Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Computers and Chemical Engineering 126 (2019) 22–34

Contents lists available at ScienceDirect

Computers and Chemical Engineering


journal homepage: www.elsevier.com/locate/compchemeng

Optimal demand response scheduling of an industrial air separation


unit using data-driven dynamic modelsR
Calvin Tsay a, Ankur Kumar b, Jesus Flores-Cerrillo b, Michael Baldea a,∗
a
McKetta Department of Chemical Engineering, The University of Texas at Austin, Austin, TX 78712, United States
b
Smart Operations–Praxair Digital, Praxair, Inc., Tonawanda, NY 14150, United States

a r t i c l e i n f o a b s t r a c t

Article history: Managing electricity demand has become a key consideration in power grid operations. Industrial de-
Received 19 December 2018 mand response (DR) is an important component of demand-side management, and electricity-intensive
Revised 16 March 2019
chemical processes can both support power grid operations and derive economic benefits from electric-
Accepted 16 March 2019
ity price fluctuations. For air separation units (ASUs), DR participation calls for frequent production rate
Available online 25 March 2019
changes, over time scales that overlap with the dominant dynamics of the plant. Production scheduling
Keywords: calculations must therefore explicitly consider process dynamics. We introduce a data-driven approach
Integrated scheduling and control for learning the DR scheduling-relevant dynamics of an industrial ASU from its operational history, and
Scale-bridging models present a dynamic optimization-based DR scheduling framework. We show that a class of low-order
Demand-side management Hammerstein-Wiener models can accurately represent the dynamics of the industrial ASU and its model
Electricity markets predictive control system. We evaluate the economic benefits of the proposed scheduling framework, and
Air separation units
analyze their sensitivity to electricity price uncertainty.
© 2019 Elsevier Ltd. All rights reserved.

1. Introduction ior is an apparent “win-win”, reducing costs for electricity users


while also mitigating the (temporary) mismatch between supply
The fundamental mission of the power grid, a large network of and demand in the power grid. Industrial manufacturing plants
generators, transmission lines, and transformers that delivers elec- are a natural candidate for load shifting operation. Intuitively, they
tricity to industrial, commercial, and residential users, is to ensure may be able to overproduce product(s) when electricity prices are
that electricity demand is exactly met by the available supply at low, store excess product(s), and fill orders using stored product(s)
all times. This task is not trivial, as user electricity demand fluctu- when grid load peaks and electricity prices are high. Participa-
ates considerably throughout the day. The increased contribution tion in a real-time electricity market is largely voluntary for in-
of intermittent renewable energy sources to the electricity gen- dustrial consumers, but can become economically attractive when
eration portfolio introduces additional short-term and long-term price variations are sufficiently high, excess production capacity is
(daily to seasonal) fluctuations in electricity supply, with the time available, and excess product(s) can be stored safely and at reason-
of daily peak renewable generation rates being typically desyn- able cost.
chronized from demand-side peaks. As a result, electricity prices In the field of chemical processing, cryogenic air separation
can fluctuate by several orders of magnitude during a 24-h period. units (ASUs) have been identified as a prime candidate for DR op-
In turn, this provides a strong incentive for demand-side manage- eration, as they consume large amounts of electricity. Air sepa-
ment, or measures of improving the energy system on the user ration units are treated as utility suppliers by their downstream
side (Conejo et al., 2002). In this context, demand response (DR), customers (Tsay et al., 2018b) who may place explicit expecta-
or “load shifting”, refers to intentional modification by an electric- tions on quality and availability of products (oxygen, nitrogen,
ity user of its power consumption over time in order to exploit and argon), but are otherwise agnostic to time of production.
time-dependent electricity prices (Zhang et al., 2015). This behav- Load shifting behavior can be achieved by ramping up produc-
tion during periods of low electricity prices and storing excess
products as cryogenic liquids, which are then vaporized to sat-
R
A preliminary version of this work was presented at the 13th International isfy gas product demand during high-price periods. Traditionally,
Symposium on Process Systems Engineering (PSE 2018), San Diego, CA (Tsay et al.,
production scheduling and process control for ASUs (and chem-
2018a)

Corresponding author. ical processes in general) have operated as distinct, hierarchi-
E-mail address: mbaldea@che.utexas.edu (M. Baldea). cal decision-making layers owing to a separation of time scales

https://doi.org/10.1016/j.compchemeng.2019.03.022
0098-1354/© 2019 Elsevier Ltd. All rights reserved.
C. Tsay, A. Kumar and J. Flores-Cerrillo et al. / Computers and Chemical Engineering 126 (2019) 22–34 23

