Você está na página 1de 32

Accepted Manuscript

Title: An energy-efficient operation system for a natural gas


liquefaction process: Development and application to a 100
ton-per-day plant

Author: Wangyun Won Kwang Soon Lee

PII: S0098-1354(16)30394-5
DOI: http://dx.doi.org/doi:10.1016/j.compchemeng.2016.11.046
Reference: CACE 5635

To appear in: Computers and Chemical Engineering

Received date: 27-7-2016


Revised date: 29-11-2016
Accepted date: 30-11-2016

Please cite this article as: Won, Wangyun., & Lee, Kwang Soon., An energy-
efficient operation system for a natural gas liquefaction process: Development
and application to a 100 ton-per-day plant.Computers and Chemical Engineering
http://dx.doi.org/10.1016/j.compchemeng.2016.11.046

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
An energy-efficient operation system for a natural gas liquefaction process:
Development and application to a 100 ton-per-day plant

Wangyun Won1† and Kwang Soon Lee2*

1
Plant Research Team, GS Engineering & Construction,
33, Jong-ro, Jongno-gu, Seoul 110-121, Korea
2
Department of Chemical & Biomolecular Engineering, Sogang University,
35, Baekbeom-ro, Mapo-gu, Seoul 121-742, Korea

Submitted to Computers & Chemical Engineering

† Current address: Department of Chemical and Biological Engineering, University of


Wisconsin-Madison, Madison, WI 53706, USA

* Corresponding author: Tel.: +82-2-705-8477, Fax: +82-2-3272-0319, E-mail:


kslee@sogang.ac.kr
Highlights

> A novel energy-thrifty operation system for a natural gas liquefaction process

> Combined self-optimization control and real-time steady-state optimization

> New comprehensive procedure for the accurate evaluation of energy loss

> Performance validation in a numerical 100 ton-per-day LNG plant

Abstract

Production of liquefied natural gas (LNG) is a highly energy-intensive process. The required
liquefaction temperature is approximately −160 °C at atmospheric pressure. In this study, we propose
a two-level energy-efficient operation system for an LNG process, which consists of real-time steady-
state optimization (RTSSO) and real-time control subsystems. The RTSSO follows the typical design
to minimize set point energy losses. In contrast, the real-time control subsystem was configured to
incorporate the concept of self-optimizing control, which minimizes both implementation energy
losses and set point energy losses. Special attention was given to the loss of the evaluation step so that
the implementation energy loss is accurately estimated; thus, the designed system can be seamlessly
applied to the actual process with guaranteed optimal energy efficiency. The performance of the
proposed system was validated in a simulated LNG plant that precisely replicated an actual plant that
produces 100 tons of LNG per day.

Keywords: natural gas, liquefaction, refrigeration, control, optimization, mixed refrigerant

1
Nomenclature

d disturbance
*
d nominal disturbance
g inequality constraint
h equality constraint
u controlled variable
uopt optimized u
J operational energy consumption (kJ/kg-LNG)
J time average of J
J gradient of J
 J 2
hessian of J
LM implementation energy loss (kJ/kg-LNG)
LS set point energy loss (kJ/kg-LNG)
 execution period of real-time steady-state optimizer

Abbreviations

CV controlled variable
DOF degree of freedom
HKMR heavy-key mixed refrigerant
IMC internal model control
KSMR Korea single mixed refrigerant
LKMR light-key mixed refrigerant
LNG liquefied natural gas
MCHE main cryogenic heat exchanger
MR mixed refrigerant
MV manipulated variable
NG natural gas
PID proportional integral derivative
RGA relative gain array
RTSSO real-time steady-state optimization
SC scrub column
SCRD scrub column reflux drum
SOC self-optimizing control
SQP sequential quadratic programming

2
1. Introduction

Natural gas (NG) is the fastest growing fossil fuel for energy usage (IEA, 2015). Large amounts of
NG have been found in remote locations where pipeline transportation to consumers is infeasible
and/or uneconomical. Therefore, transporting liquefied NG (LNG) by ships is preferred (Park et al.,
2016; Son et al., 2016). LNG is obtained after conveying NG via pipelines to a nearby seashore site
and liquefying it. The liquefaction of NG can reduce the overall volume by a factor of 600, which
greatly facilitates shipment (Won et al., 2014).

The production of LNG is a highly energy-intensive process. The required liquefaction temperature is
approximately −160 °C at atmospheric pressure (Won et al., 2017). Therefore, reducing operating
energy is a key issue for LNG producers to ensure competitiveness (DOE, 2003).

The economics of the LNG process can be greatly improved by employing an optimal operation
system. However, published works on this subject are still limited. The majority of studies to date
have focused on design issues, including exergy analysis (Kanoğlu, 2002; Morosuk et al., 2015;
Remeljej & Hoadley, 2006; Roszak & Chorowski, 2013; Vatani et al., 2014), alternative process
configurations (Chang et al., 2011; Kikkawa et al., 1997; Lee et al., 2012; Mortazavi et al., 2012;
Wang et al., 2012), and new refrigerants (Aspelund et al., 2010; He & Ju, 2014; Khan et al., 2013;
Khan & Lee, 2013; Nogal et al., 2008; Xu et al., 2013). Only a few researchers have aimed to develop
energy-efficient operation systems for the refrigeration cycle only (Jensen & Skogestad, 2007) or
simplified LNG processes (Husnil et al., 2014; Michelsen et al., 2010). There is no published
literature reporting the optimal operation of a realistic industrial LNG plant.

For the optimally energy-efficient operation of a process, either the real-time steady-state optimization
(RTSSO) or the self-optimizing control (SOC) technique can be employed. RTSSO has been widely
adopted in conventional petrochemical processes because of its ability to pursue optimal values of
controlled variables (CVs) against disturbances (Diehl et al., 2002; Won et al., 2010; Zanin et al.,
2000). However, frequent on-line calculations of new set points are computationally costly. The
RTSSO is usually executed with lengthy intervals between set point calculations; therefore, the
process being optimized is vulnerable to energy losses during these intermissions. The SOC was
devised to replace the RTSSO and indirectly uses constant set points to keep the process near
optimum conditions. As a result, re-optimization is not required, even when a disturbance occurs
(Hori & Skogestad, 2007; Kariwala & Cao, 2009; Panahi & Skogestad, 2011; Skogestad, 2000). This
method is based on the idea that the energy losses depend significantly on the choice of CVs.

