations, scheduling
decisions must be made in a similar time frame as the process dynamics. However, a static plant model,
which relies on a table of transitions, is typically used to determine the optimal operating schedule. It
is assumed that the lower level control layer can then guide these transitions exactly. The result is a
potentially suboptimal schedule because the dynamic behavior is not considered in the scheduling de-
cisions. The dynamic layers of the hierarchy should be considered to find the optimal schedule. This
requires the integration of the scheduling and control layers. One way to do this is to build a detailed
dynamic model of the system and then find the optimal transitions which minimize cost. Unfortunately,
this method is time consuming and requires extensive computational time [2, 3, 6, 7]. To overcome this
problem, a reduced order model can be used to make the optimal scheduling decisions as suggested by
Park et al. [8]. A reduced order model captures the dominant process dynamics accurately with much
fewer variables. Such models give an approximation of the process behavior and greatly reduce the
computational burden [2, 6].
A novel framework for making scheduling decisions using a reduced order dynamic model identified
from historical process data is established in this thesis. Here, an Air Separation Unit (ASU) case study
is considered to illustrate the benefits of this general framework. An ASU exhibits a multi-time scale
dynamic behavior during transitions between operating states and is a major electricity consumer [9,10].
It has been shown that the operating cost savings could be realized if an ASU is operated optimally
and change production rates in response to fluctuations in electricity price [11].In past works [11], the
savings were realized by reducing production rates and utilizing a storage unit to meet demand when
1
prices are high. Nevertheless, the optimal schedule was found using a static model. Without considering
the process dynamics, it is not possible to ascertain that the system is pushed to its transition limits, or,
conversely, that all of the suggested transitions are feasible.
Including accurate process dynamics in the scheduling decisions may improve the profitability of an
ASU. The operational cost is minimized by ensuring that the production rate changes are feasible and
that the most aggressive production rate changes are made. Determining the optimal schedule using a
full dynamic plant model often requires extensive computational power. It is also often too difficult to
solve over the (extended) time horizons of interest. However, if an accurate reduced order model of the
process can be identified with significantly fewer states, then the process dynamics could be accounted
for in the scheduling optimization, without increasing the computational burden.
In the literature, several works have applied reduced order models to optimal scheduling [2, 6, 8].
However, these works have investigated case studies of relatively small systems with few state variables.
An ASU is a much more complex system consisting of several interconnected unit operations. The ASU
model displays highly nonlinear behavior and has many state variables. In this work, a framework is
established for creating accurate data-driven models of large, complex process systems which capture the
dynamic behavior of the process variables of interest. In this framework, the models can be identified
from historical process data or by obtaining a dataset from an accurate dynamic plant model. A model
of an ASU was used to obtain a dataset which captures the dynamic behavior of the plant. The behavior
was described over a wide range of production rate changes such that an accurate reduced order model
could be established to be used for optimal scheduling.
Figure 1.1: The hierarchy of decision making time scales in a chemical plant [4]
Objectives
The goal of this thesis is to build an accurate dynamic model of an ASU with a liquid nitrogen storage
unit. The operational cost should then be minimized by operating the plant dynamically when subject
to time-varying electricity prices. A reduced order model is to be found which accurately represents the
dynamic behavior of process variables of interest. The predicted operational cost and dynamic behavior
of the two models should also be compared. The main objective is to evaluate the novel scheduling
approach that considers a simplified representation of the process dynamics.
This thesis is organized as follows: Chapter 1 and 2 gives a brief introduction and background.
Concepts used are presented in appendix A. This is followed by a detailed description of the mathematical
ASU model in Chapter 3. The full dynamic model of the ASU includes a cryogenic distillation column,
an integrated reboiler/condenser, a Primary Heat Exchanger (PHX), a compressor, a turbine, and a
liquefier. Chapter 4 describes the optimal scheduling methodology. The steady state plant operation
is first optimized at several production levels. Next, dynamic optimization is used to find the optimal
transitions between the different production levels. The optimal transitions are used to mimic historical
process data which in turn, are used to identify a reduced order model of the system that represents
the dynamic behavior of relevant variables in the plant. An optimal schedule is determined by dynamic
optimization using the reduced order model when subject to time-varying electricity prices. Finally, the
optimal schedule is implemented on the full dynamic model to determine the effectiveness of the identified
schedule. The results are presented and discussed in Chapter 5 with the conclusions and future work in
Chapter 6.
2
Chapter 2
Background
In this chapter, a brief overview of the concepts and theories used in the report are given. First, an account
over an air separation process is given together with the most used control and scheduling strategies.
Thereafter, a short summary of theoretical concepts regarding optimization and model identification is
provided.
3
2.1.2 Energy and Cost Reduction in ASUs
Optimizing the design of an ASU aims to lower both the energy use of the plant and the overall cost.
Strategies to reduce the energy usage in the column include using heat pumps to increase thermal
efficiency [19] and structured packing to reduce pressure drop in the distillation column [20]. Aneke and
Wang [21] suggested using a Rankine cycle to generate electricity from the excess heat of the compressors.
Their results showed a possible 11 % reduction in the overall electricity requirement. Rizk et al. [22]
studied three different distillation column designs and compare their exergy losses. The results suggested
that the plant efficiency can vary significantly depending on the column design.
Research regarding the optimization of the design of ASUs have often utilized linear models [17, 19,
21–23]. However, the equations describing the ASU’s behavior are actually highly nonlinear. Detailed
nonlinear models should be used for accurate optimal design calculations. Zhu et al. [17] considered the
optimal design of ASUs under uncertainty with a detailed nonlinear model, and their results showed a
reduction in the capital cost of the plant.
The inlet air compressor is the main electricity consumer of an ASU. Improvements in the compres-
sor and turbine manufacturing process using computational fluid dynamics in combination with three
dimensional machining to find the optimal form of the different parts has increased the efficiency [24].
Turbines are used in an ASU to utilize pressure drops to generate electricity. Efficiency improvements
include the use of active magnetic bearings so that the frictional loss is reduced [25]. Multi-stream heat
exchangers are used to further reduce the energy consumption of the plant and allow for multi-stream
use [26]. Computer simulations further improve the design of the process by utilizing and integrating
the existing units [24].
ASUs are commonly integrated with their supplier and/or customers to reduce the energy consump-
tion and the capital cost [24]. Integrated gasification combined cycle, coal-based iron making, and
combined power or industrial gas production facilities can all benefit from integrating their usage of
oxygen or nitrogen with an ASU. The overall efficiency, power output, and cost saving can improve
compared to a non integrated process [27]. Elevated pressure (above 6.5 bar) ASUs could also be used,
which could lead to the usage of smaller and more compact cryogenic components and the reduction of
capital cost [13].
4
that would minimize the operational cost. Their results showed that air separation plants with a great
flexibility in the production capacity benefit the most. However, time-varying electricity pricing has an
appeal even with a low degree of flexibility. Mitra et al. [31] investigated production scheduling under
time-varying demands and prices using a mixed-integer linear programming approach. They used a
deterministic surrogate model for the process based on the production mode, transition logic, and mass
balances. They defined each feasible production region by a convex region of already known feasible
operating points. Transition logics were used to handle the transition between regions. Their result
showed a 10 % improvement of the cost savings compared to a heuristic model. Storage capacity is also
shown to be an important factor in reducing operational cost. It is worth noting that most research to
date on ASU operation subjected to time-varying electricity prices has used simplified static linear plant
models to describe the system. These kinds of models do not capture the dynamic behavior of the plant
during production rate transitions, which remains an open research problem.
5
N2 Gas Product
Integrated Reboiler/Condenser
Evaporator
Waste
Auxiliary Heat Exchanger
Storage
tank
Air Feed Liquefier Drain
Compressor
Storage Unit
Turbine 2
Expansion Valve
Column
Turbine 1
6
Chapter 3
As described in the introduction, an ASU consists of the interconnection of several unit operations. Each
unit model is described by a system of Differential Algebraic Equations (DAE). In this work, gPROMS
4.0.0 [37] was used to construct the full dynamic model and formulate (and solve) the optimization
problems. In this chapter, an overview of the process flowsheet is described followed by a detailed
description of the model equations.
7
3.3 Model equations
In this section the equations describing the ASU model are presented. First, the dynamic models of
the column and the integrated reboiler/condenser are presented followed by the PHX model and the
compressor and turbine models. Lastly, the the liquefier and storage unit models are presented.
3.3.1 Column
Several authors have developed ASU column models [15,16,38]. In this work, the model has been derived
based on that developed by Huang et al. [14] with the following assumptions:
• Constant pressure drop on each stage
• Ideal vapor phases
• Well-mixed streams entering each stage
• Vapor liquid equilibrium is established on each tray
For every stage a DAE system was derived which included differential material and energy balances as
well as algebraic equations describing the equilibrium, hydraulics, and physical properties. An overview
of the column is shown in Figure 3.1.
Lcondenser , xcondenser,j
V1 , y1,j
T1 , P1 , M1
i=1
V2 , y2,j L1 , x1,j
Ti , Pi , Mi i
F30 , xf30,j i = 30
T30 , P30 , M30
L30 , x30,j
8
where Mi is the molar hold up on stage i [mol], Li is liquid flow from stage i [mol s−1 ], Vi is vapor flow
from stage i [mol s−1 ] and, Fi is the feed flow at tray i [mol s−1 ] (which is nonzero only at the feed stage).
dxi,j (Li−1 (xi−1,j − xi,j ) + Vi+1 (yi+1,j − xi,j ) − Vi (yi,j − xi,j ) + Fi (xfi,j − xi,j )
= (3.2)
dt Mi
where xi,j is the liquid mole fraction of component j at stage i, yi,j is the vapor mole fraction of com-
ponent j at stage i and xfi,j is the liquid mole fraction of component j in the feed.
