You are on page 1of 9

Applied Thermal Engineering 30 (2010) 1538e1546

Contents lists available at ScienceDirect

Applied Thermal Engineering


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

Assessment of the controlling envelope of a model-based multivariable


controller for vapor compression refrigeration systems
Leonardo C. Schurt a, Christian J.L. Hermes a, *, Alexandre Trono Neto b
a
b

POLO Research Laboratories for Emerging Technologies in Cooling and Thermophysics, Federal University of Santa Catarina, 88040-900 Florianpolis, SC, Brazil
Department of Automation and Systems Engineering, Federal University of Santa Catarina, 88040-900 Florianpolis, SC, Brazil

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 17 February 2010
Accepted 19 February 2010
Available online 3 March 2010

This paper explores the controlling characteristics of a rst-principles model-based controller specially
developed for vapor compression refrigeration systems. Mathematical sub-models were put forward for each
of the system components: heat exchangers (condenser and evaporator), variable-speed compressor and
variable-orice electric expansion device. The dynamic simulation model was then used to design a multivariable controller based on the linear-quadratic-Gaussian technique using a Kalman lter for the estimator
design. A purpose-built testing apparatus comprised of a variable-speed compressor and a pulse-width
modulated expansion valve was used to collect data for the system identication, and model and controller
validation exercises. It was found that the model reproduces the experimental trends of the working pressures and power consumption in conditions far from the nominal point of operation (30%) with a maximum
deviation of 5%. Additional experiments were also performed to verify the ability of the controller of
tracking reference changes and rejecting thermal load disturbances. It was found that the controller is able to
keep the refrigeration system running properly when the thermal load was changed from 340 to 580 W
(460 W nominal), and the evaporator superheating degree was varied from 9.5  C to 22  C (16.6  C nominal).
2010 Elsevier Ltd. All rights reserved.

Keywords:
Control
Refrigeration system
Modeling
Experimentation
Controlling envelope

1. Introduction
Household and commercial refrigerators consume by 10% of the
electrical energy produced worldwide, a gure that has motivated
both customers and governments to push the manufacturers for
high efciency products. However, in spite of the large effort put in
refrigeration systems advancements in the past years, just a modest
effect has been observed suggesting that the conventional vapor
compression refrigeration technology is reaching its limits. Vapor
compression refrigeration systems usually comprise a single-speed
compressor, a xed-orice expansion device, two heat exchangers
(i.e., the condenser and the evaporator), and a volatile working uid
(named refrigerant) that undergoes a reversed Rankine thermodynamic cycle. Furthermore, the temperature of the refrigerated
compartment is controlled by a thermostat that switches the
refrigeration system on and off according to a cycling pattern.
It has been advocated in the open literature that the refrigeration
systems with electronic-controlled equipment (i.e., compressor
speed, and valve opening, among others) may improve signicantly
the overall energy performance when compared to the conventional
* Corresponding author. Present address: Department of Mechanical Engineering,
Federal University of Paran, P.O. Box 19011, 81531-990 Curitiba, PR, Brazil.
Tel.: 55 41 3361 3239; fax: 55 41 3361 3129.
E-mail address: chermes@ufpr.br (C.J.L. Hermes).
1359-4311/$ e see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.applthermaleng.2010.02.024

refrigeration appliances. For instance, variable-speed compressors


have been employed in the past decades to obtain a continuous
matching between the cooling capacity and the thermal load with
remarkable gains in energy performance [1,2]. Moreover, it has been
also reported [3,4] that the system performance improves further in
cases where a variable-orice expansion valve is additionally
employed, as the evaporator is kept fully activated (i.e., ooded)
during most of the system runtime.
It is noteworthy that such refrigeration systems require proper
control strategies to account for the simultaneous operation of
a variable-speed compressor and a variable-orice expansion valve.
Albeit there have been in the past years some publications with
regard to the analysis and development of control strategies for
variable-speed compressors and variable-orice expansion valves
operating simultaneously [5e8], their focus was on large capacity
refrigeration applications (>6 kW).
Pottker and Melo [4] carried out an experimental study to evaluate the effect of variable-opening electric expansion valves on the
performance of variable-speed (VSC) and single-speed (SSC) vapor
compression refrigeration systems. A purpose-built breadboard
refrigeration apparatus consisting of a variable-speed compressor, an
electric expansion valve (EEV), and two secondary ow rate and
temperature controlled aqueous solution loops (condensing and
evaporating media) was designed and constructed to emulate
refrigeration systems with capacities ranging from 0.3 to 1.5 kW