(Baldea and Harjunkoski, 2014). Conventional methods for com- operating points and identify the optimal production schedule.
puting optimal production schedules therefore implicitly assume Zhao et al. (2019) proposed a state-transition network model for
that a chemical process is at a steady state before and after a scheduling ASUs similar to that of Basán et al. (2018), and ap-
change in production targets. However, electricity prices in current, plied it to the simultaneous scheduling of two multi-product ASUs.
deregulated markets may change at hourly (or faster) time inter- Obermeier et al. (2019) defined a mode-based scheduling approach
vals, and production scheduling decisions must therefore be made to examine relationships between DR scheduling and equipment
over a time scale in which process dynamics and control become fatigue.
highly relevant (Zhang and Grossmann, 2016). Specifically, produc- On the other hand, simplified (e.g., by reducing dimension-
tion targets must change over hourly intervals, and process safety ality) dynamic models can be employed for computationally
and product purity must be accounted for when improving process tractable scheduling calculations as an alternative to assuming
economics by such rapid production modulation. The decision- quasi-stationary modes. Dias et al. (2018) showed that model pre-
making time scale is further shortened in faster electricity markets dictive control (MPC) can be integrated in production scheduling
(e.g., the fifteen minute market), when it is desired to provide an- calculations with low computational effort using a nested decision-
cillary services, or place electricity bids in real-time (Dowling et al., making structure (Zhuge and Ierapetritou, 2015). A simulation-
2017; Otashu and Baldea, 2018; Schäfer et al., 2019b). optimization framework was proposed, where simple state-space
In light of these developments, the integration of scheduling models derived via system identification were employed to predict
and control for chemical processes is an important research area. the open-loop process dynamics. We note that economic MPC can
Recent advances in this field have been the subject of several directly incorporate the economic objectives usually considered in
recent review papers (Baldea and Harjunkoski, 2014; Daoutidis scheduling into the control layer (Baldea and Harjunkoski, 2014);
et al., 2018; Dias and Ierapetritou, 2019), but, for the sake of com- however, optimizing the (large-scale) process model in real-time
pleteness, a brief overview of relevant works is presented here. and considering a time horizon relevant to production scheduling
In an initial effort, Flores-Tlacuahuac and Grossmann (2006) pro- requires some measure of compromise, e.g., a sub-optimal update
posed explicitly including the dynamic process model in the approach (Caspari et al., 2018).
scheduling problem, representing the scheduling and control lay- Recent work from our group has suggested system identifica-
ers as a single, unified, simultaneous dynamic optimization prob- tion as a means to accurately capture closed-loop input-output re-
lem cast as a mixed-integer nonlinear program (MINLP). This lationships with low-order models referred to as “scale-bridging
discretized-time approach was later implemented in closed-loop models” (SBMs), which can then be naturally included in pro-
by Zhuge and Ierapetritou (2012). Beal et al. (2018) provided a fur- cess scheduling calculations (Baldea et al., 2015; Du et al., 2015).
ther extension to allow for time-dependent parameters and con- Pattison et al. (2016) demonstrated these ideas on a simulated,
straints, while still accounting for the full dynamic process model. single-product ASU, including a closed-loop scheduling, moving-
Nyström et al. (2005) simplified the integrated problem by sep- horizon implementation (Pattison et al., 2017). The work was later
arating it into a scheduling master problem and a control sub- extended by Kelley et al. (2018), where the authors expedited the
problem, achieving integrated scheduling and control by iterating computational solution of the scheduling problem by using a fully-
between the two. Similarly, Chu and You (2012) simplified the discretized time domain and an exact MILP reformulation.
scheduling and control problem by using linear control concepts, In this work, we present a related methodology for creating
computing optimal tuning parameters for a set of PI controllers of- low-order, scheduling-relevant models of a multi-product indus-
fline. Burnak et al. (2018) proposed capturing fast process dynam- trial ASU using historical operating data provided by our indus-
ics with surrogate models in a multiparametric scheduling frame- trial partner. Focusing on demand-response operation, we iden-
work. tify dynamic models for economically-relevant variables (produc-
For large-scale models, such as those that describe ASUs, com- tion rates, power consumption) and the process variables found to
puting schedules through dynamic optimization with detailed first- limit the dynamic agility of the process and its ability to shift op-
principles models is typically computationally intractable in practi- erations to fast-changing electricity price signals. Furthermore, this
cal amounts of time (Dias and Ierapetritou, 2017). To examine the work accounts for the fact that the industrial ASU operates un-
feasibility of DR production-rate modulation in ASUs, Cao et al. der model predictive control (MPC). The industrial ASU also pro-
(2011, 2016a) presented initial results on dynamic modeling and duces multiple products, in contrast to related works investigating
optimization using a first-principles process model by restricting a simulated ASU process that produces only nitrogen (Dias et al.,
the time horizon to individual production rate transitions. The 2018; Pattison et al., 2017; 2016). We incorporate the scheduling-
same authors subsequently examined alternative approaches for relevant models into a DR scheduling problem formulated as a dy-
the collection and usage of liquid storage using collocation-based namic optimization and present an extensive case study to exam-
dynamic models (Cao et al., 2016b). More recently, Schäfer et al. ine the benefits of DR operation. Although re-scheduling (Gupta
(2019a) applied a compartmentalization-based model reduction et al., 2016; Pattison et al., 2017) is not considered in this study,
approach for ASUs and demonstrated improved computational per- the identified low-order models allow the scheduling problem to
formance compared to collocation-based approaches. be solved fast enough for online scheduling and can be easily up-
In an effort to balance computational complexity with captur- dated if the controller and/or the plant are modified.
ing (some of) the relevant process dynamics in production schedul-
ing, many works assume quasi-stationary modes of operation with 2. Background
additional constraints tailored to reflect the transition capabili-
ties of the plant and control considerations (Misra et al., 2017; 2.1. Description of the industrial air separation unit under
Zhang et al., 2015, 2016; Zhou et al., 2017). Zhang et al. (2016) ex- consideration
tended the mode-based scheduling framework presented by Mitra
et al. (2012, 2014) by introducing surrogate sub-process mod- ASUs separate the components of air predominantly for use
els that allow for computationally efficient scheduling of contin- in other manufacturing processes. For example, argon is used in
uous process networks. Misra et al. (2017) used a state-task net- welding, oxygen in steel production, and nitrogen in food and met-
work to model production constraints. With a similar motiva- als processing. High-purity oxygen also has medical applications.
tion, Zhou et al. (2017) defined a set of ASU operating modes Although several air separation technologies are available, the bulk
from historical data, and used convex hulls to reflect feasible of high-purity, high-volume industrial gas production is dominated
24 C. Tsay, A. Kumar and J. Flores-Cerrillo et al. / Computers and Chemical Engineering 126 (2019) 22–34

Fig. 1. Industrial air separation unit process flowsheet.

by cryogenic ASUs, which consume immense amounts of energy tional refrigeration in the MHEX. The waste nitrogen stream is then
and primarily rely on electric compressors to handle and compress vented to the atmosphere. Purified oxygen product is collected at
their air feed streams. The industrial cryogenic ASU considered in the bottom of the UC. A portion of the oxygen product is vaporized
this work produces five products: liquid nitrogen (LN2 ), gaseous ni- through the MHEX, while the remainder is passed through the SH.
trogen (GN2 ), liquid oxygen (LO2 ), gaseous oxygen (GO2 ), and liq- The air fed to the process contains a small amount of argon, which
uid argon (LAr). The process operates under model predictive con- accumulates on an intermediate stage close to the bottom of the
trol (MPC) (Seborg et al., 2010), with the multivariable control sys- UC. At this location, an intermediate vapor stream is drawn and
tem having 12 manipulated variables and 35 controlled variables. fed to the argon column (ARC). The ARC operates at a very high
As previously noted, many previous studies (Cao et al., reflux ratio, and the “crude” argon stream removed at the top of
2011, 2016a; Dias et al., 2018; Pattison et al., 2016; Schäfer the ARC has a very small flow rate. High purity argon is then ob-
et al., 2019a) focusing on ASU scheduling involve a simulated tained in a separate device that is not considered here.
plant that produces only nitrogen. Multi-product ASUs have in
some instances been considered, but using reduced-order mod-
els (Cao et al., 2016b) or assuming quasi-stationary operating 2.2. Scheduling under dynamic constraints
points/modes (Misra et al., 2017; Zhang et al., 2016) – in essence,
another dimensionality-reduction device. One should recognize Historically, production scheduling calculations relied on the
that such model reduction measures are computationally efficient, simplifying assumption that the transition of a chemical process
and have even allowed extending production scheduling calcula- from one operating point to another can be defined in terms of
tions to sites that are served by several multi-product ASUs (Zhao a transition period, and that the process predominantly operates
et al., 2019; Zhou et al., 2017); however, to our knowledge, there at steady states corresponding to the aforementioned operating
are no results available in the open literature on the explicit incor- points. Under the assumption that the durations of the transi-
poration of process dynamics in scheduling calculations for multi- tions periods are known, are short compared to periods of steady-
product ASUs. state operation, and do not change in time, scheduling calculations
The flowsheet of the process considered in this work is shown can effectively ignore the dynamics of the process, and schedul-
in Fig. 1. Air enters the feed compressor (FC), which is driven by ing decisions can be separated from control/operational decisions
a large electric motor. Impurities such as water and hydrocarbons (Seborg et al., 2010).
are removed in the prepurifier (PP). The air stream is then cooled Recent works (Baldea and Harjunkoski, 2014; Dias et al., 2018)
in the multistream heat exchanger (MHEX) against warming cryo- have demonstrated that these assumptions may no longer be valid
genic product streams and enters the high-pressure lower column when market conditions change frequently. For DR scheduling of
(LC), where it is separated into nearly pure nitrogen “shelf gas” (at ASUs, this is particularly true given that electricity prices change at
the top of the LC) and an oxygen-rich bottoms stream. A portion frequencies similar to (or faster than) the slowest dynamic modes
of the shelf gas provides the reboiler duty to the low-pressure up- of the process. Fig. 2 illustrates a hypothetical example, where, in
per column (UC) before being returned to the LC as reflux, while schedule 1, a process violates a dynamic constraint for a particu-
the rest of the shelf gas passes through the MHEX and is com- lar state variable owing to operators imposing a large step change
bined with the nitrogen product of the UC at the liquefier (LQ). in operation too quickly. In schedule 2, the production schedule
The oxygen-rich bottoms stream of the LC passes through the ni- accounts for the dynamic agility of the process when making the
trogen superheater (SH) before being used as a condensing utility step change(s). Note the transition times are different in schedules
for the argon column (ARC) and passed to the UC. 1 and 2, and accounting for process dynamics may lead to “slower”
A portion of the combined nitrogen streams is liquefied in the (i.e., with less frequent/abrupt changes), but dynamically feasible
liquefier (LQ), which provides both gaseous and liquid nitrogen schedules. The production scheduling problem considering process
products. A “waste” stream of nitrogen is drawn at an interme- dynamics can generally be stated as a dynamic optimization prob-
diate stage close to the top of the UC in order to provide addi- lem (Pattison et al., 2016):
C. Tsay, A. Kumar and J. Flores-Cerrillo et al. / Computers and Chemical Engineering 126 (2019) 22–34 25