Energy balance
f
dhL
i (Li−1 (hL L V L V L L
i−1 − hi ) + Vi+1 (hi+1 − hi ) − Vi (hi − hi ) + Fi (hi − hi )
= (3.3)
dt Mi
−1
where hLi is the molar enthalpy of the liquid at stage i [kJ mol ], hVi is the molar enthalpy of the vapor
f
at stage i [kJ mol ] and hi is the molar enthalpy of the feed stream [kJ mol−1 ]. The liquid and vapor
−1
where aL L V V
j , bj , aj , and bj are constants for liquid and vapor specific to each component j and Ti is the
temperature at stage i [K]. aL L V V
j , bj , aj , and bj were calculated by data regressed from ASPEN [39]
Soave-Redlich-Kwong physical property method.
Summation equation X
1= yi,j (3.6)
j
Hydraulic equation
Li = kd Mi (3.7)
where kd is a tuning constant set to 0.5 [14].
Vapor-liquid equilibrium
yi,j pi = γi,j xi,j psat
i,j (3.8)
where pi is the total pressure on stage i [bar], psat
i,j is the saturation pressure of component j [bar] and
γi,j is the activity coefficient for component j. psat
i,j was calculated using Antoine’s equation:
βj
log psat
i,j = αj − (3.9)
δj + Ti
where αj , βj and δj are the Antoine’s coefficients for each component j [40] and Ti is the temperature at
each stage [◦C]. γi,j is the activity coefficient describing the nonideal behavior of the liquid for component
j. It was calculated using Margules equation:
AN2 ,O2 x2i,O2 + AN2 ,Ar x2i,Ar + (AN2 ,O2 + AN2 ,Ar − AO2 ,Ar )xi,O2 xi,Ar
γi,N2 = exp (3.10a)
RTi
AN2 ,O2 x2i,N2 + AO2 ,Ar x2i,Ar + (AN2 ,O2 + AO2 ,Ar − AN2 ,Ar )xi,N2 xi,Ar
γi,O2 = exp (3.10b)
RTi
AN2 ,Ar x2i,N2 + AO2 ,Ar x2i,O2 + (AN2 ,Ar + AO2 ,Ar − AN2 ,O2 )xi,N2 xi,O2
γi,Ar = exp (3.10c)
RTi
9
where Aj,k are the Margules constants for components j, k (j 6= k) [41] and R is the gas constant
[J mol−1 K−1 ].
Equations (3.1)-(3.10) described the dynamic model of a single stage i. However, the system was
an index 2 DAE system because the algebraic variable Vi could not be explicitly calculated from the
equations above.
The following procedure was used to reduce the index (notation is the same as presented by Huang
et.al [14]). As seen in equation (3.4) and (3.5), the enthalpy was a function of both temperature and
composition. Differentiating the left hand side of the energy balance, equation (3.3), gave [14]:
f
L X ∂hL Li−1 (hL L V L V L L
i−1 − hi ) + Vi+1 (hi+1 − hi ) − Vi (hi − hi ) + Fi (hi − hi )
∂hi T̄i + i
x̄i,j = (3.11)
∂Ti j
xi,j Mi
∂hL
i
X
= xi,j aL
j (3.12a)
∂Ti j
∂hLi
= aL L
j Ti + bj (3.12b)
xi,j
An expression for T̄i could be found through derivation of the summation equation, equation (3.6), with
respect to time [14]:
X dyi,j X dKi,j dxi,j
0= = xi,j + Ki,j (3.13)
j
dt j
dt dt
γi,j psat
i,j
where Ki,j = pi is a function of temperature and composition. After applying the chain rule to
(3.13) and rearranging, an equation for dT dt could now be established [14]:
i
P h P ∂Ki,j i
dTi j x i,j x̄
k ∂xi,k i,k + K x̄
i,j i,j
T̄i = =− ∂Ki,j
(3.14)
dt
P
j xi,j ∂Ti
∂Ki,j ∂Ki,j
With ∂xi,k and ∂Ti being explicitly calculated as:
∂Ki,j psat
i,j ∂γi,j
= (3.15a)
∂xi,k pi ∂xi,k
∂Ki,j psat
i,j ∂γi,j γi,j ∂psat
i,j
= + (3.15b)
∂Ti pi ∂Ti pi ∂Ti
Vapor hold-up
Vapor holdup is usually assumed to be negligible [14] but to capture the actual process dynamics in
the model, the holdup was considered in this case. The vapor flow into every tray was calculated using
an artificial first order dynamic expression:
dVi 1
= (Vi+1 − Vi ) (3.16)
dt τ
where τ is a time constant equal to 30 s
Equation (3.1), (3.2), (3.4)-(3.12) and (3.14)-(3.16) were used to describe each stage i and component
j in the column model. 30 equilibrium stages were assumed in the column and they are numbered in a
10
increasing order starting from the top stage (1). The total pressure drop in the column was estimated
to 0.2 bar. The feed entered at the bottom of the column (stage 30). A nomenclature of the indices, the
variables, and the parameters used in the column model are given in Table 3.1
Table 3.1: Indices, variables and parameters used in the column model
Indices
i Tray number
j Component, j=1,2,3
k Component, k=1,2,3
L Liquid
V Vapor
f Feed
Variables
M Molar hold up
L Liquid flow
V Vapor flow
F Feed flow
x Molar liquid fraction
y Molar vapor fraction
γ Activity coefficient
h Molar enthalpy
p Pressure
psat Saturation pressure
T Temperature
dT
T̄ dt
dx
x̄ dt
Parameters Value Reference
a Enthalpy constant See reference [39]
b Enthalpy constant See reference [39]
kd Tuning constant 0.5 [14]
Aj, k Margulesconstant See reference [41]
α Antoine’s constant See reference [40]
β Antoine’s constant See reference [40]
δ Antoine’s constant See reference [40]
τ Time Constant 30 Estimated
R Gas Constant 8.314 [42]
Condenser
The condenser model assumed zero material and energy holdup and that saturated liquid leaves the
condenser.
Mass Balance
0 = V1 rgasdraw − Vcondenser (3.17)
−1
where V1 [mol s ] is the top stage vapor rate, rgasdraw is the fraction of the top stage vapor flow that
is sent to the condenser and Vcondenser is the vapor molar flow to the condenser [mol s−1 ] which was
equivalent to the liquid molar flow leaving the condenser, Lcondenser
11
Vreboiler , yreboiler,j
Reboiler
Treboiler , Preboiler , Mreboiler
L30 , x30,j , preboiler
Qcondensor , UA
Condenser
Tcondenser , Pcondenser
Vcondenser , y1,j
Lcondenser , xcondenser,j
rgasdraw=Vcondenser/V1
Lreboiler , xreboiler,j Drain
Fproducts
V1 , y1,j L30 , x30,j , p30
Products, to PHX Waste, to PHX From top tray To top tray From bottom tray
Energy Balance
0 = V1 rgasdraw hV1 − Lcondenser hL
condenser − Qcondenser (3.18)
−1
where hV1is the top tray molar vapor enthalpy [kJ mol ], hL
is the liquid molar enthalpy in the
condenser
condenser [kJ mol−1 ] and Qcondenser is the heat extracted from the condensation of vapor [kW]. The
liquid enthalpy was calculated using equation (3.4) and Qcondenser was calculated as:
Because all of the inlet vapor is condensed, the composition of the outlet liquid is the same as the
inlet vapor:
y1,j = xcondenser,j (3.20)
where y1,j is the top stage composition of component j and xcondenser,j is the liquid molar composition
in the condenser of component j.
Furthermore, because the liquid is assumed to be saturated, the condenser pressure can be calculated
using the saturation pressure of each component and the liquid composition:
X
pcondenser = xcondenser,j psat
j (3.21)
j
Reboiler
The reboiler is modeled as an additional equilibrium stage with a heat input in the energy balance. The
expansion valve is modeled as a static adiabatic expansion, and was included in the reboiler section for
12
simplicity.
Mass Balance
dMreboiler
= L30 − Lreboiler − Vreboiler (3.22)
dt
where L30 is the liquid flow from the bottom tray in the column [mol s−1 ], Lreboiler is the drain flow
from the reboiler [mol s−1 ] and, Vreboiler is the waste vapor flow from the reboiler [mol s−1 ]. To keep the
molar hold up at a set point, a proportional controller with saturation was implemented on the drain
flow:
(
Lsp,reboiler − Kc (Msp,reboiler − Mreboiler ) if Mreboiler ≥ Msp,reboiler
Lreboiler = (3.23)
0 if Mreboiler < Msp,reboiler
Energy Balance
L L
∂hreb T̄reb +
X ∂hreb L30 (hL L V L
30 − hreb ) − Vreb (hreb − hreb ) + Qreb
x̄reb,j = (3.25)
∂Treb j
xreb,j Mreb
where Qreb is the heat transferred from the condensor to the reboiler [kW] and is equal to Qcondenser ,
−1
hLreb is the liquid molar enthalpy in the reboiler [kJ mol ] calculated using equation (3.4) and hVreb is
−1
the vapor molar enthalpy in the reboiler [kJ mol ] calculated using equation (3.5), and hL 30 is the molar
enthalpy at the bottom of the column [kJ mol−1 ]. T̄reb and x̄reb,j were defined in an equivalent way as
for the column model, equation (3.11).
For the vapor liquid equilibrium, equation (3.8) was used in a similar way as for a tray to calculate
the composition of the waste stream. The pressure in the reboiler , preboiler was held constant at 2.5
bar. The indices, the variables and the parameters used in the integrated reboiler/condenser model are
given in Table 3.2.