L.C. Schurt et al. / Applied Thermal Engineering 30 (2010) 1538e1546

Nomenclature

Roman
A
cross-sectional area, m2
A, B, C, E, W state-space matrices
valve opening, %
Av
D
coil inner diameter, m
h
specic enthalpy, J/kg
J
cost function
K
controller gain matrix
L
length, m
L
observer gain matrix
m
mass ow rate, kg/s
N
compressor speed, s1
NTU
number of transfer units
p
pressure, Pa
Q
heat transfer rate, W
q
heatux, W m2
Q, R
weighting matrices
T
temperature, K
t
time, s
u
controllable inputs matrix
UA
overall conductance, W/K
V
secondary coolant ow rate, L/min
v
specic volume, m3/kg
w
disturbance input matrix

(see Fig. 1). It was found that, independently of the secondary uid
temperature entering the heat exchangers, the VSC/EEV system
always operates with a maximum coefcient of performance (COP)
once it keeps the evaporator ooded and also delivers a cooling
capacity that exactly matches the thermal load.
Later, Marcinichen et al. [9] put forward an empirical dual-SISO
(single-input, single-output) strategy for the simultaneous control of
compressor speed and expansion valve opening using the experimental rig constructed by Pottker and Melo [4]. The control strategy
was devised to obtain a maximum COP within a range of cooling
capacity from 0.3 to 0.8 kW. The refrigeration system was identied
using the step-response method, which provided rst order linear
models for both the evaporator superheating and the brine outlet
temperature. The empirical models were used to derive two singleinput, single-output (SISO) proportional-integral (PI) controllers, one
for the evaporator superheating as a function of the EEV opening, and
another for brine outlet temperature as a function of the compressor
speed. Both controllers were implemented into a dual-SISO control
strategy that operated satisfactorily in terms of reference tracking
and disturbance rejection. Nonetheless, it was found that the
empirical identication has constrained the controller to a region
close to the point of operation.
In order to design controllers that can be applied to a broad range
of operation, the use of rst-principles simulation models have been
suggested in the literature [10] for the dynamic identication of vapor
compression refrigeration systems. Albeit there are several publications concerning the modeling for controlling vapor compression
refrigeration systems [11e16], all of them conducted the system
identication and model validation exercises for regions close to the
point of operation (w5%).
Therefore, in a prior study, Schurt et al. [17] put forward
a multivariable model-driven controller for vapor compression
refrigeration systems. Mathematical sub-models were developed for
each of the system components: heat exchangers (condenser and
evaporator), variable-speed compressor and variable-orice

x
y

1539

dynamic state matrix


output matrix

Greek

a
3ll-sl
g
l
r

heat transfer coefcient, W/m2 K


internal heat exchanger effectiveness
void fraction
boundary position, m
specic mass, kg/m

Subscripts
c
condenser
e
evaporator, static
i
inlet, integral
K
Kalman
l
saturated liquid
o
outlet
r
refrigerant
ref
reference
s
secondary coolant, isentropic process
sur
surroundings
v
saturated vapor
Other symbols
hxi
volume average of x
x_ dx=dt time derivative of x

expansion valve. The dynamic simulation model was then used to


design a MIMO (multi-input, multi-output) controller based on the
linear-quadratic-Gaussian (LQG) technique using a Kalman lter for
the estimator design. The purpose-built testing apparatus constructed by Pottker and Melo [4] was then used to collect data for the
system identication and model validation exercises. It was found
that the model reproduces the experimental trends of the working
pressures and power consumption in conditions far from the operation point (30%) with a maximum deviation of 5%. However,
Schurt et al. [17] have not explored the controlling envelope of the
MIMO control system, which is, therefore, the main focus of the
present study.
2. Dynamic simulation model
Modeling of the refrigeration system relies on the development
of sub-models for each of the cycle components (see Fig. 1), which
have followed both dynamic (refrigerant ow through the heat
exchangers) and quasi-steady (compressor, expansion valve, internal
heat exchanger, and condensing and evaporating media) modeling
approaches. Further details can be found in Schurt et al. [17].
2.1. Heat exchangers
The model for the refrigerant ow through the condenser and
evaporator was based on the following key assumptions: (i) 1-D
ow; (ii) straight, horizontal and constant cross-sectional coil; (iii)
negligible diffusion effects; (iv) negligible pressure drop; and (v)
negligible thermal inertia of the walls. The governing equations,
derived from the mass and energy conservation principles, are
represented by

v
1 v
rf
mf S
vt
A vz

(1)

1540

L.C. Schurt et al. / Applied Thermal Engineering 30 (2010) 1538e1546

Fig. 1. Schematic representation of the test rig (Pottker and Melo [4]).