3
Appropriate CVs are selected that result in the least amount of energy loss due to various disturbances
when kept at their constant set points. However, for the LNG process, the performance of the constant
set point policy alone does not seem to be satisfactory (or, more specifically, the energy loss is not
acceptable) since the optimal operating points are severely affected by disturbances. In addition,
previous studies of the SOC have only considered steady-state energy losses, and the resulting
operating system is often infeasible when applied to a real process requiring trial-and-error corrections
(Larsson et al., 2001).

In this paper, considering the unique traits and operational requirements of a cryogenic process, a
comprehensive energy-efficient operation system for an LNG process is proposed by combining the
RTSSO with the SOC-based real-time control system. The RTSSO was designed to summon a
sequential quadratic programming (SQP) method for the computation of new operating points. A
surrogate model was used to provide gradient and hessian information for optimization. In the case of
the control loop design, the available manipulated variables (MVs) of the process were determined
through a degree-of-freedom (DOF) analysis. Appropriate CVs were selected from operational
requirements and paired. Candidate CVs were chosen for the unpaired free MVs, and the most
pertinent CVs were determined such that the energy loss due to various disturbance changes were
minimized, which is the idea of the SOC. Rigorous steady-state and dynamic process simulators were
used for the SOC design.

The performance of the proposed system was investigated in a simulated 100 ton-per-day LNG plant,
which was constructed to replicate an actual plant in Incheon, Korea. It serves as not only an operator
training system but also as a virtual plant to investigate energy-efficient operation systems that could
be seamlessly transferred to the actual plant. Therefore, all ancillary features of the real plant, such as
the safety interlock logic and the hard constraints of the instruments and equipment, are accurately
reflected in the simulated plant. The case study revealed that compared to the worst feasible case, a
substantial amount of energy (259.7 kJ/kg-LNG) can be saved with the aid of the proposed operating
system.

This paper is organized as follows: In Section 2, the concept of energy loss from an operational point
of view is first described. The solution method to improve the operational economics is proposed in
Section 3. In Section 4, the performance of the developed operation strategy is validated in a step-by-
step manner by applying it to the simulated LNG plant. Finally, in Section 5, the paper concludes with
a brief summary and some discussion of the evaluation of the implementation energy loss.

2. Energy loss from an operational point of view

4
We consider the situation where the optimization is performed above the process control level. Under
this situation, the MVs are not free variables, and the set points of the CVs are adjusted by the
optimizer. Let J denote the operational energy consumption of the compressors (kJ/kg-LNG). In
general, J is a function of u, d, and v, which are the CVs, disturbance, and measurement noise,
respectively.

The optimal operating conditions that minimize energy consumption can be obtained by solving the
following constrained optimization problem for a given disturbance d:

J (uopt ( d ), d , 0)  min J (u , d , 0)
u (1)
subject to h(u , d )  0 and g (u )  0

where J is the time-average of J over an appropriate temporal window, the equality constraint h = 0
denotes the process model, and g  0 defines the operational and safety limits of the process. In Eq.
(1), u is used as the decision variable for the optimization under the assumption that CVs perfectly
track their set points.

In the operation of an actual plant, it is hard for the CVs to continuously pursue the optimal value
uopt(d) mainly because of the variation of the disturbance d and the presence of v. This results in an
energy loss L, which is defined as the difference between the actual value J (u , d , v ) and the optimal

value J (uopt ( d ), d , 0) of the cost function as follows:

L(u, d , v)  J (u, d , v)  J (uopt (d ), d , 0) (2)

Now, we can separate the loss L into the set point energy loss caused by slowly and permanently
changing disturbances and the implementation energy loss caused by the fast and temporarily
fluctuating disturbances. Of course, the two different types of disturbances are mixed in a single
disturbance. The reasoning behind this separation is illustrated in Fig. 1.

In Fig. 1, uopt(d) represents the optimal value of the decision variable, which is continuously updated
according to the changes in d. The value of uopt(d*) denotes the optimized u computed for a nominal
disturbance value, d*. The solid and dashed lines indicate the energy consumptions when u is
maintained at uopt(d) and uopt(d*), respectively. The dotted line represents the energy consumption
under actual dynamic operation when u = uopt(d*). As shown in Fig. 1, the loss can be separated into
two parts. The first part is the set point energy loss, LS, which is the operational energy loss caused by
not faithfully tracking d and computing uopt(d) accordingly. This loss can be observed when uopt
remains constant at uopt(d*) without being re-optimized against a new disturbance d. The second part is
the implementation energy loss, LM, which is caused by the imperfect control performance that occurs
during real-time operation. These two losses should be treated independently considering their nature.

5
Based on the considerations above, L can be written as a sum of two terms as follows:

L (u , d , v )  LS (u , d , 0)  LM (u , d , v ) (3)

3. Energy-efficient operation system

The energy-efficient operation system for an LNG plant is designed and implemented in two steps.
First, a real-time controller (optimal control structure) is designed off-line (Fig. 2(a)). Next, the
designed real-time controller is implemented in the plant with the RTSSO in the upper level (Fig.
2(b)).

The off-line design includes the SOC-based selection of CVs (denoted as u in Fig. 2(a)). The energy
consumption, J , is evaluated for different choices of u using a rigorous process model. Different
disturbance and measurement noise scenarios are assumed for this evaluation. Based on the energy
consumption, the set point loss, LS, and the implementation loss, LM, are calculated, and optimal CVs
that yield the lowest energy loss are selected. The real-time controller is constructed based on the
selected CVs. To reduce computational time, nonlinear regression is used to derive a steady-state
surrogate model from the rigorous process model using the principal component analysis and the
neural network (Lang et al., 2009). The steady-state surrogate model is used to approximately
evaluate the gradient and hessian of J based on J  J ss , where the subscript ss means steady state.

The surrogate model is also used for the RTSSO during on-line operation.

The designed control structure is then implemented in the on-line operation system, which consists of
two major parts: the RTSSO and the real-time controller; the RTSSO includes a surrogate model. The
RTSSO regularly computes the optimum set points, uSP, for the controller such that the average energy
consumption measurement, J m , is minimized. The real-time controller pursues the optimum set
points using process measurements, um.

In the following subsections, each part of Fig. 2 is described in greater detail.

