Вы находитесь на странице: 1из 23

VOLUME 15, NUMBER 1 HVAC&R RESEARCH JANUARY 2009

77
Fault Detection and Diagnostics
for Commercial Coolers and Freezers
Adam Wichman James E. Braun, PhD
Student Member ASHRAE Fellow ASHRAE
Received December 29, 2007; accepted April 16, 2008
Most of the previous diagnostics research for vapor-compression equipment has been per-
formed for packaged air conditioners and chillers. This paper describes application of a
decoupling-based diagnostic technique that was originally developed for air conditioners to
equipment used in walk-in coolers and freezers. Testing was performed on a small, restau-
rant-style walk-in cooler and a small walk-in freezer for faulted and unfaulted conditions. The
data was then used to characterize how the fault features respond to faults and to evaluate the
performance of the decoupling-based diagnostic technique.
INTRODUCTION
In general, HVAC&R systems are not well maintained (Proctor and Downey 1995; Cowan
2004; Li and Braun 2006) because of the relatively high cost of service and low cost of energy.
Recently, there has been a growing interest in the development of automated fault detection and
diagnostics (FDD) for HVAC&R equipment. For vapor-compression cooling equipment, most of
the methods presented in the literature (Grimmelius et al. 1995; Stylianou and Lau 1996; Rossi and
Braun 1997) utilize differences between measurements and model predictions (residuals) of state
variables to perform FDD. Although these methods have good performance for individual faults
(Breuker and Braun 1998; Li and Braun 2003), they do not handle multiple-simultaneous faults. In
addition, these methods require measurements over a wide range of conditions for training refer-
ence models, the development of which can be time consuming and cost prohibitive.
Recently, a diagnostic method was developed that handles multiple-simultaneous faults (Li
and Braun 2007a) through the use of decoupling features (Li and Braun 2007b). Decoupling fea-
tures are parameters that are uniquely influenced by individual faults and are insensitive to vari-
ations in ambient conditions. For example, air mass flow rate through the condenser is a feature
that is strongly influenced by the level of fouling and condenser fan problems but is nearly inde-
pendent of other faults that can occur for an air-conditioning system that incorporates
fixed-speed fans.
In developing decoupling features, it is important to utilize low-cost sensors, such as tem-
perature sensors (see Table 4 for a summary of the temperature measurements required to per-
form the diagnostics presented in this paper). These low-cost measurements are used in simple
models as virtual sensors to infer other system measurements. For example, as described by Li
and Braun (2007b), condenser airflow can be estimated using an energy balance with air-side
and refrigerant-side measurements. For this energy balance, the refrigerant flow is estimated
using a compressor map as a virtual sensor. Furthermore, virtual evaporating and condensing
pressure sensors utilize surface-mounted temperature measurements at locations where satu-
rated conditions exist and property relations to estimate saturation pressures.
Adam Wichman is a reactor engineer at PSEG Nuclear LLC, Hancocks Bridge, NJ. James E. Braun is a professor in
the School of Mechanical Engineering, Purdue University, West Lafayette, IN.
2009, American Society of Heating, Refrigerating and Air-Conditioning Engineers, Inc. (www.ashrae.org). Published in HVAC&R Research, Vol. 15,
No. 1, January 2009. For personal use only. Additional reproduction, distribution, or transmission in either print or digital form is not permitted
without ASHRAEs prior written permission.
78 HVAC&R RESEARCH
The virtual and physical measurements are used to determine the decoupling features for
diagnostic purposes. When a decoupling feature deviates significantly from its normal value, a
fault is indicated (e.g., low condenser airflow for fouling or fan problems).
The decoupling-based FDD method was originally developed for air-conditioning (AC) sys-
tems. Although the vapor-compression equipment used for commercial refrigerators and freezers
is very similar to that used for air conditioning, operation occurs over a different range of temper-
atures and the systems utilize different refrigerants. In addition, commercial refrigeration equip-
ment typically utilizes liquid-line receivers that are not generally employed for air conditioners.
In this paper, the decoupling-based FDD method was applied to equipment used in
small-scale walk-in coolers and freezers. Faults were artificially introduced in the laboratory,
and the performance of the diagnostic method was evaluated.
WALK-IN COOLER AND FREEZER EXPERIMENTS
Walk-in cooler and freezer units were tested within psychrometric chambers to allow control
of the condenser air inlet conditions. Figure 1 shows a cooler unit and walk-in cabinet. The
walk-in cooler and freezer experiments utilized the same refrigerated space but employed differ-
ent refrigeration equipment. The cooler system utilized R-22 as the refrigerant, whereas the
freezer unit employed R-404A. Both systems were equipped with a thermal expansion valve
(TXV) and a liquid-line receiver. The receiver provides a volume for liquid refrigerant to collect
after it exits the condenser. This volume keeps the refrigerant exiting the condenser in a saturated
liquid state during operation, unless the unit is overcharged to the point where the receiver is
completely filled with liquid refrigerant. When the receiver is full, then additional charge added
to the system will back up in the condenser and lead to subcooling of the refrigerant exiting the
condenser under steady operating conditions. The liquid-line receiver is shown in Figure 2.
Figure 3 shows a schematic of the refrigerant cycle depicting the methods used for simulating
faults and the measurements taken. Table 1 provides the list of faults along with a description of
the fault simulation approaches. Refrigerant undercharge, refrigerant overcharge, liquid-line
restriction, compressor valve leakage, condenser fouling, and evaporator fouling were consid-
Figure 1. The walk-in cooler inside the psychrometric chamber.
VOLUME 15, NUMBER 1, JANUARY 2009 79
Table 1. Fault Simulation Methods
Fault Type Simulation Method
Compressor valve leakage Partially open bypass valve
Condenser fouling Slow compressor fan using installed speed controller
Evaporator fouling Slow evaporator fan using installed speed controller
Liquid-line restriction Partially close liquid-line restriction valve