Fig. 2. Left: Hypothetical process schedule that violates a process constraint while tracking a step change imposed in a scheduling-relevant variable. Right: Hypothetical
process schedule with constraints correctly enforced while making the desired transition.

dynamics of a process, which can be used in scheduling calcula-


 t=t f
max t=0
(x, t )dt (1) tions (Du et al., 2015). The SBM scheduling framework is illustrated
u(t ) conceptually in Fig. 3, along with a comparison with the tradi-
s.t. x˙ = f (x, u, θ ) (2) tionally separate scheduling and control layers. In the latter, the
process operates under closed-loop control, while the scheduling
xL ≤ x ≤ xU (3) layer provides the targets given to the supervisory controller. Solv-
ing a dynamic optimization problem in this paradigm using a dy-
Storage system and product demand constraints (4) namic model (which can be low order (Dias et al., 2018)) requires
a closed-loop simulation to compute the process performance for
a given schedule. In contrast, the SBM approach (Pattison et al.,
where u(t) are the time-varying production targets and/or other 2016) directly models the closed-loop process dynamics. Although
set points to be fed to the control system, x are the process state re-scheduling is not considered in this work, the process states at
variables with lower bounds xL and upper bounds xU , and θ are the end of the scheduling time horizon in either approach could be
the process parameters. The function f(x, u, θ ) is the dynamic pro- fed back to the scheduling layer to naturally enable re-scheduling
cess model and the bounds xL and xU are enforced at all times calculations (Pattison et al., 2017).
on the process state variables x during the scheduling horizon There are two broad approaches to deriving low-order dynamic
t ∈ [0, tf ]. The economic objective function (x, t) includes the rev- models: model reduction and system identification. Model reduc-
enue from selling product and the process operating costs. Storage tion refers to the derivation of a low-order model from a more de-
and demand constraints are typically enforced to ensure customer tailed dynamic process model. This can be performed using, e.g.,
demand can be met at all times and that the amount of product singular perturbation arguments (Baldea and Daoutidis, 2012) for
stored does not exceed the physical capacity of the storage sys- system models exhibiting multiple time scales. Alternative tech-
tem. A terminal constraint or penalty may also be added on some niques such as empirical Gramians (Hahn and Edgar, 2002) and
variables to prevent artificial economic gains from selling product proper orthogonal decomposition (Willcox and Peraire, 2002) can
already present in inventory at t = 0. also be used, although these methods can produce models with
While formulating the scheduling problem as a dynamic opti- states that are not physically meaningful. The second approach,
mization allows for enforcing path constraints, the dynamic op- system identification, refers to “learning” a low-order process model
timization is difficult to solve quickly for use in practical situa- from operating data; many techniques are available. We direct the
tions. Recent works have highlighted the importance of online re- interested reader to, e.g., the book by Zhu (2001) for an overview
scheduling (Beal et al., 2017; Gupta et al., 2016), further empha- of system identification techniques and their applications to pro-
sizing the desire for computational expedience. Given the above, cess systems. Learning process dynamics from data is a highly ac-
the process dynamic model, embedded in the dynamic optimiza- tive area of research (Klus et al., 2018) and is the approach we take
tion problem as f(x, u, θ ), should preferably be low-dimensional in this work.
(Baldea et al., 2015). Replacing detailed first-principles models with
low-order models, termed “scale-bridging models” or SBMs, for
scheduling applications has been the subject of several recent 3. Constructing data-driven scale-bridging models of the
works: the reader is referred to (Baldea et al., 2015; Pattison et al., industrial ASU
2017, 2016; Tsay et al., 2018a) for a detailed discussion. In the fol-
lowing section, we describe the development of SBMs relevant to 3.1. System identification framework
the industrial problem considered here.
Chemical processes typically have many sensors that record
2.3. Scheduling-relevant scale-bridging models measurements at frequencies in the order of minutes, generating
“big data” sets that can be exploited to understand the process dy-
In this work, we exploit the data-driven scale-bridging model namics. Pattison et al. (2016) observed that the set of scheduling-
scheduling framework proposed by Pattison et al. (2016) to sched- relevant variables can be a small subset of the process variables,
ule a large-scale industrial process under MPC. We identify SBMs reducing the dimensionality of the scheduling problem. In particu-
using historical operating data that reflect routine operations. In lar, they suggested that only (i) variables relevant to the scheduling
the course of the time interval spanned by the data, process op- objective function, demand constraints, and inventory constraints
erators imposed some changes to the operation of the plant, but and (ii) variables that are near their bounds during historical static
the dataset lacks any deliberate excitation that could be construed or transient operation should be modeled and included in schedul-
as being part of a system identification experiment. An SBM is an ing optimization calculations. The remaining variables are assumed
explicit, low-order representation of the closed-loop input-output to not be relevant to optimal scheduling calculations, as historical
26 C. Tsay, A. Kumar and J. Flores-Cerrillo et al. / Computers and Chemical Engineering 126 (2019) 22–34

Fig. 3. A comparison between scheduling using detailed closed-loop process models (A) and scheduling with scale-bridging models (Baldea et al., 2015; Pattison et al., 2016)
(B). The dashed box in (A) reflects the system whose dynamics are represented by the SBMs.

data suggest that they do not hinder process agility. Briefly, the ap- scheduled to avoid consuming electricity during price peaks, they
proach for learning SBMs comprises the following steps: were not considered in this study. However, the shutdown of the
liquefier was considered, as described later.
1. Obtain historical process operating data
The data correspond to summer operation, as the summer
2. Identify scheduling-relevant variables to be modeled
months tend to have larger fluctuations in electricity prices. The
3. Determine model form and fit a dynamic model for each
first 88 days (80%) of the data were selected as training data, with
scheduling-relevant variable
the remaining 22 days (20%) used as test data to evaluate the qual-
In the application of the above system identification framework ity of the identified SBMs. While the choice of the functional form
to the industrial ASU considered here, several important factors of data-driven model can be arbitrary, due to the lack of deliberate
had to be considered. On the input side of the models, while the system excitation, we chose to identify SBMs in the Hammerstein-
process operates under MPC with six operator set points (targets), Wiener (HW) form. This choice was supported by successful ap-
we found through trial-and-error that the production rates and plication of HW models to capture the relevant process dynamics
other scheduling-relevant state and output variables could be ac- in a smaller-scale problem focused on DR scheduling of a single-
curately modeled using only two of the six operator set points, product ASU (Pattison et al., 2016). HW models are structured, re-
[SP1, SP2]. This finding suggests that the remaining four operator quiring fewer parameters to be regressed. They comprise a linear
set points either have negligible impact on the variables of inter- dynamic component flanked by static, nonlinear input and output
est for scheduling, or that they have historically been set such that transformations. They can be formulated for multiple inputs and
they are strongly correlated with the first two set points. On the a single output (MISO) with the linear dynamic component repre-
other hand, we found that ambient temperature T, which is a mea- sented as a state-space model:
sured disturbance variable, has a large effect on the process vari-
ables and must be included as an input to the SBMs. Given these hi = Hi (ui ) (5)
findings, we identified multiple-input, single-output (MISO) SBMs.
The SBM used to predict the behavior of each scheduling-relevant z˙ i = Aizi + Bi hi (6)
variable was modeled with three inputs, [T, SP1, SP2].
While in our previous works (Dias et al., 2018; Pattison et al., yi = Cizi (7)
2016) we relied on relatively small simulated datasets (whose hori-  
zons span a few days to a week) for system identification, in this w=W yi (8)
work we use a large set of actual process historical operating data.
Since in the present application the data lack the deliberate ex- where H and W are, respectively, the Hammerstein and Wiener
citation imposed during system identification experiments (as in blocks corresponding to the static, nonlinear input and output
our previous work), we included 110 days of historical operating transformations. Ai , Bi , and Ci are the matrices defining the linear
data from the industrial process to ensure that sufficiently rich in- state-space dynamical system, ui denotes the ith model input (for
formation is available for identifying the desired SBMs. The data the ASU, ui ∈ [T, SP1, SP2]), and w is the model output. The linear
were recorded at one-minute intervals in the process historian state-space system is of order nd,i , with zi ∈ IRnd,i . The structure of
database, during periods of regular, unforced operation. Periods of a generic MISO Hammerstein–Wiener model with three inputs and
start-up, shut-down, and process or measurement faults were ex- one output is shown in Fig. 4.
cluded from this study. These periods were identified easily in the Although the industrial ASU produces five products, we found
historical data, as the sensors are either off or produce readings that the liquid oxygen production rate was relatively constant
that are, e.g., outside the physical bounds for the respective vari- throughout the operating period represented in the data. As ac-
ables. Although shut-downs of the entire ASU could potentially be tual industrial data were used in this study, the argon production
C. Tsay, A. Kumar and J. Flores-Cerrillo et al. / Computers and Chemical Engineering 126 (2019) 22–34 27