3.1. Energy loss minimization by the SOC-based control structure design

Fig. 3 shows the energy loss for two different choices of CVs. Some CVs may show a relatively large
set point loss (Fig. 3(a)). Other CVs may exhibit a smaller set point loss but also show a
comparatively large implementation loss (Fig. 3(b)). The latter case is especially important in the
LNG process primarily due to severe control loop interactions and the consequent limitation in
controllability. This leads to a large implementation loss.

The effect of CV selection on energy loss is demonstrated in Appendix A using a toy problem.

Fig. 4 explains the overall procedure of the SOC-based control structure design. First, a rigorous

6
process model was constructed using a commercial process simulator. All of the design data,
including interlock logic, were reflected in the process model. Then, the steady-state surrogate model
was derived. Second, the DOF available for the MVs was analyzed following the rules suggested by
Jensen and Skogestad (2009). Third, disturbances that have significant effects on the energy loss were
identified. To obtain a reliable prediction of the set point and implementation losses, the disturbance
signal was designed such that

d  d*  p  w (4)

where d* denotes the nominal value of d, p indicates the slowly and permanently varying signal in d,
and w represents the fast and temporarily changing signal in d. A permanently drifting signal is one
that cannot be recovered after a change, and a temporarily changing signal is one that can be restored
to the original state after a while. The two disturbance signals p and w affect the set point and
implementation losses, respectively. A first-order filtered step signal with a filter time constant of
2,000 sec was chosen for p, and a first-order filtered zero-mean random signal with a filter time
constant of 600 sec was chosen for w. A uniform distribution over a finite window was assumed for
the random signal. The magnitude of each signal was determined based on plant operation data.

In the fourth step, the set point energy loss was calculated such that

1
 J (u (d * ), d ,0)  J (uopt (d ), d ,0)  dt , d  d *  p


LS (u, d ,0)  opt
(5)
t 0

for all possible choices of u, i.e., CVs. In Eq. (5), τ represents the execution period of the RTSSO. The
MV-CV pairing for decentralized control loops was made using the relative gain array (RGA) analysis
between the MVs from the DOF analysis and all possible choices of CVs. The implementation energy
loss was then evaluated as

1
 J (u, d , v)  J (u (d * ), d , 0)  dt , d  d *  w


LM (u, d , v)  opt
(6)
t 0

PID controllers with IMC tuning (Skogestad, 2003) were assumed for all MV-CV pairs. Based on the
evaluation of LS and LM, the most efficient CVs and control loop configuration were determined in the
final step.

3.2. Energy loss minimization by the RTSSO

The RTSSO updates the optimal set points of the CVs to minimize the energy consumption, J . The
set point loss in Fig. 1 is reset to zero whenever the optimizer is invoked.

A SQP method was used to solve the constrained optimization in Eq. (1). In the programming, the

7
gradient, J ss , and the hessian,  2 J ss , are computed using the surrogate model, as shown in Fig.

2(b). The value of J is measured at the LNG plant. The surrogate model greatly reduces the
computational burden by abstaining from frequent recourse to the complex rigorous model.

4. Application study to a 100 ton-per-day industrial LNG process

The proposed energy-efficient operation strategy for the LNG process was applied to a 100 ton-per-
day industrial LNG plant constructed in Incheon, Korea, and its performance was investigated. In this
section, the process and development of the required models are described. The detailed design and
evaluation procedures of the operation system are also described.

4.1. Process description

In the 100 ton-per-day LNG plant, the Korea Single Mixed Refrigerant (KSMR) cycle, which was
developed and patented by Korea Gas Corporation (KOGAS), was adopted for the liquefaction
process (Lee et al., 2013). The overall process is composed of three major parts: (i) feed gas intake
facility; (ii) scrub and fractionation columns; and (iii) the liquefaction unit to cool and liquefy NG.
The liquefaction unit is the core part and consumes the largest amount of energy.

Fig. 5 shows the process flow diagram of the liquefaction unit and the peripheral equipment. In this
figure, two process circuits are shown; the NG circuit and the mixed refrigerant (MR) circuit. The
steady-state heat and material balance under the nominal operating condition is presented in the
Supplementary Material.

The pre-treated NG (stream 1) from the feed gas mixing drum is a mixture of nitrogen, methane,
ethane, propane, butane, and pentane at approximately 63.4 bar and 12.1 °C. This information is given
in Table 1. The NG passes through the top section of the main cryogenic heat exchanger (MCHE) and
is cooled to −16.3 °C prior to entering the scrub column (SC). The SC performs two major functions:
first, the removal of heavy hydrocarbons to avoid freezing at the cold end of the liquefaction unit, and
second, the recovery of ethane and propane for refrigerant make-up. The bottom stream of the SC
(stream 3) is routed to the fractionation section, where it is separated into ethane, propane, butane, and
heavier hydrocarbons. The vapor above the SC (stream 4) is partially condensed in the middle section
of the MCHE before being fed to the scrub column reflux drum (SCRD), where vapor and liquid are
separated. A small portion of the vapor from the SCRD is used as methane make-up for the MR. The
remaining vapor (stream 5) is sent to the bottom tube bundle of the MCHE. As it passes through the
MCHE, its temperature reduces to less than −150 °C, which condenses the NG to a liquid. The LNG

8
(stream 6) is then sent to the separator SR3 after expanding to atmospheric pressure using the control
valve V3. The bottom liquid from SR3 is pumped into an LNG storage tank. The boil-off gas from the
top is used as a fuel or is otherwise routed to the flare header.

The MR circuit is a closed refrigeration loop that supplies the cooling demands of the MCHE. The
MR consists of nitrogen, methane, ethane, and propane. It is pressurized by a series of compressors,
which are driven by independent motors and run at different speeds. The MR is pressurized up to
approximately 52.5 bar in the separator (SR2). The liquid (stream 15) and vapor (stream 17) streams
from SR2 consist of heavy and light components, respectively. The liquid is cooled further as it passes
through the top tube bundle of the MCHE and the Joule-Thomson (JT) valve V1. This stream reenters
the MCHE (stream 16) as a cooling medium, and provides the refrigeration duty required to pre-cool
the feed gas and the vapor from SR2. In addition, this stream sub-cools the liquid from SR2 and is
finally vaporized and superheated at the end of the MCHE. The fully vaporized MR is then recycled
to the compressor CP5 (stream 7). The vapor from SR2 is routed to the MCHE, where it condenses
and is sub-cooled. It exits at the cold end of the MCHE and passes through the JT valve V2 (stream
18), where its pressure and temperature are reduced. This stream is then completely vaporized and
superheated by exchanging heat with the feed NG stream and with the vapor and liquid streams from
SR2. Finally, it is returned to the compressor CP1 (stream 9).