Refrigerant undercharge Purposely undercharge the system
Refrigerant overcharge Purposely overcharge the system
Figure 2. The liquid-line receiver after the condenser.
Figure 3. Cooler cycle diagram with added instrumentation.
80 HVAC&R RESEARCH
ered. A leaky compressor valve allows high-pressure refrigerant to flow back to the low-pressure
side of the system, which lowers the volumetric efficiency of the compressor and lowers the mass
flow rate of the system. To simulate a leaky compressor valve, a bypass line with a flow control
valve was added around the compressor that can be opened to allow refrigerant to flow from the
high- pressure side of the system to the low-pressure side. Heat exchanger fouling in vapor-com-
pression equipment can be characterized as a decrease in airflow across the coils. Previous work
(Pak et al. 2005; Yang et al. 2007) has demonstrated that the primary effect of air-side heat
exchanger fouling is an increased pressure drop leading to a reduced airflow rate. Fouling was
simulated for both the condenser and evaporator by connecting fan speed controllers to control
the airflow rates across the heat exchangers. During operation, a vapor-compression system can
experience clogging of the filter dryer, which restricts the flow through the liquid line. In a sys-
tem with a fixed-orifice expansion device, a liquid-line restriction can lead to a reduced mass
flow rate. The walk-in cooler tested uses a TXV, which can compensate for the pressure drop
incurred for moderate restrictions and keep the refrigerant flow rate relatively constant. For
severe restrictions, the TXV will saturate, opening fully, and act like a fixed orifice. A flow con-
trol valve was added to the liquid line, which was partially closed to simulate this fault.
The instrumentation depicted in Figure 3 includes condenser inlet and outlet air temperatures
(T
cai
and T
cao
) and evaporator inlet and outlet air temperatures (T
eai
and T
eao
) that were mea-
sured using thermocouple grids. Refrigerant pressures and temperatures were measured at the
inlets and outlets of all components to accurately determine the refrigerant states. To understand
where the two-phase regions in both the condenser and evaporator are located, eight thermocou-
ples were soldered to the return bends of both the condenser (T
cond,1-8
) and the evaporator
(T
evap,1-8
) and insulated from the surrounding air to obtain an indication of the refrigerant tem-
peratures inside the tubes. Compressor power and refrigerant mass flow rate were also measured
in order to determine system cooling capacity and efficiency.
Table 2 provides a description of the test matrix employed for the cooler. For the cooler, each
individual fault was simulated at three different ambient conditions (T
cai
). Five fault levels were
tested and characterized by percent of nominal cooling capacity. The nominal cooling capacity
was determined in the unfaulted condition or 0th fault level test. Fault levels were chosen so that
the degradation in cooling capacity would occur in even increments with the maximum degrada-
tion depending on the fault type. The compressor valve leakage and liquid-line restriction faults
degraded cooling capacity by more than 50%. Undercharging the system degraded capacity by
around 20%. Slowing the heat exchanger fans over the test range degraded the capacity by only
10%. Overcharging the refrigerant had very little effect because of the presence of a receiver.
Four combinations of multiple faults were tested for the cooler. For these tests, the system was
undercharged by 50%, and four other faults were applied to the system, all at maximum fault
levels defined by the single-fault tests.
Based on knowledge gained during testing of the walk-in cooler, a smaller test matrix was
employed for the freezer; it is described in Table 3. Only one ambient condition was tested. The
overcharging tests on the cooler showed that a liquid-line receiver prevents overcharging from
being a fault that affects the performance of the system. Therefore, no overcharging tests were
performed on the freezer. Multiple fault tests were run on the freezer combining evaporator foul-
ing with the other four faults. One of the concerns with a freezer is ice accumulation and distin-
guishing it from evaporator fouling. To gather data on this phenomenon, an ice accumulation
test was performed.
Except for the ice accumulation test, all of the data were obtained at steady-state conditions.
Heaters were controlled within the cooled spaced in order to add the necessary load to achieve
steady-state indoor conditions of 37.4F. Data were collected and averaged over 10- to 15-minute
intervals with the unit operating at a steady-state condition.
VOLUME 15, NUMBER 1, JANUARY 2009 81
DECOUPLING FEATURES
Compressor Valve Leakage
Compressor valve leakage is meant to include any fault within the compressor that leads to a
loss in volumetric efficiency and refrigerant flow rate. As this fault develops, the volumetric
efficiency of the compressor decreases. This decreases the refrigerant mass flow rate, which in
turn decreases the discharge enthalpy. In order to detect this type of fault, a simple model is used
to estimate the nominal discharge temperature (T
dis
) from the compressor. This expected value
is then compared to the measured T
dis
. The normal discharge enthalpy is calculated as
, (1)
Table 2. Single- and Multiple-Fault Test Matrix for the Walk-In Cooler
Fault
Ambient
Temperatures
Fault
Levels
Capacity
Degradation
Tests
Compressor valve leakage 55F, 75F, 95F 05 0%55% 18
Liquid-line restriction 55F, 75F, 95F 05 0%55% 18
Condenser fouling 55F, 75F, 95F 05 0%9% 18
Evaporator fouling 55F, 75F, 95F 05 0%10% 18
Low refrigerant charge 55F, 75F, 95F 05 0%20% 18
High refrigerant charge 55F, 75F, 95F 05 0%4% 18
Low charge, evaporator fouling 75F 5 20% 1
Low charge, condenser fouling 75F 5 20% 1
Low charge, compressor valve leakage 75F 5 15% 1
Low charge, liquid-line restriction 75F 5 40% 1
Table 3. Single- and Multiple-Fault Test Matrix for the Walk-In Freezer
Fault
Ambient
Temperatures
Fault
Levels
Capacity
Degradation
Tests
No fault 55F, 75F, 95F 0 n/a 3
Compressor valve leakage 75F 05 0%10% 6
Liquid-line restriction 75F 05 0%60% 6
Condenser fouling 75F 05 0%2% 6
Evaporator fouling 75F 05 0%8% 6
Low refrigerant charge 75F 05 0%22% 6
Ice accumulation 75F n/a n/a 1
Evaporator fouling, compressor valve leak 75F 5 20% 1
Evaporator fouling, condenser fouling 75F 5 4% 1
Evaporator fouling, liquid-line restriction 75F 5 25% 1
Evaporator fouling, system undercharge 75F 5 5% 1
h
dis normal ,
h
suc
W