where f 1 and S 0 retrieve the conservation of mass, whereas


f h and S dp/dt 4q/D retrieve the conservation of energy,
where q is the heat ux [W/K] and D the coil inner diameter [m]. To
be solved Eq. (1) must be integrated in both time and space
domains. The latter was performed following the so-called movingboundaries approach [11], according to which the spatial domain is
divided into two zones namely single and two-phase ow regions,
as illustrated in Fig. 2. Since the interface position varies with time,
the Leibnitz's rule has to be invoked for the integration of Eq. (1),
l2 t
Z

l1 t

supplying a heat ux with the same magnitude of that calculated


assuming two-phase refrigerant [17].
The averaged terms of Eq. (2) were approximated by a 2nd order
scheme (single-phase zone) and by the mean void fraction (twophase zone), respectively:

v
d
rf dz l2  l1 hrf i
vt
dt
dl
dl


hrf i  rl2 fl2 2  hrf i  rl1 fl1 1
dt
dt

(3)

hrf i rv fv hgi rl fl 1  hgi

(4)

The condenser and evaporator mean void fraction hgi were


tted to experimental data as a part of the system identication
exercise [17]. The spatial integration of the mass and energy
conservation equations (1) in each ow region represented in Fig. 2
provided three ordinary differential equations that were re-organized in the form of a linear set of differential equations:

where l1 0 and l2 l(t) for the two-phase region, whereas l1 l(t)


and l2 L for the single-phase region (i.e., evaporator superheating
and condenser subcooling). It is worth noting that the condenser
gascooling region was modeled as if it was saturated since the
vapor provides a lower heat transfer coefcient, but at higher logmean temperature difference which compensate each other, thus


Fx; u; w A1 mi  mo

1
r fo rl fl
2 o

hrf i

Cx$x_ Fx; u; w

(5)

where x [p l ho]T, u [mi mo]T, w [Qtp Qsp]T, and F(x, u, w) is


obtained from

mi hi  hl Qtp

mo hl  ho Qsp

T

(6)

Qsp

Qtp
mi, hi

m,, h

(t)
L

Fig. 2. Schematic representation of the spatial discretization of the heat exchangers.

mo, ho

L.C. Schurt et al. / Applied Thermal Engineering 30 (2010) 1538e1546

1541

The elements of matrix C(x) are shown in Table 1. It is noteworthy


that, l v for the evaporator, l l for the condenser, and the
superscripts ()0 and ()00 represent the pressure and enthalpy derivatives, respectively. Moreover, it is worth noting that mi meev and
mo mvsc for the evaporator, whereas mi mvsc and mo meev for
the condenser. The heat transfer rates were calculated according to
the 3-NTU approach presented in [17].

the evaporator inlet, improving the cooling capacity and avoiding


compressor slugging. The internal heat exchanger sub-model
determines the states of the refrigerant at the compressor (1) and
valve (4) inlet sections (see Fig. 1) based on the information available at the condenser (3) and evaporator (6) exits (see Fig. 1),
respectively. On one hand, the compressor inlet temperature T1 was
obtained from the temperature effectiveness:

2.2. Compressor

T1 3llsl T3 1  3llsl T6

(12)

h1 hv Te cp;v Te T1  Te

(13)

The compressor sub-model provides the refrigerant mass ow


rate suctioned from the evaporator and discharged to the
condenser, and the power consumed by the compression process
[18], being respectively calculated from

mvsc



St  cSt pc =pe cv;v =cp;v 1 N=v1

(7)

Wvsc a b$mvsc h2s  h1

h4 h3 h6  h1

(14)

T4 Tc h4  hl pc =cp;l pc

(15)

(8)

where cp,v and cv,v were calculated for the saturated vapor at the
evaporating pressure, a, b, c and St were tted to experimental data
as a part of the model identication exercise [17], and T2s was
obtained from