13
Table 3.2: Indices, variables and parameters used in the integrated reboiler/condenser model
Indices
i Tray number
j Component, j=1,2,3
L Liquid
V Vapor
1 Top tray in column
30 Bottom tray in column
sp Set point
reb reboiler
Variables
M Molar hold up
L Liquid flow
V Vapor flow
x Molar liquid fraction
y Molar vapor fraction
γ Activity coefficient
h Molar enthalpy
p Pressure
psat Saturation pressure
T Temperature
UA Product of overall heat transfer coefficient and heat
transfer area
rgasdraw Fraction of product entering the condenser
Q Heat flow
dT
T̄ dt
dx
x̄ dt
Parameters Value Reference
Kc Controller gain 10−4 Estimated
preboiler Reboiler Pressure 2.5 [bar] Estimated
14
Turbine
High
pressure
Low product
Pressure
product
Zone 1 Zone 2
Zone 1
Zone 1 consisted of differential equations describing the wall and stream temperature in every segment.
In each segment, the streams were assumed to be well mixed and the wall temperature uniform. Figure
3.4 shows an overview of a segment in zone 1.
15
Segment n
where mk,n is the metal wall mass in zone k and segment n [kg], Cpmetal is the specific heat capacity
of the metal wall [J kg−1 K−1 ], Tk,j,n is the temperature in zone k and segment n of stream j [K], Tk,nm
is the metal wall temperature in zone k and segment n, and U Ak,j,n is the product of the overall heat
transfer coefficient and the heat transfer area in zone k and segment n for stream j [J K−1 ]. The streams
were numbered in increasing order as: the high pressure product stream, low pressure product stream,
the waste stream, and the feed air stream. The metal wall mass was calculated as:
rk mmetal
mk,n = (3.27)
Nsegments
where rk is the fraction of the PHX that is in zone 1, Nsegments is the number of segments in zone 1 and
mmetal is the total metal mass in the PHX [kg] calculated as:
NS
!
X
mmetal = ρmetal W LH − Vj (3.28)
r
where ρmetal is the metal density [kg m−3 ], W, L and H are the width, length and height of the PHX [m]
(see Figure 3.5), and Vj is the total volume occupied by stream j in the PHX [m3 ]. Vj was estimated
from the volume occupied by stream j in each slot:
1 1
Vjslot = bj − bj tj + − tj tj (L − 2tedge ) (3.29)
Nj Nj
16
H
where Vjslot is the volume occupied by stream j in each slot [m3 ], N1j is the width of a single slot for
stream j [m], bj is the distance between the parting sheets for stream j [m], tj is the fin thickness for
stream j [m] and tedge is the thickness of the PHX wall [m] (see Figure 3.6 and 3.7).
The volume occupied by stream j in each layer was then calculated as:
Finally, the total volume occupied by each stream in the PHX could then be estimated as:
Vj = Vjlayer nj (3.31)
where nj is the number of layers for each stream.
Energy Balance
The energy balance in each segment of zone 1 was:
dE1,j,n
= F1,j (hV1,j,n+1 − hV1,j,n ) + U A1,j,n (T1,n
m
− T1,j,n ) (3.32)
dt
where E1,j,n is the energy hold up in stream j in segment n and zone 1 [kJ], Fk,j is the molar flow of
stream j in zone 1 [mol s−1 ], and hV1,j,n is the vapor enthalpy of stream j in segment n [kJ mol−1 ]. The
enthalpy of the streams in each segment was calculated using equation (3.5). E1,j,n was defined as:
The molar density of stream j in zone 1 and segment n, ρV1,j,n [mol m−3 ], in equation (3.33) was calculated
from the ideal gas law:
p1,j,n
ρV1,j,n = (3.35)
RT1,j,n
17
Figure 3.6: Front view of the PHX
where p1,j,n is the pressure of stream j in segment n and zone 1 [Pa], T1,j,n is the temperature of stream
j in zone 1 and segment n [K], and R is the gas constant [J mol−1 K−1 ]. Zone 1 was divided into 50
segments.
Zone 2
The air feed stream, stream 4, was divided at the end of zone 1. One fraction was extracted as vapor
and the other continued into zone 2 and phased changed into liquid air, see Figure 3.8. To simplify the
calculations only one segment was considered in zone 2 i.e Nj,segment = 1. The temperature for each
stream in zone 2 was calculated as an average of the inlet and outlet temperature of that stream:
18
Segment 1
kwithdrawn=F2,4/F1,4
Fvapor
Zone 1 Zone 2
Tj,in + Tj,out
T2,j = (3.36)
2
where Tj,in is the inlet temperature of stream j and Tj,out is the outlet temperature of stream j.
The change in metal wall temperature was captured by the same equation as in zone 1, equation
(3.26). The energy balance was formulated equivalent to that of zone 1 for streams 1-3:
dE2,j
= F2,j (hV2,j,in − hV2,j,out ) + U A2,j (T2m − T2,j ) (3.37)
dt
where hV2,j,in is the vapor enthalpy of stream j at the inlet of zone 2 [kJ mol−1 ] and hV2,j,out is the vapor
enthalpy of stream j at the outlet of zone 2 [kJ mol−1 ]. The inlet and outlet enthalpies were calculated
using equation (3.5) and the respective inlet and outlet temperature.
A fraction of stream 4 changed phase from gaseous air into liquid air and the energy balance was
calculated as:
dE2,4
= F2,4 (hV2,4,in − hL m
2,4,out ) + U A2,4 (T2 − T2,4 ) (3.38)
dt
where hliq
2,3,out is the liquid enthalpy of stream 4 at the outlet of zone 2 [kJ mol
−1
] calculated using
equation (3.4) and the outlet temperature of that stream. F2,4 is the liquid air feed stream [mol s−1 ]
calculated as:
19
E2,4 = hL L
2,4 V2,4 ρ2,4 (3.40)
−1
where hL2,4 is the liquid enthalpy for stream 4 in zone 2 [kJ mol ] calculated using equation (3.4) and
the average temperature, T2,4 . V2,4 is the volume occupied by stream 4 in zone 2 [m3 ] calculated in the
−3
same way as in zone 1, equation (3.34). ρL 2,4 is the liquid molar density of stream 4 in zone 2 [mol m ]
and was calculated using the Rackett equation [38, 44]:
" 2/7 #
RTic
1 T2,4
ln = ln c + 1+ 1+ c ln(ZiRa ) (3.41)
ρLi p i T i
NC
X
ρL
2,4 = xi ρL
i (3.42)
i=1
−3
where ρLi is the liquid molar density of component i in stream 4 [mol m ], Tic is the critical temperature
for component i in stream 4 [K] [45], pi is the critical pressure for component i in stream 4 [Pa] [45], ZiRa
c
is the Rackett compressibility factor of component i in stream 4 [46], and xi is the liquid molar fraction
of component i in stream 4.
mk,n and ρV2,j was calculated in the same way as in zone 1 but with only 1 segment (equation (3.27)
and (3.35), respectively). Variables, indices and parameters used in the PHX model are given in Table
3.3.
20
Table 3.3: Indices, variables and parameters used in the PHX model
Indices
k Zone
n Segment
j Stream
i Component
L Liquid
V Vapor
in Inlet
out Outlet
Variables
Tm Metal wall temperature
T Stream temperature
mk,n Metal wall mass in zone k, segment n
mmetal Total metal mass
Vj Volume occupied by stream j
Vslot
j Volume occupied by stream j in each slot
Vlayer
j Volume occupied by stream j in each layer
E Energy hold up
F Molar flow
h Molar enthalpy
ρ Molar density
p Pressure
psat Saturation pressure
kwithdrawn Fraction liquid air entering zone 2
x Molar fraction
T Temperature
Parameters Value Reference
Cp Specific heat capacity of metal wall 892 [J kg−1 K] Estimated
UA1,j,n Product of overall heat transfer coefficient and heat 40000 [J K−1 ] Estimated
transfer area in zone 1
UA2,j,n Product of overall heat transfer coefficient and heat 10000 [J K−1 ] Estimated
transfer area in zone 2
rk Fraction of PHX that is in zone 1 0.954817 Estimated
Nk,segments Number of segments in zone 1 50 Estimated
ρmetal Metal density [kg m−3 ] 2731 Estimated
W Width of PHX [m] 1.5 Estimated
L Length of PHX [m] 8.2 Estimated
H Height of PHX [m] 3 Estimated
Nj Number of fins of stream j [70 70 70 100] Estimated
bj Distance between parting sheets for stream j [m] 6.75e-3 Estimated
tj Fin thickness for stream j [m] 0.4e-3 Estimated
tedge PHX wall thickness [m] 1e-2 Estimated
nj Number of layers for each stream 59 Estimated
R Ideal gas constant 8.314 [42]
Tci Critical temperature of component i See reference [45]
pci Critical pressure of component i See reference [45]
ZRa
i Rackett compressibility factor of component i See reference [46]
21
GHpoly
W = (3.43)
ηpoly
where W is the power demand [W], G is the mass flow [kg s−1 ], Hpoly is the polytropic head [J kg−1 ] and
ηpoly is the polytropic efficiency. Hpoly was calculated as:
" (n−1)/n #
n pout
Hpoly = ZRTin −1 (3.44)
n−1 pin
where n is the polytropic exponent, Z is the compressibility factor, R is the gas constant [J mol−1 K−1 ],
Tin is the temperature of the inlet stream [K], and pout and pin are the pressure exiting and entering the
compressor [bar], respectively. ηpoly in equation (3.43) was calculated as:
(k − 1)/k
ηpoly = (3.45)
(n − 1)/n
where k is defined as the ratio of specific heats:
Cp
k= (3.46)
Cv
Values for Cp and Cv was found in CRC handbook of chemistry and physics [42] for oxygen and nitrogen
and are assumed to be independent of temperature.
When a fluid is compressed the relationship between pressure, p, and Volume, V , can be described
using the polytropic exponent:
pV n = Constant (3.47)
For a polytropic process the exponent is constant and equation (3.47) can be rearranged as follows:
ln p + n ln V = Constant (3.48)
An empirical relationship can be used if no data are available. Perry and Green [44] suggest a relationship
between the adiabatic efficiency and the polytropic exponent:
pout (k−1)/k
pin −1
ηad = pout (n−1)/n
(3.49)
pin −1
where ηad is the adiabatic efficiency. It was approximated to 80 % in this study.