comp
Q

l oss

ref
----------------------------------- + =
82 HVAC&R RESEARCH
where h
suc
is suction enthalpy, is refrigerant mass flow rate, is compressor power,
and is compressor heat loss. The suction pressure is determined using a virtual pressure
sensor (described in the next section), and suction temperature is measured so h
suc
can be calcu-
lated. and are calculated using a compressor map and is estimated using a
simple model as described later. The normal discharge temperature is determined using property
relations with h
dis,normal
and the condensing pressure (P
c
) as inputs. Compressor maps are read-
ily available from compressor manufacturers.
The discharge temperature residual is calculated as the difference between the measured and
normal discharge temperatures as
. (2)
This residual is used as the fault feature for loss in compressor volumetric efficiency.
The refrigerant mass flow rate and compressor power used to determine this fault feature
would normally be obtained using a compressor map available from the manufacturer. The com-
pressor power consumption is insensitive to the compressor valve leakage fault. However, the
compressor map is not accurate for estimating refrigerant flow rate with this fault. To handle this
situation, Li (2004) proposed a refrigerant mass flow virtual sensor that is based on a compres-
sor energy balance and described as
. (3)
Condenser Fouling
Both the walk-in cooler and walk-in freezer use a constant-speed fan to force ambient air
across the condenser. When fouling develops, the air side of the condensing coils becomes dirty
and/or blocked by debris. This reduces the airflow over the coil, which increases the condensing
pressure of the refrigerant and wears down the compressor at a greater rate while decreasing the
cooling capacity and coefficient of performance (COP).
Li and Braun (2007b) used the air volumetric flow rate across the condenser coil as the fea-
ture to detect this fault. Airflow measurements are expensive and difficult to implement by a
technician in the field, so a virtual sensor was developed for estimating the airflow rate. Using
an energy balance on the condenser, the volumetric airflow rate can be estimated as
, (4)
where is the condenser air volumetric flow rate, v
ca
is the condenser air specific volume,
is the refrigerant mass flow rate, h
ll
is the liquid-line enthalpy, T
ll
is the liquid-line temperature,
and c
p,ca
is the specific heat of the condenser air.
Since the change in temperature of the air across the condenser is relatively small, it is reason-
able to assume that v
ca
and c
p,ca
are constant across the coils and can be calculated using the
inlet conditions. Equation 4 is an approximation because it is assumed that there are no refriger-
ant pressure drops across the condenser and liquid line so that P
c
is equal to the pressure in the
liquid line (P
ll
).
m