T2s Tc h2s  hv pc =cp;v pc

(9)

where h2s h(pc,s1) and s1 s(pe,h1). The refrigerant enthalpy at


the compressor discharge was obtained from the following energy
balance:

h2 h1 Wvsc  UAvsc T2s  Tsur =mvsc

(10)
 C),

where Tsur is the surrounding air temperature (25


and UAvsc is
the conductance of the compressor obtained empirically from the
system identication exercise [17].
2.3. Expansion valve
The refrigerant mass ow rate through the expansion valve,
meev, was calculated from the orice equation as follows

meev

On the other hand, the refrigerant enthalpy at the evaporator


inlet h4 was obtained from the following energy balance:

p
Ceev Ae 2$r4 $pc  pe

(11)

where meev is given in [kg/s], Ae AoAv/100 is the effective passage


ow area, Av is the valve opening [%], Ao is the nominal orice crosssectional area (0.1238 cm2), r4 is the specic mass at the valve inlet,
and Ceev is the discharge coefcient calculated as a power-law function of the subcooling degree, Ceev d(T3  Tc)e, where the constants
d and e were obtained from the system identication exercise [17].
2.4. Liquid line suction line heat exchanger
The liquid line/suction line heat exchanger (also named internal
heat exchanger) aims to increase the amount of liquid refrigerant at
Table 1
Elements of matrix C(x).
c11

lhgir0v 1  hgir0l 1=2L  lr0l r0o

c12

hgirv 1  hgirl  1=2rl  ro

c13

1=2L  lr00o

c21

lhgirv h0v r0v hv  hl 1  hgirl h0l r0l hl  hl  1

c22

hgirv hv  hl 1  hgirl hl  hl

c23
c31

0
1=2L  lrl h0l  hl r0o  2

c32

1=2ro hl  ho

c33

1=2L  lro r00o ho  hl

2.5. Solution algorithm


The overall system simulation relies on the time integration of
an equation set comprised of six ordinary differential equations.
The initial conditions are the evaporating and condensing pressures, the refrigerant enthalpies at the evaporator and condenser
outlet sections, and the positions of the evaporation and condensation boundaries, all obtained from the prior steady-state solution
of the equation set enforcing the time-derivatives to be nil. The
boundary conditions are the compressor speed, the valve opening,
and the temperatures and ow rates of the secondary coolants, all
of them obtained from experimental data. The numerical integration of the differential equations was carried out through the
ODE23S adaptive method for a stiff set of equations [19]. The
thermodynamic properties were obtained from REFPROP7 [20].
3. Controller design
In order to simplify the controller design, a linear state-space
representation was adopted, which required the linearization of the
dynamic simulation model using a Taylor expansion series, yielding

dx_ A$dx B$du W$dw


dy C$dx E$dw

(16)

The vectors dx, dy, du and dw indicate respectively the difference


between the static (x0, y0, u0, w0) and the dynamic values of the state
variables, x [pe le ho,e pc lc ho,c]T, the controlled variables, y [DTsup
Ts,e,o]T, the input variables, u [N Av]T, and the perturbations, w [Vc
Ve]T. The matrices A, B, W, C and E were obtained from the model
linearization exercise and their values are summarized in [17].
The MIMO controller was designed based on a LQG scheme with
an integrator, as illustrated in Fig. 3. The controller design was
focused not only on matching the cooling capacity to the thermal
loads but also on keeping the evaporator superheating at a predened level by driving the compressor speed and the electric valve
opening simultaneously. In Fig. 3, yref [DTsup,ref Ts,e,o,ref]T is the
reference input signal, x_ yref  y is the tracking error, x is the
output of the integrator, xK is the estimated state, u Ke$xK Ki$x
is the control signal, and Ke and Ki are the state-feedback and integral gain matrices, respectively.
The LQG controller consists of a combination of a linearquadratic-regulator (LQR) and an optimal state estimator, which
have been developed independently according to the separation
principle [21]. The former consists of nding the matrix K [KejKi]

L.C. Schurt et al. / Applied Thermal Engineering 30 (2010) 1538e1546

Ki
-

++

x = Ax + Bu
y = Cx

Ke

xK

x
y

Kalman
Filter

Fig. 3. Schematic representation of the controller.

of the control action, u(t) K$xa(t), where xa [xK x]T, that


minimizes the following quadratic criterion:

ZNh

i
xTa t$Q $xTa t uT t$R$ut dt

5. Results and discussion

(17)

where Q and R are positive-denite Hermitian matrices determined through the Bryson method [22]. It is noteworthy that Q and
R lead to the elements of matrix K straightforwardly through the
positive-denite-solution P of the Riccati equation [17].
A discrete-time version of the LQG controller was designed and
the controller was then implemented on a digital computer with
a sample time of 2 s. It is worth of note that the controller was
designed to reject perturbations due to the integrator, and also to
keep the time-response of the closed-loop system approximately the
same that of the open-loop system. The values of matrices Q, R and K
of the discrete-time LQ regulator are summarized in [17].
Since some states cannot be measured directly (e.g., position of
the evaporation and condensation boundaries), a state estimator
was also required. The estimator was developed using a Kalman
lter as indicated below,

x_ K t A  L$C$xK t B$ut L$yt

Moreover, the compression power was measured with an uncertainty of 0.5% (full scale). For steady-state conditions, the difference
between the refrigerant-side and the secondary uid-side heat
transfer rates was less than 2%. A control and data acquisition system
was used for monitoring the experimental variables and for setting
the compressor speed and valve opening. Further details of the
experimental apparatus can be found in [4].
The facility was used to collect data for the system identication
and model validation exercises. It was found that the model reproduces the experimental trends of the working pressures and power
consumption in conditions far from the operation point (30%) with
a maximum deviation of 5%, as illustrated in Fig. 4.