Turbine
The turbine model was based on the same equations as for the compressor and the model was used for both
turbines in the process flowsheet. It was assumed that the polytropic exponent could be approximated
in the same way as in equation (3.49) and that the adiabatic efficiency was the same for the turbine as
for the compressor. The power supply from the turbine was then calculated as:
22
Table 3.4: Indices, variables and parameters used in the compressor and turbine models
Indices
poly Polytropic
in Inlet
out Outlet
ad Adiabatic
Variables
W Power
G Mass flow
H Polytropic head
η Efficiency
p Pressure
T Temperature
k Ratio of specific heats
n Polytropic exponent
Parameters Value Reference
Z Compressibility factor 1 Estimated
R Ideal gas constant 8.314 [42]
Cp,air Specific heat capacity for air [kJ kg−1 K] 1.007 [42]
Cp,N2 Specific heat capacity for nitrogen [kJ kg−1 K] 1.123 [42]
Cv,air Specific heat capacity for air [kJ kg−1 K] 0.7181 [42]
Cv,N2 Specific heat capacity for nitrogen [kJ kg−1 K] 0.7710 [42]
ηad Adiabatic efficiency [%] 80 Estimated
3.3.5 Storage
The storage unit consisted of a liquefier, a storage tank, and an evaporator. The liquefier was approx-
imated by a steady state model as the dynamics are typically faster than the ASU. Furthermore, the
evaporator was assumed to work with an external water cooled heat exchanger and was not included as
a model, see Figure 3.9.
Liquefier
The liquefier was modeled as an ideal cycle with an isothermal compressor and an isentropic turbine at
steady state [47], see Figure 3.9. The energy balance was derived as:
Wc + F h1 = WT + Qr + F hf (3.51)
where Wc is the work required by the compressor [W], WT is the work done by the turbine [W], Qr is
the heat produced by the compression [W], F is the molar flow to the liquefier [mol s−1 ], h is the molar
enthalpy [J mol−1 ]. Rearranging equation (3.51) gave the total energy consumption of the liquefier:
Qr = T1 F (s1 − sf ) (3.54)
−1 −1 −1 −1
where s1 is the entropy at the inlet [J mol K ], sf is the entropy at the outlet [J mol K ], and T1
is the temperature at the inlet [K].
23
Gaseous N2 in
F, h1 , T1 , s1
Gaseous N2 out
Evaporator
Compressor
Turbine Fout
F, hf , sf
Liquid N2 Storage tank
Liquefier
Combining equation (3.51) and (3.53) gave an equation for the overall work required by the liquefier:
where Fproduced is the molar flow of nitrogen produced from the ASU [mol s−1 ] and Fdemand is the product
demand [mol s−1 ]. The demand was assumed to be constant at 20 mol s−1 corresponding to about 50
t day−1 .
The flow of nitrogen produced changed in response to the variability of the electricity prices. To
ensure a constant flow of nitrogen product out from the system, the extra nitrogen produced at low
prices was sent to the liquefier and stored as liquid product in the storage tank. When prices are high,
stored liquid nitrogen was evaporated to satisfy the demand.
24
Storage Tank
The liquefied nitrogen was stored in a cryogenic storage tank at the same pressure as the inlet pressure
to the liquefier. The tank was assumed to be well insulated and no heat was neither added nor lost from
the tank. It was modeled as an accumulation term:
dMstorage
= F − Fout (3.58)
dt
where Mstorage is the molar holdup in the tank [mol] and Fout is the molar flow out from the tank
[mol s−1 ]. The liquid flow out from the tank was evaporated in an external heat exchanger to the same
temperature as the product flow from the primary heat exchanger.
The indices, variables, and parameters used in the storage tank model are presented in Table 3.5
Table 3.5: Indices, variables and parameters used in the storage model
Indices
c Compressor
T Turbine
f Outlet
1 Inlet
net Net work
Variables
W Power
Qr Heat from compression
h Enthalpy
F Molar flow
s Entropy
k Ratio of specific heats
n Polytropic exponent
Mstorage Hold up in storage
Parameters Value Reference
η Efficiency [%] 40 Estimated
Fdemand Molar demand [mol s−1 ] 20 Estimated
25
Chapter 4
The framework includes several steps to establish the optimal schedule, see Figure 4.1. Since no historical
dataset was available for this process, the dynamic model established in Chapter 3 was used to simulate
the historical process data. First, the optimal steady state process operation was determined at several
production levels (from +20% of the nominal operation to −20% of the nominal operation). Then the
behavior of process variables during transitions between each production level were determined using
dynamic optimization. The reduced order model was then identified from the transition data. The
reduced order model was then used to find the optimal schedule of the production rate and liquid
inventory management. Finally, the optimal schedule was validated using the full dynamic model.
Training data
Model identification
Scheduling
Dynamic optimization
Verification
26
production level. Several constraints were imposed to ensure feasible operation at each production levels.
The solution to the optimization problem was found using gPROMS 4.0 [37] and the NLPSQP solver.
where pdew is the dew point pressure at the outlet of zone 1 [bar] and p1,4,50 is the pressure of the air
feed stream at the outlet of zone 1 [bar].
Additionally, a constraint was set to ensure that the outlet of the air feed stream from zone 2 was in
the liquid phase:
p2,4,1 − pbubble ≥ 0 (4.3)
where p2,4,1 is the pressure at the outlet of zone 2 [bar] and pbubble is the corresponding bubble point
pressure at the outlet of zone 2 [bar].
The dew point pressure was calculated as:
X
pdew = pdew,j xj (4.4)
j
where pdew is the dew point pressure of the mixture [bar], pdew,j is the dew point pressure of component
j [bar] calculated using equation (3.9), and xj is the liquid molar fraction of component j in the mixture.
27
Likewise, the bubble point pressure was calculated as:
X
pbubble = pbubble,j xj (4.5)
j
where pbubble is the bubble point pressure of the mixture [bar], pbubble,j is the bubble point pressure of
component j [bar] calculated using equation (3.9), and xj is the liquid molar fraction of component j in
the mixture.
The nitrogen product stream was at high pressure, and to fully utilize the available heat, the stream
was withdrawn from the heat exchanger and expanded in a turbine to produce electricity, see Figure 3.3.
To ensure that the nitrogen exiting the turbine was in the gas phase, the pressure was constrained to be
below the corresponding saturation pressure for nitrogen:
psat
N2 − pout ≥ 0 (4.6)
where psat
N2 is the saturation pressure for nitrogen [bar] calculated using equation (3.9) and pout is the
pressure exiting the turbine [bar].
Column Constraints
The flooding capacity constraint had to be met in the distillation column. At 100 % flooding, liquid is
excessively carried over to the adjacent stage due to a high vapor flow rate. This typically results in
drastically reduced stage efficiencies and high pressure drops. The flooding fraction is typically set at
80-90 % during nominal production [48].
The flooding constraint was adapted from the work by Coulson and Richardson [48] and Cao [38].
To ensure that no flooding occurred, the flooding fraction was constrained to be below 1:
ui
αif looding = <1 (4.7)
uf looding,i
where αif looding is the fractional flooding at stage i, ui is the vapor velocity at stage i [m s−1 ], and
uf looding,i is the flooding velocity at stage i [m s−1 ]. ui is calculated by:
Vi
ui = (4.8)
ρV AActive
where Vi is the molar vapor flow at stage i [mol s−1 ], ρV is the vapor molar density of the mixture at
stage i [mol m−3 ] calculated using equation (3.35), and AActive is the active area of stage i [m2 ].
The flooding velocity can be calculated as [38]:
s
ρL ML − ρV MV
uf looding,n = K1 (4.9)
ρV MV
where K1 is a capacity constant and ρL is the liquid density [mol m−3 ] obtained from equation (3.42).
ML is the molar weight of the liquid [kg mol−1 ] and MV is the molar weight of the vapor [kg mol−1 ].
K1 was calculated as: σn
K1 = K2 −
(4.10)
20 ∗ 10 3
where σn is the surface tension of the liquid mixture [N m−1 ] and K2 is a capacity factor calculated as:
where Ts is the tray spacing [m] and F is a flow parameter. The tray spacing was estimated from work
done by Thorogood et al. [49] and F was calculated as:
s
GL ρV MV
F = Vi (4.12)
Gi ρL ML
28
−1
where GL i is the mass flow rate of liquid at stage i [kg s ] and GVi is the mass flow rate of vapor at stage
−1
i [kg s ]. F was assigned to 0.1 if it was calculated to be lower than 0.1 [38].
The surface tension of the liquid mixture was estimated from the surface tension of the pure compo-
nents:
X 3
r
σm = xj σjr (4.13)
j
where σm is the surface tension of the mixture, σj is the surface tension of the pure component j
[mN m−1 ] and r is an adjustable parameter set to 0.25 [50]. The surface tension of the pure component
was calculated using an empirical relationship given by Kai et al. [50]:
Ti p
σj = σ0 (1 − ) (4.14)
Tc
where σ0 and p are parameters for each component j, Ti is the temperature at stage i, and Tc is the
critical temperature of component j [50].
Constraints
The constraints of the steady state optimization must hold during the transitions as well (equation (4.2)
- (4.7), (4.15) - (4.17)). The maximum allowable impurity during transitions was set to 1000 ppm. To
ensure that these constraints are satisfied throughout the time horizon, they were transformed into path
constraints. An example of this transformation for the dew point pressure constraint, equation (4.2), is
given by:
29
Z tf
max(0, p1,4,50 − pdew ) ≤ (4.18)
t0
where is a small positive parameter. These path constraints are implemented in gPROMS as endpoint
inequality constraints.