Fig. 4. MISO Hammerstein–Wiener model with three inputs.

Table 1
Details of Hammerstein–Wiener scale-bridging models (refer to Fig. 4 for the model struc-
ture).

Output np for inputs nd for inputs np for output Training Test


variable [T, SP1, SP2] [T, SP1, SP2] NMSE (%) NMSE (%)

PLN2 [2,1,3] [1,3,2] 5 83.04 74.53


PGN2 [4,3,3] [3,2,1] 5 76.17 61.52
PGO2 [3,3,2] [4,3,3] 3 66.39 65.55
Power [4,3,3] [3,3,1] 5 82.18 82.67
CV1 [2,2,2] [3,3,3] 4 67.53 61.86
CV2 [4,4,2] [2,3,2] 4 42.22 42.65

rate was not disclosed and modeled for confidentiality reasons. The the connection of adjacent line segments: bpi is the value of
production rates that are therefore considered in DR scheduling the breakpoint at segment i, and pwi is the value of the non-
calculations (and whose dynamic responses to changes in [T, SP1, linear function at the breakpoint bpi . To determine the num-
SP2] are modeled) are thus those of liquid nitrogen (PLN 2 ), gaseous ber of piecewise-linear segments np for each transformation, the
nitrogen (PGN 2 ), and gaseous oxygen (PGO 2 ). The MPC implementa- normalized Akaike information criterion (nAIC) was minimized
tion includes constraints on several process variables, but we found while using a large number of segments for all other piecewise-
that only two controlled process variables, denoted here as CV1 linear transformations and a high-order linear state-space model.
and CV2, closely approach their bounds during operational transi- The order of each linear state-space model was similarly deter-
tions, indicating they are the primary factors limiting the dynamic mined using the nAIC. Table 1 also reports the normalized mean
agility of the ASU. In the interest of protecting industrial intellec- square error (NMSE) for each model on training and test data,
tual property, their true nature cannot be disclosed. Together with defined as:
the power consumption of the plant, the scheduling-relevant out-
 xref − x2
puts of the models are w ∈ [PLN2 , PGN2 , PGO2 , CV1, CV2, Power]. NMSE = 1 −   (11)
 xref − mean xref 2
3.2. System identification results
The SBM predictions for all six variables listed in Table 1 for
a week within the training dataset are plotted in Fig. 5. The
A MISO SBM in Hammerstein–Wiener form (5)–(8) was identi-
low NMSE values are largely due to high-frequency dynam-
fied for each scheduling-relevant variable, and the pertinent details
ics in the historical data: the MPC operates (and the data are
are shown in Table 1. The models were obtained using the Sys-
recorded) on a 1-min interval, while the operator set points are
tem Identification Toolbox in MATLAB. Kelley et al. (2018) found
changed much less frequently. The model predictions are close
that the dynamics of an ASU could be accurately represented us-
to the time-averaged behavior of the process variables in most
ing Hammerstein–Wiener models with piecewise linear input and
cases, and the accuracy of model predictions could most likely
output functions. Consequently, the nonlinearities were all repre-
be improved in future studies by collecting process (or simu-
sented as piecewise linear functions of the form:
lated) data with higher excitation levels such as through sys-
pw j+1 − pw j
Hi (ui ) = (ui − bp j ) + pw j ; ui ∈ [bp j , bp j+1 ) (9) tem identification experiments. Additional data may also help
bp j+1 − bp j improve model accuracy by affording the selection of a differ-
    ent model representation/structure. HW models are inherently
pwk+1 − pwk  restricted to a linear representation of the dynamics; although
W yi = yi − bpk + pwk ;
bpk+1 − bpk this feature does not pose limitations for this study, it may be-
 come limiting in cases where the dominant time constant of
yi ∈ [bpk , bpk+1 ) (10)
the process changes significantly over time or as a result of
where j = 1, . . . , n p + 1 is the set of piecewise linear segments changes in operating regime. We note that all output and in-
in an input nonlinear transformation; k is used to similarly in- put variables listed or plotted in this article have been scaled
dex the output nonlinear transformation. The piecewise linear and filtered to preserve the confidential nature of industrial
functions are each parameterized by np breakpoints that define data.
28 C. Tsay, A. Kumar and J. Flores-Cerrillo et al. / Computers and Chemical Engineering 126 (2019) 22–34

Fig. 5. Scale-bridging model predictions for one week of test data.

4. Optimal demand response operation of the ASU


CV 2L ≤ CV 2 ≤ CV 2U (15)
4.1. Problem formulation  t=t f
t=0
PLN2 (t )dt ≥ DLN2 (16)
The continuous-time ASU demand-response scheduling prob-  t=t f
lem was formulated as a dynamic optimization of the type (1)– t=0
PGN2 (t )dt ≥ DGN2 (17)
(4), using the identified SBMs to predict the transient behavior of  t=t f
relevant process variables. The objective is to minimize the cost t=0
PGO2 (t )dt ≥ DGO2 (18)
of electricity purchased from the grid over a four-day horizon by
manipulating the evolution in time of the two operator set points where Celec (t) is the time-varying price of electricity, fHW denote
SP1, SP2. Constraints are added such that product demand is met the identified Hammerstein-Wiener models, and Di is the (known)
and the scheduling-relevant CVs remain within their prescribed demand for product i over the four-day window. (14) and (15) are
bounds. The disturbance variable, ambient temperature, is also ac- path constraints on the scheduling-relevant CVs, and (16)–(18) are
counted for. For this initial study, we assume that, (i) production production constraints on the ASU products.
constraints are known and can be enforced on total production, (ii) A four-day period from the test data, where operator set point
the storage capacity is sufficiently large to retain products to meet values were constant in time, was identified as a base case. Ex-
demand for the four-day period considered, and (iii) perfect fore- tended periods of operation with constant operator set points in
casts of ambient temperature and electricity price are available for the historical data occurred only with the plant running at maxi-
the entire scheduling horizon. While these assumptions may not mum capacity (maximum here refers to the highest value present
hold true in practical applications, the solutions we obtain pro- in the available dataset). In order to simulate over-production
vide valuable insight into the maximum benefit that can be de- without extrapolating, the same four-day period was simulated for
rived from DR scheduling of industrial ASUs. The DR scheduling operation at 95% (in terms of power consumption) of the plant
optimization problem can be written as: capacity. Historical ambient temperature profiles were obtained
from the process historian, and historical electricity prices from
 t=t f the regional independent system operator (ISO). We note that, al-
min t=0
Celec (t )Power (t )dt (12) though the operator set points are held constant in the base case,
SP1(t ),SP2(t )
the process states are still affected by fluctuations of the distur-
s.t. x˙ HW = f HW (SP 1, SP 2, T ) (13) bance variable (temperature) in time. The real-time and day-ahead
electricity prices over the selected four-day horizon are shown
CV 1L ≤ CV 1 ≤ CV 1U (14) in Fig. 6. The real-time electricity prices are subject to larger
C. Tsay, A. Kumar and J. Flores-Cerrillo et al. / Computers and Chemical Engineering 126 (2019) 22–34 29