The experimental analyzes were carried out to assess the


controller performance in terms of reference tracking, controlling
envelope, and disturbance rejection. The reference tracking test was
conducted imposing step changes to the controlled variables (i.e.,
references), for instance the evaporator superheating and the brine
temperature at the evaporator outlet. The controlling envelope was
evaluated by changing the references until the maximum and
minimum controlled states were reached. The disturbance rejection
was investigated by varying the ow rate of the evaporating and
condensing media.
5.1. Reference tracking and disturbance analysis
The test was initiated when the steady-state condition was achieved for the following operating conditions: compressor speed of
3000 rpm, valve opening of 45%, evaporating medium ow rate and
temperature of 1.24 L/min and 10  C, respectively, and condensing
medium ow rate and temperature of 1.17 L/min and 35  C,

(18)

where L and S are calculated according to the procedure described


in [17].
4. Experimental facility
Fig. 1 shows a schematic diagram of the HFC-134a refrigeration
system used in this study. The compressor is of hermetic variablespeed reciprocating type, with minimum and maximum speeds of
1800 and 4500 rpm, respectively. The condenser and evaporator are
both tube-in-tube heat exchangers, connected to two distinct
temperature and ow controlled secondary heat transfer loops.
Brine of 27% ethylene-glycol aqueous solution is used for the evaporator, whereas pure water is used for the condenser. The expansion
device is a pulse-width-modulation (PWM) electric expansion
device with variable opening from 0 to 100%. The condenser and
evaporator heat transfer rates are obtained from energy balances on
the refrigerant and secondary uid sides. The refrigerant heat
transfer rate was calculated by multiplying the refrigerant mass ow
rate, measured by a Coriolis mass ow meter (measurement
uncertainty of 0.1%) by the enthalpy difference between the heat
exchanger inlet and outlet, which provided a maximum measurement uncertainty of 2% [4].
Immersion T-type thermocouples (measurement uncertainty of
0.2  C) and pressure transducers (measurement uncertainty of
0.15%) were installed at selected points of the circuit to obtain the
refrigerant enthalpies and the evaporator superheating. The
secondary uid heat transfer rate was calculated using the volumetric ow rate, measured by a turbine ow meter (measurement
uncertainty of 1%), the uid density, the specic heat and the
temperature difference between the heat exchanger inlet and outlet.

a 3.0
Evaporating pressure [bar]

yref +

Experimental
Simulation

2.5

2.0

1.5

1.0

b 16
Condensing pressure [bar]

1542

50

100

150

200
250
Time [min]

300

350

400

150

200
250
Time [min]

300

350

400

Experimental
Simulation

15

14

13

12

11

50

100

Fig. 4. Model validation exercise: (a) evaporating pressure; and (b) condensing
pressure.

L.C. Schurt et al. / Applied Thermal Engineering 30 (2010) 1538e1546

1543

Table 2
Changes imposed to the evaporator superheating and brine outlet temperature during the reference tracking exercise.
Time [min]

t < 20

20 < t < 40

40 < t < 60

60 < t < 80

80 < t < 100

100 < t < 120

t > 120

DTsup [ C]

16.6
4

20.2
4

12.5
4

16.6
4

16.6
5

16.6
3

16.6
4

Ts,e,o [ C]