Discretization
To simulate the control moves, a piecewise linear control profile is used, see Figure A.1. To obtain the
piecewise linear control profile, the rate of change for each manipulated variable in each control interval
was determined:
dui,j
= ai,j (4.19)
dt
where ui corresponds to manipulated variable i (see Table 4.2) and ai,j is the rate of change of variable i
in control interval j. Thus, the optimization decision variables were actually piecewise constants, which
correspond to the slope of the manipulated variables in each control interval.
five control intervals were established for each transition with an extra interval at the end to ensure
that steady state is reached (fix ai,j = 0 in the last interval). An additional constraint was imposed to
ensure that the optimal steady state values of the manipulated variables were reached at the end of the
time horizon [38]:
ui 2
(1 − ) ≤ (4.20)
uop
i,j
where uop
i are the optimal values of manipulated variable i at production rate j and is a very small
parameter.
Objective function
The objective of the dynamic optimization was to minimize the transition time between production
levels. The objective function was formulated as [38]:
" #2
Z tf
Fproduction (t)
φDO = 1− ∗ dt + t2f (4.21)
t0 Fproduction
∗
where Fproduction is the production rate at the end of the transition.
30
Relevant Process Variables
The desired production rate was the scheduling decision, and was therefore the input to the reduced
order model. The outputs were the process variables relevant to scheduling decisions. These include the
actual production rate and energy use, as well as the constrained operating variables. The constrained
variables must be included to ensure that the production rate changes are physically realizable. The
energy use was directly proportional to production rate, so only one model must be identified for these
two variables. The constrained operating variables included the product impurity, column flooding, the
phase of the inlet stream exiting zone 1 of the PHX, the phase of the inlet stream exiting zone 2 of the
PHX, the product of the overall heat transfer coefficient and area in the reboiler, and the reboiler liquid
level. The output variables and their corresponding identified model order are presented in Table 4.3.
Table 4.3: Outputs in the scale bridging model and model order
Due to the fact that each individual transfer function for each output j, Gi,j , was second order, the
overall transfer function Gj was 14th order for each output j.
The parameters for the transfer functions were found using the System Identification Tool box in
MATLAB [51]. It was determined that the 14th order model had poor conditioning (i.e., the ratio of
31
G1,j
+
+
u1
ysp G2,j G5,j + y
+
u2
G3,j G6,j
u3
G4,j G7,j
the largest and smallest eigenvalue is large). This causes difficulties for the simulator and results in very
long simulation times [4]. Therefore, the model is further reduced to eliminate the nonessential poles.
The balanced Gramian state space realization are obtained using the MATLAB [51] function balreal [52].
The diagonal entries of the joint Gramian, g, are evaluated and poles corresponding to g < 0.002 were
removed. The resulting model was of varying order between seven and ten, see Table 4.3.
The identified reduced models were transformed into state space models. The complete reduced order
process model also included a mass balance over the storage tank, equation (3.58), and an equation
ensuring that the demand is satisfied, equation (3.57). Furthermore, the energy requirement given by:
32
4.4 Scheduling Using the Reduced Order Model
The reduced order model was used to obtain the optimal schedule which minimizes the operating cost
of the plant while meeting product demand throughout the horizon. The scheduling problem setup was
formulated as a dynamic optimization problem to determine the optimal production rate set point. The
model was implemented and optimized using gPROMS 4.0.0 [37] and the NLPSQP solver.
where Etotal is the power usage of the plant at time t [MW h] and price is the hourly electricity price
[EUR/MWh].
The electricity price schedule was taken from the Nord Pool Spot market for UK [53]. N2EX Day
Ahead Auction prices from April 11 - 13 2015 are presented in Figure 4.3. According to Ierapetritou et
al. [11] the price can be known with reasonable certainty over a 3 day time horizon. Predicted prices on
a longer time horizon are more uncertain and require optimal scheduling under uncertainty [54].
110
100
90
Electricity Price [Eur/MWh]
80
70
60
50
40
30
0 4 8 12 16 20 24 28 32 36 40 44 48 52 56 60 64 68 72
Time [hour]
33
Impurity ≥ 0 (4.30)
Constraints are also imposed on the storage tank inventory. At the initial state the storage tank was
assumed to be 50 % full and at the end of the time horizon the storage tank was constrained to be at
least the same.
M (t0 )storage
∗ = 0.5 (4.31)
Mstorage
M (tf )storage
∗ ≥ 0.5 (4.32)
Mstorage
where M (t0 )storage is the liquid level in the storage tank at the beginning of the time horizon, M (tf )storage
∗
is the liquid level in the storage tank at the end of the time horizon and Mstorage is the maximum storage
tank capacity. Furthermore, constraints were implemented to ensure that the tank was not emptied or
filled beyond its capacity:
M (t)storage
0≤ ∗ ≤1 (4.33)
Mstorage
where M (t)storage is the liquid level in the storage tank at time t.
The scheduling decision variable, the production rate, was allowed to vary every hour by up to ±20%
of the nominal production rate. Furthermore, constraints were imposed on the rate of change of the
decision variable. In one hour, the production rate set point could not change by more then 3 mol s−1 :
dFproduction
≤3 (4.34)
dt
The decision variables were equivalent to the dynamic optimization, equation (4.19). The time horizon
was divided into 6 (variable length) control intervals for every hour. First, 5 control intervals were used
where ai,j was free to be optimized, and the last where ai,j = 0. At the sixth interval in every hour,
constraint (4.20) was imposed to ensure that the optimal values of ui,j were reached. To find the optimal
control values for each level, a polynomial was fit to interpolate the optimal values of ui,j found at each
discrete operating level in the steady state optimization, see Table 4.1. A first order polynomial was
used for the feed flow and a second order polynomial fit rgasdraw and kwithdraw .
34
Chapter 5
The results from the dynamic optimization, model identification and schedule optimization are presented
in this chapter, along with a simulation of the full dynamic model when subjected to the optimal schedule.
The results are normalized except when otherwise noted.
35
24
22
21
20
19
18
17
16
0 10 20 30 40 50 60 70 80
Time [hour]
Figure 5.1: Step changes in production set point considered during the dynamic optimization
As seen in Figure 5.2a, 5.2b, and 5.3 the impurity, the column flooding and the ratio between the dew
pressure and the pressure exiting zone 1 of the PHX (denoted dew pressure ratio henceforth) are all close
to or temporarily violate their corresponding constraints during the transition horizon. The impurity
level, which has a very slow dynamic behavior, is the most challenging constraint to satisfy during the
transitions. Moreover, the ratio between the pressure of the feed stream exiting zone 2 of the PHX and
the bubble pressure, see Figure 5.2c, (henceforth denoted bubble pressure ratio) also approaches the
constraint bounds. The heat transfer constraint, and the reboiler liquid level are not close to the bounds
throughout the trajectory, see Figure 5.2e and 5.2f.
The dashed line in Figure 5.2 shows the outputs of the reduced model, equation (4.24). The NRMSE
between the training data outputs and the reduced model prediction is also given. For production,
flooding and heat transfer the identified reduced model predictions fit the data very well with NRMSE
values between 75 % - 88 %. Especially the model for the production rate, figure 5.2d, has a good fit
with an NRMSE value of 88 %. The NRMSE between the bubble pressure ratio training data and the
corresponding reduced order model prediction was relatively low. At 25 hours and 60 hours, a significant
offset is observed between the data and the model prediction. However, most of the transitions are
accurately predicted, and the transfer function model for the bubble pressure ratio captures the dynamic
behavior of the training data acceptably. The reduced transfer function model for the impurity level
did not have a good fit (NRMSE of 34 %), due to a slight mismatch in the gain, but the time constant
appears to mimic the dynamic behavior of the process. It was determined that the fit was acceptable,
and if the schedule suggests a transition which may lead to a breach of the impurity constraint, the
underlying control system should be able to prevent these occurrences.
The reduced order transfer function model of the dew pressure ratio did not have a sufficient fit with
a NRMSE value of -74 %. In Figure 5.3, the model predicts constraint violations at several points when
the training data does not violate the constraints. A nonlinear Hammerstein-Wiener model was therefore
identified to better predict the dew point pressure. As can be seen in Figure 5.3 the nonlinear reduced
order model for the dew pressure ratio gave a significantly better prediction than the transfer function
model.
36
1
2
Data 0.8 Data
Transfer Function model Transfer Function model
1.5 Constraint level 0.6 Constraint Level
NRMSE: 34.1463 % NRMSE: 75.9094 %
0.4
1 0.2
Flooding
Impurity
0
0.5
−0.2
0 −0.4
−0.6
−0.5
−0.8
−1 −1
0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70 80
Time [hour] Time [hour]
Production
0
0
-2
−0.5
-4 Data
Transfer Function model
Constraint Level −1
NRMSE: 88.1496 %
NRMSE: 38.7112 % Data
-6
Transfer Function model
−1.5
0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70 80
Time [hour] Time [hour]
1.5
2 1
Reboiler Liquid level
0.5
1
0
UA
−0.5
0
−1
−1 −1.5
−2
−2 NRMSE: 83.2343 %
Data −2.5
Transfer Function model
−3 −3
0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70 80
Time [hour] Time [hour]
Figure 5.2: Data and transfer function models for step change responses.
37
4
Data
Transfer Function model, NRMSE: −74.6451 %
3.5 Non−linear model, NRMSE: 61.3946 %
Constraint level
3
2.5
Dew pressure
1.5
0.5
−0.5
−1
0 10 20 30 40 50 60 70 80
Time [hour]
Figure 5.3: Dew pressure data compared to transfer function model and non-linear model
Overall, the transition dataset obtained through dynamic optimization resulted in highly nonlinear
behavior of the model outputs. This is not surprising given the complex, highly coupled, and nonlinear
behavior of the equations that describe an ASU. In light of this fact, it is very encouraging that the
identified reduced order models obtained using transfer functions and Hammerstein-Wiener models were
able to predict the important dynamic behaviors of the process variables during open-loop production
rate transitions. The fits are not perfect, but the models are suitable for optimal scheduling calculations
with a significantly smaller model. Additionally, a control system would be utilized to ensure that the
constraints are satisfied during the optimal schedule.