Fig. 6. Electricity prices for the selected four-day horizon.

variations, with a sharp peak of over $500/MWh in the second day,


but also some periods of negative electricity price. In contrast, the
day-ahead electricity prices follow a “smoother” periodic pattern.
The models were implemented in gPROMS (Process Systems
Enterprise, 1997–2018), and the optimal scheduling calculations
were performed therein using a sequential dynamic optimization
solver. An alternative approach could involve full discretization of
the time domain and solution of the resulting large-scale MILP
(Kelley et al., 2018). The schedules reported herein represent the
local optima found using the base case as the initial point. The
calculations were performed on a 64-bit Windows system with an
Intel Core i7-8700 CPU at 3.20 GHz and 16GB RAM. The optimiza-
tion computations required about 600 s each, making the method
amenable to online scheduling and re-scheduling calculations. The
reader is referred to the papers by Pattison et al. (2017, 2016) for
details on the formulation of the optimal scheduling problem un-
der dynamic constraints represented as HW models, and to the Fig. 7. Power consumption predicted by SBMs for the optimal schedules in the real-
work of Vassiliadis et al. (1994a,b) for an in-depth discussion on time market using the MPC bounds (P1, top) and backoff bounds (P2, bottom).
sequential dynamic optimization.

4.2. Results for the case of real-time electricity pricing 1994b), using two-hour piecewise constant profiles. As expected,
the optimal schedules found for both P1 and P2 involve over-
The optimization problem in (12)–(18) was solved using the producing (and increasing power consumption with respect to the
prices Celec (t) given by the historical real-time electricity price base case) when electricity prices are low. Conversely, power con-
shown in Fig. 6. As their name implies, these prices are set in sumption is decreased when electricity prices are high. The power
real time and, as a consequence, their values for a 96-h future demand profiles for P1 and P2 are shown in Fig. 7. Although the
time horizon would not be available in practice at the time when identified SBMs are (piecewise) continuous, a visual comparison
scheduling calculations are performed. Rather, predictions would of the results against historical data suggests that the sharp drop-
need to be used to generate the optimal schedule. The bounds de- off in power consumption between hours 40–45 in both computed
fined in the MPC system for CV1 and CV2 were used as their re- schedules corresponds to a liquefier shutdown, where power con-
spective upper and lower bounds in (14)–(15). We will refer to this sumption is drastically decreased at the cost of producing no LN2
problem as P1. The same problem was solved using a 10% backoff (PLN2 ≈ 0).
constraint for both CV1 and CV2 to generate a more conservative The LN2 production profiles for the computed optimal sched-
schedule, noting that the SBMs representing process dynamics may ules are shown in Fig. 8. In the optimal solution for both P1 and P2,
not be completely accurate (Table 1). Such “backoffs” for active only the LN2 production constraint (16) is exactly met, as the other
constraints have been used for SBM-based scheduling (Dias et al., two products are over-produced throughout the four-day window
2018; Pattison et al., 2016), and are in effect implemented in many to allow for excess LN2 production. LN2 production is decreased to
practical situations to avoid constraint violations in the presence of practically zero during the liquefier shutdown episodes, as can be
disturbances and/or model inaccuracy. We direct the reader to the easily seen in Fig. 8. We again note that the variables are scaled
works by Narraway and Perkins (1993) and Aske et al. (2008) for and that a value of PLN2 = 0 actually corresponds to a “negative”
further information. We will refer to the second problem, which net LN2 production rate, i.e., LN2 is transferred from storage to the
includes the aforementioned backoff constraints, as P2. For each UC (Fig. 1) when the liquefier is turned off. The profiles of CV1
variable, the 10% backoff was computed by adding/subtracting 10% at the optimal points are also shown in Fig. 8; CV2 did not reach
of the scaled variable range to/from its lower/upper bound. its bounds in the optimal schedules. It can be seen that using the
Optimization of the operator set points was treated using a tighter backoff constraint results in CV1 reaching its lower bound
control vector parameterization approach (Vassiliadis et al., 1994a; at multiple points in the solution of P2. The backoff constraint
30 C. Tsay, A. Kumar and J. Flores-Cerrillo et al. / Computers and Chemical Engineering 126 (2019) 22–34

Fig. 9. Real-time (top) and day-ahead (bottom) electricity prices under uncertainty.
The shaded area denotes one standard deviation for i ∼ N(0, 50% ). Five Monte
Carlo (dotted) samples are shown.

a cryogenic liquid storage capacity of 1.92 times the hourly aver-


age production, while the solution for P2 requires a tank capacity
of 1.68 times hourly average production. Industrial ASUs typically
have ample storage capacity available, and we do not expect stor-
age of cryogenic liquids to become a limitation in practical imple-
mentations of this DR strategy.
A sensitivity analysis can be conducted to explore the impact of
imperfect price forecasts. Denoting the electricity price prediction
for hour j as Pˆj , ∀ j = 1, . . . , 96, we define the relative error of the
electricity price forecast as:

Pˆj − Pj
j = (19)
Pj

where Pj is the “true” electricity price in hour j. For the purpose


Fig. 8. Profiles of PLN2 and CV1 predicted by SBMs for the computed optimal sched- of this study, we assume that  j are independent and identically
ules in the real-time market. distributed, although, e.g., Gaussian process models (Kou et al.,
2015) could give smoother electricity-price prediction profiles by
accounting for hour-to-hour error correlations. Fig. 9 (top) shows
limits liquefier shutdowns compared to the case with no backoff, the electricity prices for the four-day horizon with  j ∼ N(0, 50% ),
and thus the economic savings are decreased. This demonstrates where N(μ, σ ) denotes the normal distribution with mean μ and
that the SBM-based scheduling approach is capable of identifying standard deviation σ . The relative error representation for real-
schedules that are dynamically feasible, with respect to the given time pricing is optimistic, as the occurrence and timing of the
process and variable bounds. peak(s) and period(s) of negative price are still predicted properly
We define CRT 0 as the incurred cost for the four-day operation
(in terms of expected value).
period in the base case given the real-time electricity prices. The To explore the effect of price uncertainty on schedule optimal-
optimal schedule found by solving P1 results in a 13.1% savings ity, Monte Carlo simulations were performed with  j ∼ N(0, σ ).
compared to CRT 0 , while the schedule found by solving P2 results
Each Monte Carlo simulation was run with 50 0 0 samples, and
in a 6.8% savings. While the savings decrease considerably when uncertainty was coarse-grained into 4-h blocks to reduce the di-
the dynamic agility of the process is used conservatively (con- Pˆj −P j
mensionality of the uncertainty space (i = P j , {i = 1, . . . , 24; j =
strained with the 10% backoff), both savings values represent a sig-
nificant economic advantage in the commoditized industrial gas 4i − 3, . . . , 4i}). The first five samples for σ = 50% are shown in
sector (Tsay et al., 2018b). These large savings are due to perfect Fig. 9. Generating a Monte Carlo sample involves in fact two sim-
predictions (including of price spikes and negative prices), perfect ulations: one with the base case set points (which creates the
temperature forecasts, and no constraint(s) on the amount of ma- basis for comparison) and one with the optimal schedule. The
terial that can be stored. If large spikes in electricity price are pre- output datum from each sample simulation is the ratio of the
dicted correctly, the ASU can shut down its liquefier and sell LN2 cost incurred by implementing the optimal schedule, denoted as
over-produced at other times to exploit its ability to store prod- CRT,m ; m = 1, . . . , 50 0 0, to the cost associated with executing the
base case schedule, denoted as CRT,m 0 ; m = 1, . . . , 50 0 0.
uct (effectively storing energy in the form of molecules). Although
no storage constraints were modeled, the required storage tank 0
The sampled values of CRT,m and CRT,m are approximately nor-
capacity was estimated by examining the maximum over/under- mally distributed, as the uncertain variable (electricity price) af-
production of LN2. Using this estimate, the solution for P1 requires fects the variable of interest (plant operating cost) linearly. The
C. Tsay, A. Kumar and J. Flores-Cerrillo et al. / Computers and Chemical Engineering 126 (2019) 22–34 31