respectively. At this condition, an evaporator superheating of 16.6  C


and a brine temperature at the evaporator outlet of 4  C were achieved and adopted as references. Independent changes in the references were imposed following the sequence shown in Table 2. The
dynamic behavior of the controlled variables is depicted in Fig. 5a,
where it can be noted that the controller drove the refrigeration
system towards the prescribed references. It is noteworthy that
a reference change in one of the controlled variables produced
a disturbance in its counterpart, which was identied and properly
rejected by the controller.
The highest disturbance level was observed during the second
change in the evaporator superheating degree, at t 40 min, reaching
a minimum value of 10.6  C, which represents an undershooting of
2  C. Such an effect was induced by the magnitude of the imposed
change (7.7  C), negligible in the rst and third changes of references
(at t 20 min and t 60 min, respectively). The disturbances caused
by changing the reference of the brine outlet temperature were also
properly rejected by the controller, which affected the evaporator
superheating in 2  C only (at t 120 min).
For the brine outlet temperature, the controller showed a better
performance, with a maximum undershooting of 0.5  C at the
instant when the highest change of reference was observed, i.e.,
2  C at t 100 min. Furthermore, the controller kept the brine
outlet temperature within a 0.5  C band during the rst 80 min.
Fig. 5b shows the variations experienced by the control actions
(i.e., compressor speed and valve opening) in order to drive the
controlled variables towards the desired references. It is worthy
noting that, during the change of reference of the evaporator
superheating (t < 40 min), the control actions presented opposite
behaviors: the compressor speed increased when the valve
opening decreased. This was so as the cooling capacity tends to
decrease inasmuch as the superheating degree increases, forcing
the controller to increase the compressor speed in an attempt to
keep the cooling capacity (see Fig. 5d), thereby reducing the
evaporating pressure (see Fig. 5c) and increasing the evaporator
log-mean temperature difference. The opposite behavior was
observed within the time interval 40 < t < 60 min.
During the change of reference of the brine outlet temperature
(t > 60 min), it was noted that both compressor speed and valve
opening showed the same behavior (see Fig. 5b). On one hand, to
raise the brine outlet temperature (i.e., lower cooling capacity)
without changing the evaporator superheating (t 80 min), the
controller increased the evaporating pressure (see Fig. 5c) by
reducing both the compressor speed and the expansion valve
opening so as to diminish the refrigerant mass ow rate. On the other
hand, the cooling capacity increased at t 100 min as the compressor
speed and the valve opening were simultaneously brought up (see
Fig. 5b).
The disturbance rejection test aims to check whether the
controller is able to keep the controlled variables at the reference
previously set, even if changes are imposed on the thermal load or
the surrounding conditions. Such changes have been respectively
emulated in the experimental apparatus by modifying the evaporator and condenser secondary coolant ow rates, as shown in
Table 3. Such variations have induced disturbances on the
controlled variables that have been recognized by the controller
and then rejected, as can be seen in Fig. 6a. The largest deviation
observed in the evaporator superheating, in comparison to its

Fig. 5. Reference tracking analysis: (a) controlled variables; (b) control actions;
(c) working pressures; and (d) cooling capacity.

1544

L.C. Schurt et al. / Applied Thermal Engineering 30 (2010) 1538e1546

Table 3
Changes imposed to the evaporator and condenser secondary coolant mass ow rates during the disturbance rejection exercise.
Time [min]

t < 20

20 < t < 40

40 < t < 60

60 < t < 80

80 < t < 100

100 < t < 120

t > 120

Vs,c [L/min]
Vs,e [L/min]

1.16
1.23

1.16
1.43

1.16
1.00

1.16
1.23

1.38
1.23

0.93
1.23

1.16
1.23

reference, was 3  C at t 40 min. Such deviation is related to the


sudden reduction of the brine ow rate from 1.43 to 1.00 L/min,
which represented a 30% reduction in the thermal load imposed
(from 550 to 385 W, see Fig. 6d).
It can be also noted in Fig. 6 that the variations observed for
t < 70 min are related to the thermal load disturbances, as the
changes in the condenser side had no practical inuence on the
controlled variables and control action, such as the compressor
speed and the valve opening (Fig. 6b), the mass ow rate (Fig. 6d),
and evaporating temperature (Fig. 6c). The condensing pressure, on
the other hand, has experienced some disturbances, although the
largest variation was due to the changes on the evaporator brine ow
rate (Fig. 6c). Albeit the cooling capacity (see Fig. 6d) was practically
not affected by the disturbance imposed in the condenser secondary
coolant ow rate (t > 70 min), it changed substantially with the
evaporator brine ow rate (t < 70 min).
5.2. Assessment of the controlling envelope
The operational envelope beyond which it is not possible to
maintain the controlled variables at their references was evaluated
by articially driven the controller through step changes promoted
in the references. In addition, changes in the evaporator brine ow
rate were also imposed to the system in order to identify the
maximum and minimum controllable thermal loads.
First, increasing step changes (from 4.0 to 6.0  C, step of 0.5  C) on
the brine outlet temperature were imposed, as shown in Fig. 7a,

Fig. 6. Disturbance rejection analysis: (a) controlled variables; (b) control actions;
(c) working pressures; and (d) cooling capacity.