Two independent temperature controllers are installed to manipulate V4 and V5, which are the bypass
valves. V4 adjusts the amount of pre-cooling feed gas, and V5 adjusts the SC top cooling. All five
compressors are of the centrifugal type and are equipped with anti-surge recycle lines, suction drums,
and fan-type discharge coolers. A distributed control system from Yokogawa Co. is used for receiving
and transmitting field signals, safety interlock logics, and local controllers.

The operation system for this process needs to be developed in a feasible form using the conditions
described above.

4.2. Model development

Aspen HYSYS® was used to develop rigorous dynamic and steady-state models. The whole design
data of the actual plant, such as the piping and instrumentation diagram, the cause and effect diagram,
the plot plan, the datasheet for equipment and instruments, and various vendor prints, were reflected
in the models. The key parameters included the compressor and motor performance curves and were
corrected using equipment operation data as delivered by the manufacturers (see Appendix B for the
results). The surrogate model was developed using MATLAB with ActiveX objects.

4.3. Design of an energy-efficient operation system

9
4.3.1. Degree-of-freedom analysis

The DOF for the LNG process in Fig. 5 is analyzed and summarized in Table 2. The anti-surge valves
and the suction drums for compressors CP1, CP2, CP3 and CP5 were not considered in the DOF
analysis because they are provided for abnormal states. On the other hand, the levels of the two flash
drums (SR1 and SR2) have steady-state effects. In other words, the MR composition and active
charge can be varied by controlling the levels in these drums. Therefore, the number of actual steady-
state DOFs can be counted. The number of MVs with no steady-state effect is zero, as indicated by the
asterisk in Table 2. Here, the active charge is defined as the total mass of refrigerant accumulated in
the MR cycle, which is mainly in the heat exchangers.

According to Jensen and Skogestad (2009), the liquefaction unit in Fig. 5 potentially has twenty DOFs.
Compared to the unit in Jensen and Skogestad (2009), the following six DOFs in the present process
are not installed or considered: the bypass flow rates of light-key MR (LKMR) and heavy-key MR
(HKMR) in the MCHE, the two levels of SR1 and SR2, and the control valves that adjust the pressure
of SR1 and SR2.

The utilization of actual DOFs for optimal operation is discussed in Section 4.3.4.

4.3.2. Definition of optimal operation

The energy consumed by the five compressors (CP1 to CP5) was formulated as the cost function
shown in Eq. (1) and minimized. The interlock logic such as emergency shutdown, process shutdown,
and compressor vendor shutdown were considered as constraints. In addition, the set values for the
pressure safety valves were also used as constraints.

4.3.3. Identification of important disturbances

In the case where NG is produced from a gas well, the feed conditions can vary with time. Moreover,
depending on the process operating condition (e.g., changes in the required amount of boil-off gas
from SR3) or market condition (e.g., changes in the required LNG heating value), the desired value of
LNG temperature and MR pressure can change. These changes have a significant effect on the energy
consumption. The variables mentioned above are listed in Table 3 with their signal characteristics.

In this study, RTSSO was assumed to be invoked every 2 hours. Magnitudes of the disturbances
during the RTSSO intermission were assumed based on this assumption and the real operation data, as
shown in Table 3. Magnitudes of measurement noises were assumed considering the sensor accuracy

10
provided by manufacturers.

4.3.4. Selection of CVs - Design of SOC

Alternative choices for the CVs were compared by evaluating the energy loss against the presumed set
of disturbances and measurement noises.

Thirteen out of the fourteen DOFs listed in Table 2 were paired with thirteen CVs, as presented in
Table 4. The MV-CV pairings for the first eleven loops are rather obvious and are in accordance with
industrial practice, where the stability of the process operation is the highest priority. The LNG
temperature and the production rate are important CVs, and they can be controlled by selecting V1,
V2 or V3. The remaining MV can be used directly for operation optimization. However, the better
strategy is to form a control loop with a CV that can yield minimum energy loss.

Unlike loops 1 to 11 (shown in Table 4), the pertinent MVs for the LNG temperature and the
production rate are unclear. Moreover, interaction among loops 12 to 14 can be varied depending on
the choice of CV to be determined. Therefore, appropriate pairings were found before evaluating the
loss. This was done using RGA analysis (Seborg et al., 2010).

Table 5 lists ten candidate CVs that were considered in this study. The nominal values were obtained
through steady-state optimization. Among the ten candidate CVs, Y1 to Y4 are shown in Fig. 5. Note
that Y3, Y4, Y8, Y9 and Y10 are designed as a combination of two measurements to explore an
opportunity to further enhance economic performance. The table also shows pairing results for each
candidate CV. Note that in all cases, V3 is paired with the LNG flow rate.

Investigation of the set point energy loss:

Fig. 6 shows energy consumption as a function of Y1 when D2 (feed gas pressure) has two different
values (nominal and permanent drift). The value of Y1 has a minimum that depends on the size of the
disturbance (i.e., 28 °C for D2 = 63.4 bar and 31.6 °C for D2 = 69.9 bar). If the HKMR temperature
remains constant, despite a change in D2 (63.4 → 69.9 bar), an energy loss will occur.

Fig. 7 exhibits the set point losses incurred by keeping alternative CVs fixed at their nominal optimal
values when there are variations in D4 (CH4 mole fraction in the feed gas) and D6 (CP4 outlet
pressure). To evaluate the set point losses, the process is first regulated at the nominal steady state and
is then perturbed to a new steady state using a pre-designed disturbance signal described in Section
3.1. The set point losses are computed using Eq. (5) for all candidate CVs in Table 5. The result shows

11
that Y5 to Y10 should be eliminated from the list of CV candidates because of large energy losses. In
contrast, Y1 to Y4 can be considered as CV candidates because the energy loss (at least for the two
disturbance cases) is relatively small.