Fig. 11. Power consumption predicted by SBMs for the optimal schedules in the
day-ahead market using the MPC bounds (P3, top) and backoff bounds (P4, bottom).

Fig. 10. Relative costs computed in Monte Carlo simulations for P1 (top) and P2 right, indicating that although the chances of exceeding constant-
(bottom) using σ = {10%, 30%, 50%}. Skewed Gaussian probability density functions
are shown as dashed lines.
operation costs may be low, the probabilities of encountering ex-
treme values are greater on the right-hand side of the distribution.

CRT,m 4.3. Results for the case of day-ahead electricity pricing


stochastic relative cost of the optimal schedule, 0 is thus the
CRT,m
ratio of two (approximately) normal distributions. The closed-form The same optimization problem given in (12)–(18) was solved
expression of the ratio distribution is complicated (Hinkley, 1969), using the prices Celec (t) given by the historical day-ahead electricity
and we approximate it using a skewed Gaussian distribution to cal- price at the same location (as shown in Fig. 6). These prices have
culate relevant statistics, which are provided in full in the Sup- a smaller range of fluctuation and are relatively more predictable,
porting Information. The approximation is accurate for low val- but are still subject to some uncertainty as they are, in fact, only
ues of σ , but the quality of the approximation deteriorates no- known a single day (rather than 96 h) in advance. The scheduling
ticeably for σ ≥ 80%. Histograms of the relative costs found in the optimization problem was again solved with the bounds specified
Monte Carlo simulations for P1 and P2 using σ = {10%, 30%, 50%} in the process MPC as the control variable bounds (P3) and with a
are shown in Fig. 10 (note the different abscissa ranges). There 10% backoff constraint (P4). The optimal schedules for the ASU in-
is a non-negligible probability that the relative cost exceeds 1 volve shutting down the liquefier once in each case, corresponding
when uncertainty is considered for electricity prices, suggesting to times of high electricity prices. The power consumption profiles
that demand-response operation may not always be economically for both optimal schedules are shown in Fig. 11. The liquefier shut-
favorable when the accuracy of price forecasts is low. The probabil- down in the solution to P3 occurs during the third electricity price
ities of relative cost exceeding 1 are plotted in Fig. 14, and the rel- peak, even though the electricity price is marginally higher in the
evant values and statistics can be found in the Supporting Informa- fourth peak, likely reflecting a locally optimal solution.
tion. Even with the optimistic assumption that the expected value Due to the scheduled liquefier shutdown periods, LN2 is again
of electricity price is the predicted value (i.e., times of electricity the limiting product in the optimal schedules, and the LN2 pro-
price peaks and negative electricity prices are properly identified), duction profiles for the optimal schedules are shown in Fig. 12.
the optimal schedule(s) may be economically worse than constant The cost savings are primarily due to shutting down the liquefier
set point operation. The distributions shown in Fig. 10 are skewed during times of peak electricity price and over-producing liquid
32 C. Tsay, A. Kumar and J. Flores-Cerrillo et al. / Computers and Chemical Engineering 126 (2019) 22–34

Fig. 13. Relative costs found in Monte Carlo simulations for P3 (top) and P4 (bot-
Fig. 12. Profiles of PLN2 and CV1 predicted by SBMs for the computed optimal tom) using σ = {10%, 30%, 50%}. Skewed Gaussian probability density functions are
schedules in the day-ahead market. shown as dashed lines.

larger than the corresponding values found for the real-time elec-
nitrogen at times of lower electricity price. We note that this is tricity price cases (P1 and P2).
likely strongly related to the large number of liquefier shutdown We again explore the effect of price uncertainty on the sched-
events recorded in the training data set. While this inherently bi- ules using the same Monte Carlo simulation with i ∼ N(0, σ ), but
nary decision is handled using the HW model, a discrete variable with no uncertainty in the first 24 h (where day-ahead prices
could also model the on/off state of the liquefier. The profiles of would be exactly known). The first five samples for σ = 50%
CV1 at the optimal points are also shown in Fig. 12; CV2 is again are shown in Fig. 9 (bottom). Histograms of the relative costs
not found to reach its bounds in the optimal schedules. The backoff found in the Monte Carlo simulations for P3 and P4 using σ =
constraint is met multiple times in the solution to P4. {10%, 30%, 50%} are shown in Fig. 13. The same abscissae as in
We again define CDA 0 as the incurred cost of four-day operation Fig. 10 were used in Fig. 13 to facilitate comparison between re-
in the base case given day-ahead electricity prices. The optimal sults for the real-time and day-ahead markets. A comparison be-
schedule found by solving P3 results in a 3.6% savings compared tween Figs. 10 and 13 reveals that, although the expected relative
0 , while the schedule given by the solution of P4 results in a
to CDA cost of DR operation in the day-ahead market is higher (demand
2.4% savings. Although lower than in the case of the real-time mar- response in the real-time market yields larger savings in the case
ket, these savings still constitute a significant economic advantage of no error in the expected values of electricity prices), the sched-
in the industrial gas sector. The economic benefits are lower when ules based on day-ahead data are less affected by price uncertainty,
the backoff constraint is used, as the tighter bound for CV1 gives a and the distributions of relative costs found for the day-ahead mar-
more conservative measure to the degree to which the ASU power ket have significantly less variance. There is still a small probability
consumption can be modulated (Fig. 12). Using the same storage that the relative cost exceeds 1. The relevant statistical data can be
capacity estimation method as above, the solution for P3 requires found in full in the Supporting information.
a liquid storage capacity of 2.62 times the hourly average produc- The probabilities of relative costs exceeding 1, estimated from
tion, while the solution for P4 requires a capacity of 1.95 times Monte Carlo frequencies for all four schedules, are shown in
the average hourly production. These storage capacities are slightly Fig. 14. The mean relative costs and standard deviations for all four
C. Tsay, A. Kumar and J. Flores-Cerrillo et al. / Computers and Chemical Engineering 126 (2019) 22–34 33

E[Pj ]; ∀ j = 1, . . . , 96, scheduling in the real-time market produces


more savings. This assumption corresponds to knowing the exact
timing of extreme electricity prices (in terms of expected value),
which may not be valid in the case of real-time electricity markets.
Using the backoff constraint (P2 and P4) results in more conserva-
tive schedules, and consequently the standard deviations in relative
cost (and probabilities of relative costs exceeding 1) are decreased.