38
1
0.8
Production set point 0.6
0.4
0.2
−0.2
−0.4
−0.6
−0.8
−1
0 10 20 30 40 50 60 70 80
Time [hour]
Figure 5.4: Schedule for production rate set point
0.9
0.8
Hold Up Storage Tank
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0
0 10 20 30 40 50 60 70 80
Time [hour]
Figure 5.5: Hold up in storage unit
The relevant process variables (reduced order model outputs) are presented in Figure 5.6. Both
the flooding and bubble pressure ratio reduced model predictions do not approach their constraints,
39
see Figure 5.6c and Figure 5.6d. The reduced model prediction of the dew pressure ratio was close to
the constraint violation throughout the schedule horizon, see Figure 5.6b. The impurity reduced model
prediction of the impurity level, Figure 5.6a, reaches its constraint bound on several occasions. Due to
the fact that the impurity reduced model prediction was the least accurate, it is important that a robust
controller is implemented to ensure that the actual system does not violate the constraint throughout
the horizon. The other reduced model output predictions are presented in the appendix.
1.5
Impurity 3 Dew Pressure
Constraint level
Constraint level
1 2.5
Dew Presure
0.5
Impurity
1.5
0 1
0.5
−0.5
0
−1 −0.5
0 20 40 60 80 0 20 40 60 80
Time [hour] Time [hour]
−1
Flooding
0 −2
−3 Bubble pressure
−4 Constraint level
−0.5
−5
−6
−1 −7
0 20 40 60 80 0 20 40 60 80
Time [hour] Time [hour]
Figure 5.6: Expected constraint and variable profiles from the dynamic scheduling using the reduced order model
for 72 hours
40
0.8 Production set point
Full dynamic model
0.6 Reduced model
NRMSE: 82.753 %
0.4
0.2
Production
−0.2
−0.4
−0.6
−0.8
−1
0 2 4 6 8 10 12 14 16 18 20 22 24 26
Time [hour]
Figure 5.7: Production of the plant using the dynamic and reduced order model for the first 24 hours of the
schedule found using the reduced order model. NRMSE between the reduced order model and the full dynamic
model is also given
0.9
0.8
0.7
Hold Up Storage Tank
0.3
0.2
0.1
−0.1
0 2 4 6 8 10 12 14 16 18 20 22 24
Time [hour]
Figure 5.8: Hold up in storage during 24 hours using the dynamic and reduced model
41
The impurity level is shown in Figure 5.9a for the full dynamic model and the reduced order model
prediction. While the NRMSE between the reduced model prediction and the full model is 36 %,
the model accurately predicts the impurity spikes, and captures similar dynamic features of the full
dynamic model. The reduced model prediction of the impurity during the optimal 3 day schedule
(Figure 5.6a) suggests that the impurity level would reach its bounds on several occasions. The impurity
level determined using the full dynamic model over the first 24 hours indicates that these bounds are in
fact reached as predicted by the reduced order model.
In Figure 5.9b, the dew pressure ratio determined by the full dynamic model and the reduced model
prediction is shown over the first 24 hours of the optimal schedule. The reduced order model accurately
predicts the time periods in which the dew pressure ratio is close to the constraint bounds. Other
model outputs including the flooding fraction, the bubble pressure ratio, and the heat transfer constraint
display a tight fit between the reduced order model predictions and the full dynamic model outputs (see
Figures 5.9c, 5.9d, and 5.9e). The reboiler liquid level reduced model is unable to satisfactorily predict
the dynamic behavior determined by the full model, however, the variable is not close to the constraint
bound throughout the horizon (see Figure 5.9f).
The results indicate that the identified reduced order models are capable of finding an optimal feasible
production rate schedule that accurately predicts the transient behavior of relevant process variables. The
reduced order model consisted of 525 equations and 67 differential variables compared to the full dynamic
model which consisted of 5079 equations and 437 differential variables. The resulting optimization
problem formulation is tractable and can be solved in a relatively short time frame (14861 CPU seconds)
compared to an equivalent optimization formulation with a full dynamic model (which would probably
take weeks to solve with the most advanced optimization solvers). The results should be compared to
those determined using a static scheduling approach, where the production rate setpoint schedule is
determined under the assumption that some transitions are infeasible. While this approach is much
simpler, it is unlikely that the constrained variables will pushed to their bounds, or that the optimal
schedule is actually feasible. Consequently, it is expected that the optimal schedule determined using a
dynamic model will find a better result.
42
1 Full dynamic model
3.5
Reduced model
0.8 Constraint level
3 NRMSE: 39.4096 %
0.6
2.5
0.4
Dew Pressure
0.2 2
Impurity
0
1.5
−0.2
1
−0.4
0.5
−0.6
NRMSE: 36.0379 %
−0.8 Full dynamic model 0
Reduced model
Constraint level
−1
−0.5
0 2 4 6 8 10 12 14 16 18 20 22 24 0 2 4 6 8 10 12 14 16 18 20 22 24
Time [hour] Time [hour]
−1
0
−2
−0.2
−3
−0.4
−4
−0.6
NRMSE: 66.3762 % −5 NRMSE: 36.5008 %
Full dynamic model Full dynamic model
−0.8 Reduced model −6 Reduced model
Constraint level Constraint level
−1
−7
0 2 4 6 8 10 12 14 16 18 20 22 24 0 2 4 6 8 10 12 14 16 18 20 22 24
Time [hour] Time [hour]
0.5 0
−0.5
UA
0
−1
−0.5
−1.5
−1
−2
−1.5 −2.5 NRMSE: −34.2085 %
Full dynamic model
−2 −3 Reduced model
0 2 4 6 8 10 12 14 16 18 20 22 24 0 2 4 6 8 10 12 14 16 18 20 22 24
Time [hour] Time [hour]
Figure 5.9: Variables and constraints using the full dynamic model and the reduced order model.
43
Chapter 6
A novel framework for making optimal scheduling decisions while incorporating process dynamics is
described in this thesis. The concept is to establish a reduced order model that predicts the dynamic
behavior of relevant process variables, e.g., those involved in operational constraints or process outputs,
when subject to changes in the production setpoint schedule. The optimal schedule is then determined
using the reduced order dynamic process model. This significantly differs from the conventional methods
using static models of full dynamic process models. The framework provides the benefit of using the
dynamic process knowledge to place process variables at their limits (if needed) in order to more rapidly
transition between production levels, while maintaining the process within its operational constraints.
The benefits of the framework are illustrated with an ASU that produces nitrogen at 99.9 % purity and
subjected to time-varying electricity prices. A detailed dynamic model of the ASU was developed which
included a cryogenic distillation column with reboiler/condenser, a PHX, a compressor, two turbines
and a liquefier, and consisted of 5079 equations with 437 differential variables. A simulated production
rate transition dataset was obtained using dynamic optimization on the dynamic model to determine the
optimal manipulated variable trajectories throughout the selected transitions. These data were used to
identify an accurate reduced order model consisting of 525 equations and 67 differential variables which
captures the dynamic behavior of important process variables (i.e., process outputs and constrained
variables). The reduced model utilizes transfer function models for all but one output and a nonlinear
Hammerstein-Wiener model for the the other output. An optimal production rate setpoint schedule
was determined for the ASU when subject to time-varying electricity prices with the use of a storage
tank. The optimization sought to minimize electricity cost throughout the time horizon by adjusting the
production rate setpoint while ensuring that the operational constraints were satisfied. As expected, the
optimal schedule utilized the stored liquid nitrogen to meet demand when electricity prices were high,
and filled the storage tank with liquid nitrogen when electricity prices were low. The optimal schedule
realized a 2.6 % savings compared to a flat production rate over the 3 day period considered. Finally,
the optimal schedule was implemented in the full dynamic model and again the optimal manipulated
variable transition trajectories were determined using dynamic optimization. The result showed that the
optimal schedule was feasible and that the predicted outputs from the reduced order model showed a
close match to their corresponding variables in the full model.
The limitation of the proposed framework is the ability to develop reduced order models that can
accurately predict the important process variables throughout production transitions. Given the complex,
highly coupled, and nonlinear nature of the equations that describe an ASU, it is encouraging that
accurate reduced order models can be obtained for the important process variables using relatively few
state variables. It was observed from the optimal trajectories that the manipulated variables followed
certain paths. When constructing the reduced order model it was advantageous to first related the input
variable to the manipulated variables with a transfer functions. The manipulated variables was thereafter
related to the output variables to obtain the overall transfer function. This structured approach gave
good results when identifying the reduced order model.
The framework presented in this thesis is generally applicable to optimal scheduling of processes
with variable product qualities and production rates. It is an alternative to the conventional scheduling
approaches which either use a static model with the process dynamics lumped into transition times, or a
full dynamic process model. In this work a full dynamic model of the plant was necessary to simulate the
44
transition data, but the framework can be applied when only historical transition data from the plant is
available (i.e., a full dynamic process model is not required if rich historical data is present). Finally, in
the presence of unexpected disturbances, the framework provides a quick and robust way to re-schedule
the plant operation on-line.
45
Bibliography
[1] M.M. El-Halwagi. Introduction to process integration. In Process Integration, volume 7 of Process
Systems Engineering, pages 1 – 20. Academic Press, 2006.
[2] M. Baldea and I. Harjunkoski. Integrated production scheduling and process control: A systematic
review. Computers & Chemical Engineering, 71:377–390, 2014.
[3] D.E. Shobrys and D.C. White. Planning, scheduling and control systems: why cannot they work
together. Computers & Chemical Engineering, 26(2):149 – 160, 2002.
[4] D.E. Seborg, D.A. Mellichamp, T.F. Edgar, and F.J. Doyle III. Process dynamics and control. John
Wiley & Sons, 2010.