In general, temperature-related variables have small losses, while flow rate-related variables result in
a larger loss. The reason for this is that a thermal balance exists between the NG and MR in the
MCHE that needs to be met, especially given the temperature and production rate of LNG. Keeping
the refrigerant flow rate constant provides an identical cold duty despite disturbance variations, and
the required thermal balance may not be satisfied.
Investigation of the implementation energy loss:

To investigate the implementation energy loss, we generated fifteen different disturbance scenarios for
each disturbance variable listed in Table 3. Then, we conducted dynamic simulation with regulatory
control for all of the disturbance scenarios and took an average of the estimated implementation losses
for each CV. Fig. 8 depicts selected real-time control results and associated implementation energy
losses due to a temporary change in D2. Regulatory control was conducted for Y1, Y4, and Y6 against
the disturbance change. As shown in Fig. 8(c), when Y6 (LKMR flow rate) was selected as a CV, a
decrease in D2 results in a decrease in CP5 suction temperature, and the emergency shutdown is
eventually activated. Here, the CP5 suction temperature is slightly different from Y1, at
approximately 0.1-0.5 °C. Y1 is the temperature of HKMR leaving the MCHE, while the CP5 suction
temperature is measured immediately upstream of the compressor and is used for the interlock logic.
During this transient stage, compared to other cases, the LNG temperature also undergoes a large
excursion, and the accumulated implementation loss grows quickly. On the other hand, the regulation
of Y1 and Y4 does not cause any noticeable trouble in process operation. An implementation loss with
limited variation around the optimized value is produced.
Selection of CVs:

The set point and implementation energy losses for all candidate CVs against all potential
disturbances (as shown in Table 3) are summarized in Fig. 9. The value of an energy loss for the CV is
the sum of the energy losses incurred by the individual disturbances. Detailed values of the losses are
given in Table C.1 in Appendix C.

As shown in Fig. 9, Y3 was found to be the most promising self-optimizing CV. In the presence of
disturbances, this variable has the lowest energy loss when held constant. The energy loss of Y3 is
259.7 kJ/kg-LNG lower than Y4, which is the worst feasible case. Note that Y1 would be regarded as
the best choice only if the set point loss was considered. This has also been demonstrated in other
publications (Larsson et al., 2001; Skogestad, 2000).

In the case of set point loss, Y4 is an alternative to Y1 and Y3. However, Y4 shows a relatively long

12
settling time and leads to a larger implementation loss compared to other feasible CV candidates. The
implementation losses of Y2 and Y5 were estimated to be infeasible, i.e., emergency shutdown was
activated during the energy loss evaluation, although their set point losses were feasible.

4.4. Evaluation of the designed operation system

Fig. 10 demonstrates the performance of the two-layered operation system (as shown in Fig. 2(b)). In
this simulation, Y3, which produced the lowest energy loss (Fig. 9), was used as a CV with V1 as its
MV (see Table 5).

As D2 drifts from its initial value for the first 5,400 sec, the set point energy loss, which is the
difference between the actual and optimized values of energy consumption, gradually increases. This
is shown in Fig. 10(e). After 7,200 sec had elapsed, the RTSSO was executed, and the set points for
all CVs, except LNG temperature and production rate, were updated to the optimized values based on
the new disturbance values. Fig. 10(b) shows an exemplary case of how Y3 responds to the new set
point. The designed operation system stabilizes the process quickly against the temporary disturbance
in approximately 9,600 sec and results in a moderate energy loss.

5. Conclusions and discussion

This study proposed a novel energy-efficient operation system for an LNG process. The key idea was
to minimize the set point energy loss by employing the RTSSO. Simultaneously, energy losses
incurred by imperfect control and non-renewed steady-state operating points during RTSSO
intermission were minimized by designing the SOC-based control subsystem. Regarding the SOC
design in particular, a new comprehensive procedure was proposed for the accurate evaluation of
energy loss. The basic concept was taken from Skogestad (2000). However, the details contain
elaborations and improvements, so any implementation losses that were overlooked in previous SOC
publications (Hori & Skogestad, 2007; Kariwala & Cao, 2009; Larsson et al., 2001; Panahi &
Skogestad, 2011; Skogestad, 2000) were precisely estimated. This improvement is considerable,
especially in the LNG plant, where the magnitude of the implementation loss is comparable with that
of the set point loss and has a significant effect on the CV selection. The case study conducted in a
simulated 100 ton-per-day LNG plant revealed that, compared to the worst feasible case, a substantial
amount of energy (259.7 kJ/kg-LNG) can be saved with the aid of the proposed operating system. The
disturbances used for loss evaluation were generated based on real operation data.

It is noted that the main issue in this study was not a simple amendment of the control algorithm but a

13
structural decision made before the control system was implemented.

The energy-efficient operation strategy proposed in this study is not limited to a specific LNG plant.
Therefore, it can be used for other liquefaction cycles after modifying the relevant process-dependent
features. Some examples include the propane pre-cooled MR (C3-MR) process offered by Air
Products and the dual MR (DMR) process supplied by Shell. These processes have a similar structure,
which includes two separate compressor loops for vapor and liquid MRs, and a multi-tube heat
exchanger that utilizes two-phase MR.

14
Appendix

Appendix A. Effect of the CV selection on the energy loss.

In this appendix, the 2x1 dynamic process is used to describe how CV selection affects energy loss:

 0.15   0.4 1.8 


 u1   ( s  0.3)( s  0.5)   s  1.5   d1 
u     z   s 1  (A.1)
 2  2.56   0.55 1   d 2 
 ( s  2)( s  0.8)   s  0.5 s  2 
 

J  2 z 2  (3d1  d 2 ) z  14.5 (A.2)

where u1 and u2 are candidate CVs; z is an MV; d1 and d2 indicate disturbances; and J represents an
objective function (e.g., to minimize energy consumption).

When [d1 d2]T=[3 1]T and z=2.5, the steady-state optimum values of candidate CVs are calculated as
[u1 u2]T=[4.9 1.2]T. A value of J=2 is obtained. If u1 is chosen to be a constant, and the disturbance is
permanently perturbed to [d1 d2]T=[2 2]T, the feedback controller will steer z to get u1=4.9. This leads
to z=1.7 and J=6.68. If the set point of u1 is re-optimized for the disturbance change, such that u1=5.2,
the MV would settle at z=2, leading to J=6.5, the set point loss in this case will be LS=0.18. Likewise,
the set point loss of u2 for the identical disturbance drift can be computed such that LS=0.5. From the
set point loss point of view, u1 is the better choice.

On the other hand, the process response of u2 is three times faster than u1. The sum of the time
constants for u1 is 5.33, while that of u2 is 1.75. When a temporary disturbance that varies at
approximately [d1 d2]T=[3 1]T is injected, the controller for u2 can stabilize the process much faster
than the controller for u1. This leads to a smaller implementation loss, and u2 is the better choice from
the implementation loss point of view.