5. Conclusions

Cryogenic air separation units (ASUs) are a prime candidate for


DR operation, as they consume large amounts of electricity, but
their products are agnostic to time of production. However, opti-
mal scheduling calculations for industrial ASUs must account for
process dynamics in order to ensure that product quality and pro-
cess safety constraints are met at all times during the execution
of the schedule. In this work, we apply a data-driven approach
to learn the scheduling-relevant dynamics of an industrial process
from its operating history without providing any deliberate ex-
citation into the input variables. We use data from operation in
the summer months, when electricity price fluctuations are typi-
Fig. 14. Effect of increasing price uncertainty on the probability of exceeding
cally the highest, to create data-driven, low-order dynamic mod-
constant-operation cost. Probabilities are estimated by frequency in 50 0 0 Monte
Carlo samples.
els (scale-bridging models) of the scheduling-relevant dynamics.
With these models, the scheduling calculation is formulated as a
dynamic optimization problem that is easily tractable from a com-
putational point of view.
We demonstrate the value of the proposed scheduling frame-
work by comparing the demand response operation of the indus-
trial ASU against a constant operation base case, finding that over
13% cost savings over a four-day horizon can be achieved in real-
time electricity markets. In the more conservative day-ahead elec-
tricity market, we still find that over 3% cost savings are possi-
ble over the four-day horizon. These cost savings estimates may
still suffer from inaccuracy due to assumptions of perfect price and
temperature forecasts, but the savings are nevertheless extremely
significant in the commoditized industrial gas market.

Disclaimer

This report was prepared as an account of work sponsored by


an agency of the United States Government. Neither the United
States Government nor any agency thereof, nor any of their em-
ployees, makes any warranty, express or implied, or assumes any
legal liability or responsibility for the accuracy, completeness, or
Fig. 15. Effect of increasing price uncertainty on relative costs of production de- usefulness of any information, apparatus, product, or process dis-
termined using 50 0 0 Monte Carlo samples. The marked lines represent the mean closed, or represents that its use would not infringe privately
relative costs, while the shaded regions represent one standard deviation. We note owned rights. Reference herein to any specific commercial product,
that the distributions are in fact skewed (see Supporting information).
process, or service by trade name, trademark, manufacturer, or oth-
erwise does not necessarily constitute or imply its endorsement,
schedules are shown in Fig. 15. We note that while the mean val- recommendation, or favoring by the United States Government or
ues of relative costs increase with electricity price uncertainty for any agency thereof. The views and opinions of authors expressed
DR operation in the real-time market (P1 and P2), they are rel- herein do not necessarily state or reflect those of the United States
atively flat for the schedules created with day-ahead prices (P3 Government or any agency thereof.
and P4). In the cases with no backoff constraint (P1 and P3), the
probabilities of relative cost exceeding 1 are almost identical be- Acknowledgments
tween day-ahead and real-time markets for all values of uncer-
tainty. However, Fig. 15 shows that DR scheduling in the real-time The authors thank the National Science Foundation (NSF) for
market increases the chances of extreme values. For the more con- funding through CAREER Award 1454433 and Award CBET-1512379.
servative schedules generated with a 10% backoff constraint (P2 This material is also based on work supported by the US De-
and P4), the probabilities of relative cost exceeding 1 are greater partment of Energy under Award Number DE-OE0 0 0 0841. The au-
in the real-time market, and extreme values are again more likely thors also gratefully acknowledge the financial support of Prax-
in the real-time market. These observations confirm the intuition air, Inc. The authors thank Drs. Jun Shi and Richard C. Patti-
that scheduling in the real-time market is riskier for the same rel- son for their guidance in the early stages of this work, as well
ative uncertainty in electricity prices, although the expected val- as the CAST division directors for their constructive comments
ues of savings are larger. With the given assumption that E[Pˆj ] = at the AIChE 2018 Annual Meeting. CT was supported by the
34 C. Tsay, A. Kumar and J. Flores-Cerrillo et al. / Computers and Chemical Engineering 126 (2019) 22–34