Fig. 7. Controlling envelope analysis with increasing brine temperature: (a) controlled
variables; (b) control actions.

L.C. Schurt et al. / Applied Thermal Engineering 30 (2010) 1538e1546

1545

as shown in Fig. 7b. When the minimum compressor speed was


reached, the evaporator superheating increased as the valve opening
ruled the system while the compressor speed became saturated.
Similarly, decreasing step changes (from 4.0 to 2.5  C, step of 0.5  C)
on brine outlet temperature were imposed, as shown in Fig. 8a,
which is equivalent to a thermal load augmentation. At each new
reference change, the controller increased the cooling capacity by
opening the valve and raising the compressor speed up to the
maximum allowed value (4500 rpm), as shown in Fig. 8b. As the
evaporator superheating was xed at 16.6  C, the brine outlet
temperature was limited within the range from 3 to 5.5  C.
Alternatively, decreasing step changes (from 16.6 to 8.5  C) on the
evaporator superheating were imposed, as shown in Fig. 9a, until the
controller is no longer maintaining the controlled variables at their
references. It is noteworthy that at superheating degrees lower than
8.5  C the system became unstable due to two factors. First, as the
proposed controller is linear, it cannot identify the non-linear
dynamics of the variables (see Fig. 9b). In addition, the evaporator
dry-out location oscillates at lower superheating degrees due to the
natural two-phase ow instabilities [11], which affected the quality
of the control signal and therefore the control action.

a
Fig. 8. Controlling envelope analysis with decreasing brine temperature: (a) controlled
variables; (b) control actions.

which is equivalent to a thermal load reduction. In order to follow


the references and also to adapt the system to lower cooling
demands, the controller decreased the valve opening while reduced
the compressor speed until 1800 rpm, the minimum allowed speed,

Fig. 9. Controlling envelope analysis with decreasing evaporator superheating:


(a) controlled variables; (b) control actions.

Fig. 10. Controlling envelope analysis with increasing and decreasing thermal load:
(a) controlled variables; (b) control actions; and (c) disturbances.

1546

L.C. Schurt et al. / Applied Thermal Engineering 30 (2010) 1538e1546

Additional tests were carried out varying the evaporator brine


ow rate, but keeping the evaporator superheating degree and the
brine outlet temperature references xed (see Fig. 10a). During the
rst 120 min, the ow rate was steeply increased to the maximum
value that can be achieved without loosing the system controllability,
as showed in Fig.10b. Therefore, the refrigeration system experienced
a thermal load variation from 460 to 580 W (see Fig. 10c). In other
words, a thermal load 26% higher than the nominal was achieved but
keeping the preset references (see Fig. 10b). Between 120 and
140 min, the system was then returned to its initial condition. After
140 min, the brine ow rate was steeply decreased in order to reduce
the thermal load to 340 W, i.e., a gure 26% lower than the nominal
value (see Fig. 10c).
6. Summary and conclusions
A dynamic simulation model for the identication and control of
vapor compression refrigeration systems was developed and validated for a broad range of operation. A breadboard refrigeration
system comprised of a variable-speed compressor and an electric
expansion valve was used to gather the required experimental data
for both system identication and model validation exercises. It
was found that the model reproduces well the experimental trends
of the working pressures even in conditions far from the operation
point used in the system identication, with maximum discrepancies between calculated and measured counterparts within 5%
error bands.
A model-driven multivariate linear strategy for controlling the
evaporator feeding and for matching the cooling capacity to the
thermal loads was devised. The dynamic simulation model was
linearized according to a Taylor expansion series and then used to
design a proportional plus integral type controller based on the
linear-quadratic-Gaussian (LQG) method, which is based on a state
estimator of the Kalman lter type. Additional experimental tests
were also carried out using the breadboard facility to verify the
controller ability for tracking reference changes and rejecting
thermal load disturbances due to the integral action of the controller.
It was found that the controller was able to drive the facility for
thermal loads spanning from 340 to 580 W, i.e., 26% in comparison to the nominal value. Similarly, the controller kept the refrigeration system running properly for evaporator superheating
degrees ranging from 9.5  C to 22  C, albeit it was designed for
16.6  C. However, it is worthy of note that the controller cannot be
applied for superheating degrees lower than 9.5  C, when both the
controlled system and the control signal became unstable.
Acknowledgements
The authors would like to thank the CNPq (Grant No. 573581/
2008-8 e National Institute of Science and Technology in Refrigeration and Thermophysics) for the nancial support. Thanks are
also addressed to Embraco S.A. for sponsoring this research project.
The authors duly acknowledge Prof. C. Melo and Prof. N. Roqueiro