Appendix B. Model parameter estimation.

For correct representation of the behavior of the LNG process, key parameters in the process model
were corrected based on operation data supplied by manufacturers. Before model tuning, the data
were first pretreated by eliminating outliers, supplementing missing data, and mitigating noise using a
non-causal low pass filter. The compressor performance and efficiency curves at various speeds were
modeled as second- to fourth-order polynomials, respectively. The corresponding coefficients were
obtained by fitting the polynomial equations to the experimental data. The surge line was also

15
determined by fitting the measured data to a third-order polynomial. Motor efficiency and torque data
were fitted to eighth- and fourth- order polynomials, respectively.

Fig. B.1 demonstrates fitting results of the operation data (symbols) for compressor CP1 and its driver
(motor) to the regression polynomials (lines). Similar results were obtained for other compressors
driven by independent motors at a different pressure levels. The power consumption for each
compressor was calculated from these estimated curves rather than sole computations using the ideal
equation.

Regarding the MCHE model, both process and utility side heat transfer coefficients, and the metal
mass, were appropriately tuned to have reasonable statics and dynamics using commissioning data
from another LNG plant. This plant has an identical capacity (100 ton-per-day) and uses the same
type of heat exchanger (plate-fin).

Appendix C. Energy loss for alternative CVs.

16
References

Aspelund, A., Gundersen, T., Myklebust, J., Nowak, M. P., & Tomasgard, A. (2010). An optimization-
simulation model for a simple LNG process. Computers & Chemical Engineering, 34, 1606-
1617.

Chang, H. M., Chung, M. J., Lee, S., & Choe, K. H. (2011). An efficient multi-stage Brayton-JT cycle
for liquefaction of natural gas. Cryogenics, 51, 278-286.

Department of Energy (DOE) (2003). The global liquefied natural gas market: status & outlook. In
Energy Information Administration (Ed.), Washington.

Diehl, M., Bock, H. G., Schlöder, J. P., Findeisen, R., Nagy, Z., & Allgöwer, F. (2002). Real-time
optimization and nonlinear model predictive control of processes governed by differential-
algebraic equations. Journal of Process Control, 12, 577-585.

He, T. & Ju, Y. (2014). Design and optimization of a novel mixed refrigerant cycle integrated with
NGL recovery process for small-scale LNG plant. Industrial & Engineering Chemistry Research,
53, 5545-5553.

Hori, E. S. & Skogestad, S. (2007). Selection of control structure and temperature location for two-
product distillation columns. Chemical Engineering Research and Design, 85, 293-306.

Husnil, Y. A., Yeo, G., & Lee, M. (2014). Plant-wide control for the economic operation of modified
single mixed refrigerant process for an offshore natural gas liquefaction plant. Chemical
Engineering Research and Design, 92, 679-691.

International Energy Agency (IEA) (2015). World Energy Outlook.

Jensen, J. B. & Skogestad, S. (2007). Optimal operation of simple refrigeration cycles. Part II:
selection of CVs. Computers & Chemical Engineering, 31, 1590-1601.

Jensen, J. B. & Skogestad, S. (2009). Steady-state operational degrees of freedom with application to
refrigeration cycles. Industrial & Engineering Chemistry Research, 48, 6652-6659.

Kanoğlu, M. (2002). Exergy analysis of multistage cascade refrigeration cycle used for natural gas
liquefaction. International Journal of Energy Research, 26, 763-774.

Kariwala, V. & Cao, Y. (2009). Bidirectional branch and bound for CV selection part II: exact local
method for self-optimizing control. Computers & Chemical Engineering, 33, 1402-1412.

Khan, M. S. & Lee, M. (2013). Design optimization of single mixed refrigerant natural gas
liquefaction process using the particle swarm paradigm with nonlinear constraints. Energy, 49,
146-155.

Khan, M. S., Lee, S., Rangaiah, G. P., & Lee, M. (2013). Knowledge based decision making method
for the selection of mixed refrigerant systems for energy efficient LNG processes. Applied
Energy, 111, 1018-1031.

17
Kikkawa, Y., Nakamura, M., & Sugiyama, S. (1997). Development of liquefaction process for natural
gas. Journal of Chemical Engineering of Japan, 30, 625-630.

Lang, Y., Malacina, A., Biegler, L. T., Munteanu, S., Madsen, J. I., & Zitney, S. E. (2009). Reduced
order model based on principal component analysis for process simulation and optimization.
Energy & Fuels, 23, 1695-1706.

Larsson, T., Hestetun, K., Hovland, E., & Skogestad, S. (2001). Self-optimizing control of a large-
scale plant: the Tennessee Eastman process. Industrial & Engineering Chemistry Research, 40,
4889-4901.

Lee, S. G., Choe, K. H., Yang, Y. M., Lee, C. G., Cha, K. S., Park, C. W., Choi, S. H., & Lee, Y. B.
(2013). Natural gas liquefaction process. US patent, US20130133362A1, USA.

Lee, S., Long, N. V. D., & Lee, M. (2012). Design and optimization of natural gas liquefaction and
recovery processes for offshore floating liquefied natural gas plants. Industrial & Engineering
Chemistry Research, 51, 10021-10030.

Michelsen, F. A., Halvorsen, I. J., Lund, B. F., & Wahl, P. E. (2010). Modeling and simulation for
control of the TEALARC liquefied natural gas process. Industrial & Engineering Chemistry
Research, 49, 7389-7397.

Morosuk, T., Tesch, S., Hiemann, A., Tsatsaronis, G., & Omar, N. B. (2015). Evaluation of the PRICO
liquefaction process using exergy-based methods. Journal of Natural Gas Science Engineering,
27, 23-31.

Mortazavi, A., Somers, C., Hwang, Y., Radermacher, R., Rodgers, P., & Hashimi, S. (2012).
Performance enhancement of propane pre-cooled mixed refrigerant LNG plant. Applied Energy,
93, 125-131.

Nogal, F. D., Kim, J. K., Perry, S., & Smith, R. (2008). Optimal design of mixed refrigerant cycles.
Industrial & Engineering Chemistry Research, 47, 8724-8740.