University Graduate Continuing Fellowship at The University of Klus, S., Nüske, F., Koltai, P., Wu, H., Kevrekidis, I., Schütte, C., Noé, F., 2018. Data–
Texas at Austin. driven model reduction and transfer operator approximation. J. Nonlinear Sci.
28 (3), 985–1010.
Kou, P., Liang, D., Gao, L., Lou, J., 2015. Probabilistic electricity price forecasting with
Supplementary material variational heteroscedastic Gaussian process and active learning. Energy Con-
vers. Manage. 89, 298–308.
Misra, S., Kapadi, M., Gudi, R.D., Srihari, R., 2017. Energy-efficient production
Supplementary material associated with this article can be scheduling of a cryogenic air separation plant. Ind. Eng. Chem. Res. 56 (15),
found, in the online version, at doi:10.1016/j.compchemeng.2019. 4399–4414.
Mitra, S., Grossmann, I.E., Pinto, J.M., Arora, N., 2012. Optimal production planning
03.022. under time-sensitive electricity prices for continuous power-intensive processes.
Comput. Chem. Eng. 38, 171–184.
References Mitra, S., Pinto, J.M., Grossmann, I.E., 2014. Optimal multi-scale capacity planning
for power-intensive continuous processes under time-sensitive electricity prices
and demand uncertainty. part i: modeling. Comput. Chem. Eng. 65, 89–101.
Aske, E.M., Strand, S., Skogestad, S., 2008. Coordinator MPC for maximizing plant
Narraway, L.T., Perkins, J.D., 1993. Selection of process control structure based on
throughput. Comput. Chem. Eng. 32 (1–2), 195–204.
linear dynamic economics. Ind. Eng. Chem. Res. 32 (11), 2681–2692.
Baldea, M., Daoutidis, P., 2012. Dynamics and Nonlinear Control of Integrated Pro-
Nyström, R.H., Franke, R., Harjunkoski, I., Kroll, A., 2005. Production campaign plan-
cess Systems. Cambridge University Press, Cambridge, UK.
ning including grade transition sequencing and dynamic optimization. Comput.
Baldea, M., Du, J., Park, J., Harjunkoski, I., 2015. Integrated production schedul-
Chem. Eng. 29 (10), 2163–2179.
ing and model predictive control of continuous processes. AlChE J. 61 (12),
Obermeier, A., Windmeier, C., Esche, E., Repke, J.-U., 2019. A discrete-time schedul-
4179–4190.
ing model for power-intensive processes taking fatigue of equipment into con-
Baldea, M., Harjunkoski, I., 2014. Integrated production scheduling and process con-
sideration. Chem. Eng. Sci. 195, 904–920.
trol: a systematic review. Comput. Chem. Eng. 71, 377–390.
Otashu, J.I., Baldea, M., 2018. Grid-level battery operation of chemical processes and
Basán, N.P., Grossmann, I.E., Gopalakrishnan, A., Lotero, I., Méndez, C.A., 2018. Novel
demand-side participation in short-term electricity markets. Appl. Energy 220,
milp scheduling model for power-intensive processes under time-sensitive elec-
562–575.
tricity prices. Ind. Eng. Chem. Res. 57 (5), 1581–1592.
Pattison, R.C., Touretzky, C.R., Harjunkoski, I., Baldea, M., 2017. Moving horizon
Beal, L., Petersen, D., Pila, G., Davis, B., Warnick, S., Hedengren, J., 2017. Economic
closed-loop production scheduling using dynamic process models. AlChE J. 63
benefit from progressive integration of scheduling and control for continuous
(2), 639–651.
chemical processes. Processes 5 (4), 84.
Pattison, R.C., Touretzky, C.R., Johansson, T., Harjunkoski, I., Baldea, M., 2016. Optimal
Beal, L.D., Petersen, D., Grimsman, D., Warnick, S., Hedengren, J.D., 2018. Integrated
process operations in fast-changing electricity markets: framework for schedul-
scheduling and control in discrete-time with dynamic parameters and con-
ing with low-order dynamic models and an air separation application. Ind. Eng.
straints. Comput. Chem. Eng. 115, 361–376.
Chem. Res. 55 (16), 4562–4584.
Burnak, B., Katz, J., Diangelakis, N.A., Pistikopoulos, E.N., 2018. Simultaneous process
Process Systems Enterprise, 1997–2018. general PROcess Modeling System
scheduling and control: a multiparametric programming-based approach. Ind.
(gPROMS). http://www.psenterprise.com/gproms.
Eng. Chem. Res. 57 (11), 3963–3976.
Schäfer, P., Caspari, A., Kleinhans, K., Mhamdi, A., Mitsos, A., 2019a. Reduced dy-
Cao, Y., Swartz, C.L., Baldea, M., 2011. Design for dynamic performance: application
namic modeling approach for rectification columns based on compartmental-
to an air separation unit. In: Proceedings of the American Control Conference
ization and artificial neural networks. AlChE J. doi:10.1002/aic.16568.
(ACC), 2011. IEEE, pp. 2683–2688.
Schäfer, P., Westerholt, H.G., Schweidtmann, A.M., Ilieva, S., Mitsos, A., 2019b. Mod-
Cao, Y., Swartz, C.L., Flores-Cerrillo, J., 2016a. Optimal dynamic operation of a high-
el-based bidding strategies on the primary balancing market for energy-intense
purity air separation plant under varying market conditions. Ind. Eng. Chem.
processes. Comput. Chem. Eng. 120, 4–14.
Res. 55 (37), 9956–9970.
Seborg, D.E., Mellichamp, D.A., Edgar, T.F., Doyle III, F.J., 2010. Process Dynamics and
Cao, Y., Swartz, C.L., Flores-Cerrillo, J., Ma, J., 2016b. Dynamic modeling and collo-
Control. John Wiley & Sons, Hoboken, NJ.
cation-based model reduction of cryogenic air separation units. AlChE J. 62 (5),
Tsay, C., Baldea, M., Shi, J., Kumar, A., Flores-Cerrillo, J., 2018a. Data-driven mod-
1602–1615.
els and algorithms for demand response scheduling of air separation units. In:
Caspari, A., Faust, J.M., Schäfer, P., Mhamdi, A., Mitsos, A., 2018. Economic nonlinear
Towler, G.P., Eden, M.R., Ierapetritou, M.G. (Eds.), Computer Aided Chemical En-
model predictive control for flexible operation of air separation units. IFAC-Pa-
gineering, Proceedings of the Thirteenth International Symposium on Process
persOnLine 51 (20), 295–300.
Systems Engineering, San Diego, CA. Elsevier, pp. 1273–1278.
Chu, Y., You, F., 2012. Integration of scheduling and control with online closed-loop
Tsay, C., Pattison, R.C., Piana, M.R., Baldea, M., 2018b. A survey of optimal process
implementation: fast computational strategy and large-scale global optimization
design capabilities and practices in the chemical and petrochemical industries.
algorithm. Comput. Chem. Eng. 47, 248–268.
Comput. Chem. Eng. 112, 180–189.
Conejo, A.J., Nogales, F.J., Arroyo, J.M., 2002. Price-taker bidding strategy under price
Vassiliadis, V.S., Sargent, R.W., Pantelides, C.C., 1994a. Solution of a class of mul-
uncertainty. IEEE Trans. Power Syst. 17 (4), 1081–1088.
tistage dynamic optimization problems. 1. Problems without path constraints.
Daoutidis, P., Lee, J.H., Harjunkoski, I., Skogestad, S., Baldea, M., Georgakis, C., 2018.
Ind. Eng. Chem. Res. 33 (9), 2111–2122.
Integrating operations and control: a perspective and roadmap for future re-
Vassiliadis, V.S., Sargent, R.W., Pantelides, C.C., 1994b. Solution of a class of multi-
search. Comput. Chem. Eng. 115, 179–184.
stage dynamic optimization problems. 2. Problems with path constraints. Ind.
Dias, L.S., Ierapetritou, M.G., 2017. From process control to supply chain manage-
Eng. Chem. Res. 33 (9), 2123–2133.
ment: an overview of integrated decision making strategies. Comput. Chem.
Willcox, K., Peraire, J., 2002. Balanced model reduction via the proper orthogonal
Eng. 106, 826–835.
decomposition. AIAA J. 40 (11), 2323–2330.
Dias, L.S., Ierapetritou, M.G., 2019. Optimal operation and control of intensified pro-
Zhang, Q., Grossmann, I.E., 2016. Enterprise-wide optimization for industrial de-
cesses challenges and opportunities. Curr. Opin. Chem. Eng. doi:10.1016/j.coche.
mand side management: fundamentals, advances, and perspectives. Chem. Eng.
2018.12.008.
Res. Des. 116, 114–131.
Dias, L.S., Pattison, R.C., Tsay, C., Baldea, M., Ierapetritou, M.G., 2018. A simula-
Zhang, Q., Grossmann, I.E., Heuberger, C.F., Sundaramoorthy, A., Pinto, J.M., 2015.
tion-based optimization framework for integrating scheduling and model pre-
Air separation with cryogenic energy storage: optimal scheduling considering
dictive control, and its application to air separation units. Comput. Chem. Eng.
electric energy and reserve markets. AlChE J. 61 (5), 1547–1558.
113, 139–151.
Zhang, Q., Sundaramoorthy, A., Grossmann, I.E., Pinto, J.M., 2016. A discrete-time
Dowling, A.W., Kumar, R., Zavala, V.M., 2017. A multi-scale optimization framework
scheduling model for continuous power-intensive process networks with vari-
for electricity market participation. Appl. Energy 190, 147–164.
ous power contracts. Comput. Chem. Eng. 84, 382–393.
Du, J., Park, J., Harjunkoski, I., Baldea, M., 2015. A time scale-bridging approach for
Zhao, S., Ochoa, M.P., Tang, L., Lotero, I., Gopalakrishnan, A., Grossmann, I.E., 2019.
integrating production scheduling and process control. Comput. Chem. Eng. 79,
Novel formulation for optimal schedule with demand side management in mul-
59–69.
ti-product air separation processes. Ind. Eng. Chem. Res. 58 (8), 3104–3117.
Flores-Tlacuahuac, A., Grossmann, I.E., 2006. Simultaneous cyclic scheduling and
Zhou, D., Zhou, K., Zhu, L., Zhao, J., Xu, Z., Shao, Z., Chen, X., 2017. Optimal schedul-
control of a multiproduct CSTR. Ind. Eng. Chem. Res. 45 (20), 6698–6712.
ing of multiple sets of air separation units with frequent load-change operation.
Gupta, D., Maravelias, C.T., Wassick, J.M., 2016. From rescheduling to online schedul-
Sep. Purif. Technol. 172, 178–191.
ing. Chem. Eng. Res. Des. 116, 83–97.
Zhu, Y., 2001. Multivariable System Identification for Process Control. Elsevier Sci-
Hahn, J., Edgar, T.F., 2002. An improved method for nonlinear model reduction using
ence, Oxford, UK.
balancing of empirical Gramians. Comput. Chem. Eng. 26 (10), 1379–1397.
Zhuge, J., Ierapetritou, M.G., 2012. Integration of scheduling and control with closed
Hinkley, D.V., 1969. On the ratio of two correlated normal random variables.
loop implementation. Ind. Eng. Chem. Res. 51 (25), 8550–8565.
Biometrika 56 (3), 635–639.
Zhuge, J., Ierapetritou, M.G., 2015. An integrated framework for scheduling and con-
Kelley, M.T., Pattison, R.C., Baldick, R., Baldea, M., 2018. An MILP framework for op-
trol using fast model predictive control. AlChE J. 61 (10), 3304–3319.
timizing demand response operation of air separation units. Appl. Energy 222,
951–966.

You might also like