[5] T.F. Edgar. Control and operations: when does controllability equal profitability? Computers &
Chemical Engineering, 29(1):41 – 49, 2004. {PSE} 2003.
[6] J. Du, J. Park, I. Harjunkoski, and M. Baldea. A time scale-bridging approach for integrating
production scheduling and process control. Computers & Chemical Engineering, 79(0):59 – 69,
2015.
[7] I. Harjunkoski, R. Nyström, and A. Horch. Integration of scheduling and control—theory or practice?
Computers & Chemical Engineering, 33(12):1909 – 1918, 2009. {FOCAPO} 2008 – Selected Papers
from the Fifth International Conference on Foundations of Computer-Aided Process Operations.
[8] J. Park, J. Du, I. Harjunkoski, and M. Baldea. Integration of scheduling and control using internal
coupling models. In J.J. Klemes, P.S. Varbanov, and P.Y. Liew, editors, 24TH European Symposium
on Computer Aided Process Engineering, PTS A and B, volume 33 of Computer-Aided Chemical
Engineering, pages 529–534, 2014.
[9] Y. Cao, C.L.E. Swartz, and M. Baldea. Design for dynamic performance: Application to an air
separation unit. In American Control Conference (ACC), 2011, pages 2683–2688, June 2011.
[10] M.H. Karwan and M.F. Keblis. Operations planning with real time pricing of a primary input.
Computers & operations research, 34(3):848–867, 2007.
[11] M.G. Ierapetritou, D. Wu, J. Vin, P. Sweeney, and M. Chigirinskiy. Cost minimization in an energy-
intensive plant using mathematical programming approaches. Industrial & engineering chemistry
research, 41(21):5262–5277, 2002.
[12] D.R. Vinson. Air separation control technology. Computers & Chemical Engineering, 30(10):1436–
1446, 2006.
[13] A.R. Smith and J. Klosek. A review of air separation technologies and their integration with energy
conversion processes. Fuel Processing Technology, 70(2):115 – 134, 2001.
[14] R. Huang, V.M. Zavala, and L.T. Biegler. Advanced step nonlinear model predictive control for air
separation units. Journal of Process Control, 19(4):678 – 685, 2009.
[15] T. Guo, J. Lu, W. Xiang, Y. Guo, and K. Wu. Dynamic modeling and control of the air separation
unit in an igcc power plant. In Sustainable Power Generation and Supply, 2009. SUPERGEN’09.
International Conference on, pages 1–7. IEEE, 2009.
46
[16] B. Roffel, B.H.L. Betlem, and J.A.F. De Ruijter. First principles dynamic modeling and multivari-
able control of a cryogenic distillation process. Computers & Chemical Engineering, 24(1):111–123,
2000.
[17] Y. Zhu, S. Legg, and C.D. Laird. Optimal design of cryogenic air separation columns under uncer-
tainty. Computers & Chemical Engineering, 34(9):1377 – 1384, 2010. Selected papers from the 7th
International Conference on the Foundations of Computer-Aided Process Design (FOCAPD, 2009,
Breckenridge, Colorado, USA.
[18] Energy Information Association. Manufacturing energy consumption survey: Total consumption of
electricity, 2010. http://www.eia.gov/consumption/manufacturing/data/2010/pdf/Table11_
1.pdf.
[19] R. Agrawal and T.F. Yee. Heat pumps for thermally linked distillation columns: An exercise for
argon production from air. Industrial & engineering chemistry research, 33(11):2717–2730, 1994.
[20] R. Agrawal, D.W. Woodward, K.A. Ludwig, and D.L. Bennett. Impact of low pressure drop struc-
tured packing on air distillation. In Institution of Chemical Engineers Symposium Series, volume
128, pages A125–A125. Hemsphere Publishing Corporation, 1992.
[21] M. Aneke and M. Wang. Potential for improving the energy efficiency of cryogenic air separation
unit (ASU) using binary heat recovery cycles . Applied Thermal Engineering, 81(0):223 – 231, 2015.
[22] J. Rizk, M. Nemer, and D. Clodic. A real column design exergy optimization of a cryogenic air
separation unit. Energy, 37(1):417 – 429, 2012. 7th Biennial International Workshop “Advances in
Energy Studies”.
[23] Yasuki Kansha, Akira Kishimoto, Tsuguhiko Nakagawa, and Atsushi Tsutsumi. A novel cryogenic
air separation process based on self-heat recuperation. Separation and Purification Technology,
77(3):389 – 396, 2011.
[24] W.F. Castle. Air separation and liquefaction: recent developments and prospects for the beginning
of the new millennium. International Journal of Refrigeration, 25(1):158 – 172, 2002.
[25] B. Keenan and K. Reuter. The design, experience and economic advantage of cryogenic turbo-
expanders with active magnetic bearings. In 3rd International Conference Kryogenica’94, 1994.
[26] R. Boehme, J.A.R. Parise, and R. Pitanga Marques. Simulation of multistream plate–fin heat
exchangers of an air separation unit. Cryogenics, 43(6):325 – 334, 2003.
[27] A.R. Smith, J. Klosek, and D.W. Woodward. Next-generation integration concepts for air separation
units and gas turbines. Journal of Engineering for Gas Turbines and Power, 119(2):298–304, 1996.
[28] B. Daryanian, R.D. Tabors, and R.E. Bohn. Optimal demand-side response to electricity spot prices
for storage-type customers. IEEE Trans. Power Syst.;(United States), 4(3), 1989.
[29] J.G. Roos and I.E. Lane. Industrial power demand response analysis for one-part real-time pricing.
Power Systems, IEEE Transactions on, 13(1):159–164, 1998.
[30] Y. Zhu, S. Legg, and C.D. Laird. A multiperiod nonlinear programming approach for operation of
air separation plants with variable power pricing. AIChE Journal, 57(9):2421–2430, 2011.
[31] S. Mitra, I.E. Grossmann, J.M. Pinto, and N. Arora. Optimal production planning under time-
sensitive electricity prices for continuous power-intensive processes. Computers & Chemical Engi-
neering, 38(0):171 – 184, 2012.
[32] J. Miller, W.L. Luyben, and S. Blouin. Economic incentive for intermittent operation of air sepa-
ration plants with variable power costs. Industrial & Engineering Chemistry Research, 47(4):1132–
1139, 2008.
[33] C. Zhongzhou, M.A. Henson, P. Belanger, and L. Megan. Nonlinear model predictive control of
high purity distillation columns for cryogenic air separation. Control Systems Technology, IEEE
Transactions on, 18(4):811–821, July 2010.
47
[34] G.Y. Zhu, M.A. Henson, and L. Megan. Low-order dynamic modeling of cryogenic distillation
columns based on nonlinear wave phenomenon. Separation and Purification Technology, 24(3):467
– 487, 2001.
[35] Y. Fu and X.G. Liu. Nonlinear control based on wave model of high-purity heat integrated air
separation column. Chemometrics and Intelligent Laboratory Systems, 146(0):1 – 9, 2015.
[36] R. Lopez-Negrete, F.J D’Amato, L.T Biegler, and A. Kumar. Fast nonlinear model predictive
control: Formulation and industrial process applications. Computers & Chemical Engineering,
51(0):55 – 64, 2013.
[37] Process Systems Enterprise. gPROMS 4.0.0. www.psenterprise.com/gproms, 1997 - 2014.
[38] Y. Cao. Design for dynamic performance: Application to an air separation unit, 2011.
[39] Aspen Technology, Inc. Aspen Plus. https://www.aspentech.com/products/aspen-plus.aspx,
1994 - 2015.
[40] C.L. Yaws. The Yaws Handbook of Vapor Pressure: Antoine coefficients. Elsevier Science, 2015.
[41] A. Harmens. Vapour-liquid equilibrium N2 -A2 -O2 for lower argon concentrations. Cryogenics,
10(5):406–409, 1970.
[42] W.M Haynes. CRC Handbook of Chemistry and Physics. CRC press, 2013.
[43] ALPEMA. The Standards of the Brazed Aluminum Plate-Fin Heat Exchangers Manufacturers’s
Association, 2000.
[44] D. Green and R. Perry. Perry’s Chemical Engineers’ Handbook, Eighth Edition. Chemical Engineers
Handbook. Mcgraw-hill, 2007.
[45] E.W. Lemmon and S.G. Penoncello. The surface tension of air and air component mixtures. In
P. Kittel, editor, Advances in Cryogenic Engineering, volume 39 of Advances in Cryogenic Engi-
neering, pages 1927–1934. Springer US, 1994.
[46] C.F. Spencer and S.B. Adler. A critical review of equations for predicting saturated liquid density.
Journal of Chemical & Engineering Data, 23(1):82–89, 1978.
[47] G. Venkatarathnam and K.D. Timmerhaus. Cryogenic mixed refrigerant processes. Springer, 2008.
[48] R.K. Sinnott and G. Towler. Chemical Engineering Design. Elsevier, 5th edition, 2009.
[49] R.M. Thorogood, D.L. Bennett, B.K. Dawson, A.L. Prentice, and R.J. Allam. Air separation process
using packed columns for oxygen and argon recovery, November 25 1992. Patent EP 0341854 B1.
[50] T. Kai, H. Nomoto, M. Deguchi, and T. Takahashi. Surface tension of ternary mixtures of nitrogen,
oxygen, and argon. Journal of Chemical & Engineering Data, 39(3):499–501, 1994.
[51] MathWorks, Inc. MATLAB Student R2014a. http://www.mathworks.com/products/matlab/,
1994 - 2014.
[52] A.J. Laub, M.T. Heath, C. Paige, and R. Ward. Computation of system balancing transforma-
tions and other applications of simultaneous diagonalization algorithms. Automatic Control, IEEE
Transactions on, 32(2):115–122, Feb 1987.