Panahi, M. & Skogestad, S. (2011). Economically efficient operation of CO2 capturing process part I:
self-optimizing procedure for selecting the best CVs. Chemical Engineering and Processing:
Process Intensification, 50, 247-253.

Park, K., Won, W., & Shin, D. (2016). Effects of varying the ambient temperature on the performance
of a single mixed refrigerant liquefaction process. Journal of Natural Gas Science and
Engineering, 34, 958-968.

Remeljej, C. W. & Hoadley, A. F. A. (2006). An exergy analysis of small-scale liquefied natural gas
(LNG) liquefaction processes. Energy, 31, 2005-2019.

Roszak, E. A. & Chorowski, M. (2013). Exergy analysis of combined simultaneous liquid natural gas
vaporization and adsorbed natural gas cooling. Fuel, 111, 755-762.

Seborg, D. E., Mellichamp, D. A., Edgar, T. F., & Doyle, F. J. (2010). Process Dynamics and Control,
3rd ed., New York, John Wiley & Sons.

18
Skogestad, S. (2000). Plantwide control: the search for the self-optimizing control structure. Journal
of Process Control, 10, 487-507.

Skogestad, S. (2003). Simple analytic rules for model reduction and PID controller tuning. Journal of
Process Control, 13, 291-309.

Son, Y., Min, K., & Lee, K. S. (2016). A liquid distribution model for a column with structured
packing under offshore conditions. Chemical Engineering Science, 153, 199-211.

Vatani, A., Mehrpooya, M., & Palizdar, A. (2014). Advanced exergetic analysis of five natural gas
liquefaction processes. Energy Conversion and Management, 78, 720-737.

Wang, M., Zhang, J., & Xu, Q. (2012). Optimal design and operation of a C3MR refrigeration system
for natural gas liquefaction. Computers & Chemical Engineering, 39, 84-95.

Won, W., Lee, S. K., Choi, K., & Kwon, Y. (2014). Current trends for the floating liquefied natural gas
(FLNG) technologies. Korean Journal of Chemical Engineering, 31, 732-743.

Won, W. & Kim, J. (2017). Bi-level optimizing operation of natural gas liquefaction process.
Computers & Chemical Engineering, 96, 87-102.

Won, W., Lee, K. S., Lee, S., & Jung, C. (2010). Repetitive control and online optimization of catofin
propane process. Computers & Chemical Engineering, 34, 508-517.

Xu, X., Liu, J., Jiang, C., & Cao, L. (2013). The correlation between mixed refrigerant composition
and ambient conditions in the PRICO LNG process. Applied Energy, 102, 1127-1136.

Zanin, A. C., Gouvêa, M. T., & Odloak, D. (2000). Industrial implementation of a real-time
optimization strategy for maximizing production of LPG in a FCC unit. Computers & Chemical
Engineering, 24, 525-531.

19
(J )

Fig. 1. Energy losses from an operational point of view.

J ss ,  2 J ss Jm

u SP

um

J u, d , v
J ss ,  2 J ss
m

Fig. 2. Proposed energy-efficient operating system: (a) off-line determination of the optimal control
structure, and (b) on-line implementation of the control structure with the RTSSO at the upper level.

20
uopt ,1 (d * )  constant
(J )

(J )
uopt ,2 ( d * )  constant

J J

Fig. 3. Set point and implementation losses for the two alternative choices of CVs.

Fig. 4. Design procedure for the energy-efficient operating system.

21
Fig. 5. A process flow diagram of the liquefaction unit and scrub column in the LNG plant. CP:
compressor, CL: cooler, MCHE: main cryogenic heat exchanger, SC: scrub column, SCRD: scrub
column reflux drum, and SR: separator.

Disturbance
1320 (D2=69.9 bar)

1315 Nominal
(D2=63.4 bar)
1310

1305 loss

1300

1295
20 25 30 35 39
Y1 (HKMR temperature to CP5) (oC)

Fig. 6. Energy consumption as a function of HKMR temperature to CP5. The other remaining
degrees-of-freedom were optimized.

20
Energy loss (kJ/kg-LNG)
Energy loss (kJ/kg-LNG)

15 Y6
Y5

10 Y7
Y10
Y3

5
Y9
Y1,Y2,
0 Y4, Y8
0.82 0.84 0.86 0.88 0.9
D4 (CH4 mole fraction in feed gas)
(a)

Fig. 7. Energy losses of alternative CVs for permanently drifted disturbances in (a) D4 (CH4 mole
fraction in feed gas) and (b) D6 (CP4 outlet pressure). A line that ends corresponds to infeasible

22
operation.

CP5 suction temperature (oC)

x104
Energy consumption (kJ/kg-LNG)

6
ESD
Accumulated energy loss

5
(kJ∙sec/kg-LNG)

0
0 2000 4000 6000
Time (sec)
(f)

Fig. 8. Results of control and energy loss against temporary change in D2 (feed gas pressure) when
Y1, Y4 and Y6 are each selected as a CV and regulated: (a) variations in D2, (b) regulation of selected
CVs, (c) CP5 suction temperature, (d) response of LNG temperature, (e) energy consumption, and (f)
accumulated energy loss. The dotted lines indicate the set points. ESD: emergency shutdown activated
by interlock logic.

23
Fig. 9. Energy losses for alternative CVs.

10.5

Y3 (warm end delta temp.) (oC)


64 RTSSO executed
D2 (Feed gas pressure) (bar)

63
62 10
Permanent
disturbance Temporary
61
disturbance
60 9.5
59
58
9
0 4000 8000 12000 0 4000 8000 12000
Time (sec) Time (sec)
(a) (b)

-151.4 4454
LNG temperature (oC)

-151.5
LNG flow rate (kg/hr)

4452
-151.6 RTSSO executed RTSSO executed

-151.7 4450

-151.8
4448
-151.9

-152.0 4446
0 4000 8000 12000 0 4000 8000 12000
Time (sec) Time (sec)
(c) (d)

24
Energy consumption (kJ/kg-LNG)

SR2 level (m)


Fig. 10. Energy loss under the designed operating system: (a) feed gas pressure, (b) warm end delta
temperature, (c) LNG temperature, (d) LNG flow rate, (e) energy consumption, and (f) SR2 level. The
dotted lines indicate the set points.

25
Normalized head

1 1

0.8

Normalized torque
0.95
Efficiency

0.6
0.9
Efficiency
0.4
Normalized torque
0.85
0.2

0.8 0
0 0.2 0.4 0.6 0.8 1
Normalized speed
(c)

Fig. B.1. Compressor (CP1 in Fig. 5) operation data (symbols) and fitting results (lines) for: (a) head,
(b) efficiency, and (c) driver motor.