(Federal University of Santa Catarina), and Dr. L. W. da Silva


(Embraco S.A.) for their valuable technical contributions.
References
[1] S.A. Tassou, T.Q. Qureshi, Comparative performance evaluation of positive
displacement compressors in variable-speed refrigeration applications. Int. J.
Refrigeration 21 (1998) 29e41.
[2] C. Aprea, R. Mastrullo, Experimental evaluation of electronic and thermostatic
expansion valves performances using R22 and R407C. Appl. Therm. Eng. 22
(2002) 206e218.
[3] S.A. Tassou, H. Al-Nizari, Investigation of the steady state and transient
performance of reciprocating chiller equipped with an electronic expansion
valve. Heat Recov Syst CHP 11 (1991) 541e550.
[4] G. Pottker, C. Melo, A study on the relationship between compressor speed
and expansion valve opening in refrigeration systems, in: Conf. on compressors and their systems, London, UK, Paper C658, 2007.
[5] A. Outtagarts, P. Haperschill, M. Lallemand, The transient response of an
evaporator fed through an electronic expansion valve. Int. J. Energy Res. 21
(1997) 793e807.
[6] J.M. Choi, Y.C. Kim, Capacity modulation of an inverter-driven multi-air
conditioner using electronic expansion valve. Energy 28 (2003)
141e155.
[7] D.S. Yang, G. Lee, M.S. Kim, Y.J. Hwang, B.Y. Chung, A study on the capacity
control of a variable speed vapor compression system using superheat information at compressor discharge, in: 10th Int. Refrig. Conf. at Purdue, West
Lafayette-IN, USA, Paper R-164, 2004.
[8] B. Lamanna, Development of an advanced control system for chillers with
inverter driven scroll compressor and comparison with traditional on/off
systems, in: IIR conference on commercial refrigeration, Vicenza, Italy, 2005,
pp. 295e302.
[9] J.B. Marcinichen, T. N. Holanda, C. Melo, A dual SISO controller for a vapor
compression refrigeration system, in: 12th Int. Refrig. Conf. at Purdue, West
Lafayette-IN, USA, Paper 2444, 2008.
[10] B. Rasmussen, A. Alleyne, Control-oriented modeling of transcritical vapor
compression systems. J. Dyn. Syst. Meas. Contr. 126 (2004) 54e64.
[11] G.L. Wedekind, W.F. Stoecker, Theoretical model for predicting the transient
response of the mixture-vapour transition point in horizontal evaporating
ow. ASME J. Heat Trans. (February 1968) 165e174.
[12] X. He, S. Liu, H. Asada, Modeling of vapor compression cycles for multivariable
feedback control of HVAC systems. J. Dyn. Syst. Meas. Contr. 119 (1997)
183e191.
[13] M. Willatzen, N.B.O.L. Pettit, L. Ploug-Sorensen, A general dynamic simulation
model for evaporators and condensers in refrigeration. Part I: movingboundary formulation. Int. J. Refrigeration 21 (1998) 398e403.
[14] D. Leducq, J. Guilpart, G. Trystram, Non-linear predictive control of a vapour
compression cycle. Int. J. Refrigeration 29 (2006) 761e772.
[15] T.L. Mckinley, A.G. Alleyne, An advanced nonlinear switched heat cycles
using the moving-boundary method. Int. J. Refrigeration 31 (2008)
1253e1264.
[16] Q. Qi, S. Deng, Multivariable control-oriented modeling of a direct expansion (DX) air conditioning (A/C) system. Int. J. Refrigeration 31 (2008)
841e849.
[17] L.C. Schurt, C.J.L. Hermes, A. Trono Neto, A model-driven multivariable
controller for vapor compression refrigeration systems. Int. J. Refrigeration 32
(2009) 1672e1682.
[18] W.C. Gosney, Principles of Refrigeration. Cambridge University Press,
Cambridge, UK, 1982.
[19] L.F. Shampine, M.W. Reichelt, The Matlab ODE Suite. SIAM J. Sci. Comput. 18
(1997) 1e22.
[20] E.W. Lemmon, M.O. McLinden, M.L. Huber, REFPROP 7.0, Standard Reference
Database 23. National Institute of Standards and Technology, Gaithersburg,
MD, USA, 2002.
[21] U. Mackenroth, Robust Control Systems. Springer Verlag, Berlin, Germany,
2004.
[22] M.J. Johnson, M.A. Grimble, Recent trends in linear optimal quadratic multivariable control system design. IEE Proceedings Part D Control Theory and
Applications 134 (1987) 53e71.

You might also like