[53] Nord Pool Spot AS. N2EX Day Ahead Auction Prices. http://www.nordpoolspot.com/
Market-data1/N2EX/Auction-prices/UK/Hourly/?view=chart. [Online; acessed 15-May-2015].
[54] Q. Zhang, I.E. Grossmann, C.F. Heuberger, A. Sundaramoorthy, and J.M. Pinto. Air separation
with cryogenic energy storage: Optimal scheduling considering electric energy and reserve markets.
AIChE Journal, 61(5):1547–1558, 2015.
[55] S.S. Rao. Introduction to optimization. In Engineering Optimization, pages 1–62. John Wiley &
Sons, Inc., 2009.
48
[56] S. Skogestad. Self-optimizing control: the missing link between steady-state optimization and con-
trol. Computers & Chemical Engineering, 24(2):569–575, 2000.
[57] W.F. Feehery and P.I. Barton. Dynamic optimization with equality path constraints. Industrial &
Engineering Chemistry Research, 38(6):2350–2363, 1999.
[58] A. Cervantes and L.T. Biegler. Optimization strategies for dynamic systems optimization strategies
for dynamic systems. In C.A. Floudas and P.M. Pardalos, editors, Encyclopedia of Optimization,
pages 2847–2858. Springer US, 2009.
[59] V.S. Vassiliadis, R.W.H. Sargent, and C.C. Pantelides. Solution of a class of multistage dynamic
optimization problems. 1. problems without path constraints. Industrial & Engineering Chemistry
Research, 33(9):2111–2122, 1994.
[60] L.T. Biegler. Nonlinear programming strategies for dynamic chemical process optimization. Theo-
retical Foundations of Chemical Engineering, 48(5):541–554, 2014.
[61] S.S. Rao. Nonlinear programming iii: Constrained optimization techniques. In Engineering Opti-
mization, pages 380–491. John Wiley & Sons, Inc., 2009.
[62] J. Sjöberg, Q. Zhang, L. Ljung, A. Benveniste, B. Delyon, P.-Y. Glorennec, H. Hjalmarsson and, A.
Juditsky. Nonlinear black-box modeling in system identification: a unified overview. Automatica,
31(12):1691 – 1724, 1995.
[63] L. Ljung. System identification. In Wiley Encyclopedia of Electrical and Electronics Engineering.
John Wiley & Sons, Inc., 2001.
[64] D. Zill and M. Cullen. Differential Equations with Boundary-Value Problems. Cengage Learning,
2008.
[65] A.P. Wardle. Chapter 7 - Process Control . In J.F. Richardson D.G. Peacock, editor, Coulson &
Richardson’s Chemical Engineering Volume 3, pages 560 – 736. Butterworth-Heinemann, Oxford,
third edition edition, 1991.
[66] A. Wills, T.B. Schön, L. Ljung, and B. Ninness. Identification of Hammerstein–Wiener models .
Automatica, 49(1):70 – 81, 2013.
49
Appendix A
Concepts
The system of ODE can be solved using several different numerical methods. Examples include the Euler
method, the trapezoidal method, the Taylor series method, the Runge-Kutta method, and the linear
multi-step method [38]. Additionally, several solvers (e.g., gPROMS [37]) are capable of integrating
index 1 DAE systems without transforming the algebraic equations. However, it is possible that higher
index DAE systems arise in process models. For example, if equation (A.1c) is replaced with:
0 = x1 − 6x2 (A.3)
The algebraic equation would need to be differentiated twice before the system can be transformed into
a set of ODE:
0 = x01 − 6x02 = 3x1 − 5x2 − 23y1 (A.4)
3x01 − 5x02
y10 = (A.5)
−23
This DAE system is of index 2. Such problems typically cannot be solved in this form. Hidden con-
straints in the algebraic equation, stability restrictions, and the difficulty in determining consistent initial
conditions are some of the problems with high index DAE. Dynamic models of distillation columns, for
example a cryogenic distillation column, are typically index 2 [14, 38]. To solve a high index DAE, the
problem needs to be reformulated into index 1 to reduce the order (see section 3.3.1 for an example of
index reduction). This can be costly depending on the size and complexity of the original DAE system.
Direct integration approaches without order reduction may be used, but the systems must have a certain
structure for this option to be viable [38].
50
A.2 Optimization
Broadly speaking, there are 2 classes of process optimization, steady state optimization and dynamic
optimization. Steady state optimization is used, for example, to find the optimal equipment design
and the optimal operating parameters, such as flow rates, temperatures, pressure, etc., to minimize the
capital and the operating cost of the nominal operation. Dynamic optimization is typically used to find
the optimal control moves between operating points over a time horizon to maximize the plant efficiency
and profit [4].
Subject to
F (x, z, θ) = 0 (A.6b)
g(x, z, θ) ≤ 0 (A.6c)
Likewise, the general form of a dynamic optimization problem is described as [57, 58]:
Subject to:
f (ẋ(t), x(t), y(t), u(t), t, θ) = 0 (A.7b)
h(x(t), y(t), u(t), t, θ) = 0 (A.7c)
g(x(t), y(t), u(t), t, θ) ≤ 0 (A.7d)
With initial conditions:
x(t0 ) = x0 (A.7e)
where φ is the objective function, x and y are time dependent differential and algebraic variables, u are the
control variables and θ are the time invariant parameters. Equation (A.7b) describes the DAE system
that the objective function, equation (A.7a), is subjected to with initial conditions of the differential
equations described in equation (A.7e). Equation (A.7c) and (A.7d) describes the equality and inequality
constraints of the system.
51
Polynomial with
continuity
Piecewise
linear with
continuity
u(t)
Piecewise
linear
Piecewise
constant
t
Control Interval
A.3.1 Scheduling
An optimal schedule can be determined using a static or dynamic plant model. Conventional approaches
use a static plant model. The static model incorporates the dynamics of the process by including a table
of transition times, τs , which are determined off-line. Allowable transitions between operating states
in the process are also included in the table. The table guarantees that no constraints are violated.
However, it is beneficial from a profitability perspective to account for the detailed process dynamics if
multiple products are produced in order to determine the minimal (feasible) transition time. This calls
for a close integration of scheduling with the control system to maximize profit and minimize the effect
of disturbances on the system [6].
Full dynamic scheduling uses a detailed dynamic plant model in the optimal scheduling problem.
A full dynamic scheduling approach can describe the dynamic behavior during transitions and likely
find the optimal production schedules, while ensuring that operational constraints are met. However,
the optimization problem can quickly become exceedingly large and difficult to find a solution in a
52
reasonable time frame. To overcome this problem, a reduced order model can be used to represent the
dynamic behavior of important process variables [2, 6].
Y (s)
G(s) = (A.9)
X(s)
where Y (s) and X(s) are usually described by polynomials in the Laplace domain. The roots of X(s) = 0
and Y (s) = 0 are called the poles and zeros of the transfer function, respectively. The number of
poles must be greater then the number of zeros for the transfer function to be physically realizable.
Furthermore, the number of poles describes the order of the system [4, 65].
A transfer function is used to describe the dynamic behavior of a system and relates one input, x, to
one output, y. The order of the transfer function corresponds to the order of the differential equation
which it represents. A series of transfer functions is commonly used in process control when relating the
input signal to the output through several “blocks” or sub-systems (e.g., relating the valve position to the
flow rate, and the flow rate to the outlet temperature), see Figure A.2. The resulting transfer function
is then the product of the two transfer functions [4]:
G(s) = G1 ∗ G2 (A.10a)
y = G(s)x1 (A.10b)
53
x1 G1 x2 G2 y
Transfer functions in parallel (e.g., Figure A.3) can be added together to obtain the overall transfer
function:
G(s) = G1 + G2 (A.11a)
y = G(s)x (A.11b)
G1
+
u y
+
G2
An example of a non-linear “black box” model is the Hammerstein-Wiener model [66]. Such models
consists of a linear transfer function dynamic block with static nonlinear blocks which transform the
inputs and outputs, as seen in Figure A.4. The nonlinear blocks can have different structures, e.g.,
piecewise linear, networks, or polynomial. An example of a Hammerstein-Wiener model with a second
order polynomial as the input and output nonlinear blocks is presented below:
u0 = Au2 + Bu + C (A.12a)
0 0
y = Gu (A.12b)
y = ay 02 + by 0 + c (A.12c)
where u is the input to the model, y is the output, G is the transfer function model, and A, B, C, a, b,
c are parameters.
u u’ y’ y
Input Non- Output Non-
Linear Block
Linearity Linearity
54
Appendix B
Results
0.5117 0.032
Data points Data points
Fitted polynomial 0.03
Fitted polynomial
0.5116
kwithdrawn [mol N2 reflux/mol N2 total]
0.028
rgasdraw [mol liquid air/mol air total]
0.5115
0.5114 0.026
0.5113 0.024
0.5112 0.022
0.5111 0.02
0.511 0.018
0.5109 0.016
0.5108 0.014
16 17 18 19 20 21 22 23 24 16 17 18 19 20 21 22 23 24
Production rate [mol/s] Production rate [mol/s]
(a) Fraction liquid air entering the column (b) Fraction of product entering the condenser
48
Data points
46
Fitted polynomial
44
42
F1,4 [mol/s]
40
38
36
34
32
30
16 17 18 19 20 21 22 23 24
Production rate [mol/s]
Figure B.1: Optimal control variables for different production rates with polynomial approximations
B.2 Outputs From the Scheduling Using the Reduced Order Model
In Figure B.2 the U A value, the reboiler liquid level and the production rate expected from the scheduling
using the reduced order model is given for 72 hours.
55
3 1.5
2 1
0 0
−1 −0.5
−2 −1
−3 −1.5
0 20 40 60 80 0 20 40 60 80
Time [hour] Time [hour]
0.5
Production
−0.5
−1
−1.5
0 20 40 60 80
Time [hour]
(c) Production
Figure B.2: Expected variable and constraint outputs from the scheduling using the reduced order model for 72
hours
56
www.kth.se