26
Table 1. LNG temperature and feed gas conditions.

Condition Value

LNG temperature at the exit of MCHE (oC) -150.7


Feed gas temperature (oC) 12.1
Feed gas pressure (bar) 63.4
Nitrogen 0.0496
Methane 0.8690
Ethane 0.0510
Propane 0.0213
Feed gas composition
i-Butane 0.0044
n-Butane 0.0045
i-Pentane 0.0001
n-Pentane 0.0001

Table 2. Degree-of-freedom analysis.

LNG valve position (V3) 1


Compressor power (CP1-CP5) 5
Choke valve position (V1,V2) 2
Pump power (Pump) 1
+ Heat exchanger fan speed (CL1-CL5) 5

Total manipulable variables 14


– Manipulable variables with no steady-state effect 0*

= Actual steady-state degrees of freedom 14

* The number of degree-of-freedom (level tanks) with no steady-state effect is zero.

Table 3. Important disturbances that have a significant effect on the energy consumption.
Magnitude of disturbance
Nominal Measurement
Description Permanently Temporarily
value noise, |v|max
drifted, |p| changed, |w|max

D1 Feed gas flow rate (kg/h) 4970 250 100 5

D2 Feed gas pressure (bar) 63.4 6.5 4 0.2

D3 Feed gas temperature (oC) 12.1 5 7 0.2

27
D4 Mole fraction of CH4 in the feed gas* 0.8690 0.05 0.05 0.01

D5 LNG temperature (oC)† -151.7 3 4.5 0.2

D6 CP4 compressor outlet pressure (bar)† 52.5 5 7 0.2


*
The mole fraction of other components is adjusted proportionally to their original value (to make the
sum of the mole fraction unity).

These variables are introduced as disturbances to consider the effects of (1) regulation error that
exists in an actual process and (2) tracking error during the transient period to a new set point on
energy losses.

28
Table 4. Decentralized control structure.

Loop MV CV

1 CP1 power CP1 suction pressure


2 CP2 power CP2 pressure ratio
3 CP3 power CP3 pressure ratio
4 CP4 power CP4 pressure ratio
5 CP5 power CP5 suction pressure
6 CL1 fan speed CP1 downstream temperature
7 CL2 fan speed CP2 downstream temperature
8 CL3 fan speed CP3 downstream temperature
9 CL4 fan speed CP4 downstream temperature
10 CL5 fan speed CP5 downstream temperature
11 Pump power SR1 level
V1 valve position LNG temperature
12-14* V2 valve position LNG production rate
V3 valve position Need to be determined†

* Pairing is required.

It will be determined through loss evaluation.

29
Table 5. Candidate CVs for a free MV among V1, V2, and V3.

Description Nominal Pairing of loops 12-14

Y1 HKMR temperature to CP5 (oC) 27.9 V1-Y1 V2-LNG T. V3-LNG F.

Y2 LKMR temperature to CP1 (oC) 25.3 V1-LNG T. V2-Y2 V3-LNG F.

Y3 * Warm end delta temperature (oC) 9.2 V1-Y3 V2-LNG T. V3-LNG F.



Y4 Cold end delta temperature (oC) 11.9 V1-LNG T. V2-Y4 V3-LNG F.

Y5 SR2 pressure (bar) 53.1 V1-Y5 V2-LNG T. V3-LNG F.

Y6 LKMR flow rate (kg/h) 9,175 V1-LNG T. V2-Y6 V3-LNG F.

Y7 HKMR flow rate (kg/h) 9,072 V1-Y7 V2-LNG T. V3-LNG F.

Y8 LKMR to HKMR flow rate ratio 1.011 V1-Y8 V2-LNG T. V3-LNG F.

Y9 LNG to LKMR flow rate ratio 0.485 V1-LNG T. V2-Y9 V3-LNG F.

Y10 LNG to HKMR flow rate ratio 0.491 V1-Y10 V2-LNG T. V3-LNG F.

* Temperature difference between HKMR from SR2 and HKMR to CP5



Temperature difference between LKMR from SR1 and LKMR to CP1
LNG T. = LNG temperature, LNG F. = LNG flow rate

Table C.1. Energy loss (kJ/kg-LNG) for alternative CVs.


Candidate CVs
Disturbance
Y1 Y2 Y3 Y4 Y5 Y6 Y7 Y8 Y9 Y10
D1 36.38 88.86 32.70 66.54 152.17 infeas 134.23 infeas infeas 121.45
D2 4.54 13.89 24.49 10.43 66.94 35.50 21.78 23.29 42.82 15.57
Set point D3 0.05 0.02 0.06 0.02 0.03 8.87 0.04 6.01 19.79 0.03
energy loss D4 5.33 3.99 5.43 6.55 14.67 infeas 8.64 17.76 infeas 8.87
D5 39.80 98.98 34.33 86.30 24.65 infeas 29.19 infeas infeas 23.07
D6 3.60 6.95 1.35 2.61 62.93 infeas 50.92 infeas infeas 45.61
Subtotal (a) 89.69 212.67 98.35 172.45 321.38 infeas 244.80 infeas infeas 214.59
D1 15.42 infeas 9.03 54.97 infeas infeas 21.07 infeas infeas 14.88
D2 2.49 12.65 0.69 8.51 13.57 infeas 5.52 53.29 infeas 2.00
D3 0.72 6.84 0.59 4.46 10.39 13.16 0.01 14.82 17.96 2.44
Implementation
D4 7.97 22.31 5.59 16.11 19.73 infeas 7.78 29.37 infeas 10.72
energy loss
D5 62.03 infeas 26.60 128.80 infeas infeas 56.66 infeas infeas 45.85
D6 2.42 17.54 1.37 11.59 21.54 infeas 8.17 infeas infeas 0.47
Meas. noise* 6.24 12.71 6.71 11.76 22.10 53.51 6.92 41.07 64.05 9.27
Subtotal (b) 97.27 infeas 50.58 236.20 infeas infeas 106.12 infeas infeas 85.63
Total (=(a)+(b)) 186.97 infeas 148.93 408.64 infeas infeas 350.93 infeas infeas 300.22
* Sum of implementation loss caused by measurement noises

30

Você também pode gostar