ref
W

comp
Q

l oss
W

comp
m

ref
Q

l oss
T
dis residual ,
T
dis measured ,
T
dis normal ,
=
m

ref
W

comp
Q

loss

h
dis
h
suc

----------------------------------- =
V

ca
v
ca
m

ref
h
dis
P
c
T
dis
, ( ) h
l l
P
c
T
l l
, ( ) ( )
c
p ca ,
T
cao
T
cai
( )
-------------------------------------------------------------------------------------------
V

ca
m

ref
VOLUME 15, NUMBER 1, JANUARY 2009 83
Evaporator Fouling
The primary effect of evaporator fouling is a reduction in airflow rate. For many air condi-
tioners, this fault would generally be caused by fouling of an evaporator air filter. However, the
walk-in cooler contains no filter, so evaporator fouling would occur due to deposits on the evap-
orator coil and through ice buildup. It could also occur, but to a lesser degree, by food products
blocking the evaporator air inlets inside the cooler.
The reduction in airflow rate is used as the feature for this fault. A virtual sensor for evapora-
tor airflow rate was presented by Li and Braun (2007b) as
, (5)
where is the evaporator air volumetric flow rate, v
ea
is the evaporator air specific volume,
P
e
is the evaporation pressure, h
eai
is the evaporator air inlet enthalpy, is the evaporator
air inlet relative humidity, h
eao
is the evaporator air outlet enthalpy, and is the evaporator
air outlet relative humidity. The use of h
ll
in Equation 5 implies that the expansion process is
isenthalpic.
For a cooler or freezer application, the air is very dry and it is not necessary to measure
humidity. The air enthalpy difference can be evaluated based on measurements of temperature
only and ideal gas behavior with constant specific heat.
Liquid-Line Restriction
Most commercial vapor-compression equipment uses a filter dryer in the liquid line to absorb
water dissolved in the refrigerant and to filter out dirt. Over time, these devices can become sat-
urated with water and dirt, which creates a large pressure drop.
In a typical air-conditioning system, there is usually about 15F of subcooling at the con-
denser exit. This makes it difficult to use a temperature drop across the filter dryer (T
ll
) as a
fault indicator, because it would require a large pressure drop to get the refrigerant to change
phase and generate a large enough T
ll
for detection. However, for systems with a liquid-line
receiver, the refrigerant at the condenser exit is a saturated liquid. Therefore, it requires a very
small pressure drop in the liquid line to get the refrigerant to change phase. For detection of a
liquid-line restriction, using two temperature measurements across the filter dryer should be
adequate since a change in temperature can be directly correlated to pressure drop.
Refrigerant Charge
Undercharge and overcharge of a vapor-compression system usually occurs during service of
the system. Refrigerant leakage is another fault that fits into this category because in doing fault
detection, leakage is detected in the same manner as system undercharge. During fault diagno-
sis, they can be separated because the refrigerant leakage will become more severe over time
while the system undercharge will remain constant. In this analysis, refrigerant leakage and sys-
tem undercharge will be classified together as system undercharge.
Li (2004) suggested that a good measure of system charge would be the difference between
the suction line superheat and the liquid-line subcooling, defined as
, (6)
where superheat and subcooling are defined as
V

ea
v
ea
m

ref
h
suc
P
e
T
suc
, ( ) h
ll
P
c
T
ll
, ( ) ( )
h
eai
T
eai

eai
, ( ) h
eao
T
eao

eao
, ( )
--------------------------------------------------------------------------------------------
V

ea

eai

eao
T
sh sc
T
sh
T
sc
=
84 HVAC&R RESEARCH
and (7)
. (8)
Values of T
sh
and T
sc
for normal refrigerant charge levels will vary to some extent based on
the ambient temperatures. To include the effects of driving conditions, Li and Braun (2007b)
developed a new charge feature described as
, (9)
where k
sc,sh
is the ratio of subcooling change to superheat change with the ambient temperatures
and T
sh,normal
and T
sc,normal
are values of superheat and subcooling degree for the normal refrig-
erant charge at a rating condition. For air-conditioning equipment, the feature in Equation 9 is
sensitive to refrigerant charge but is relatively insensitive to ambient conditions and other faults.
The rating condition for Equation 9 is somewhat arbitrary and could be chosen as the standard
rating condition for the air-conditioning unit.
Liquid-line receivers cause the refrigerant in the liquid line to be saturated, so the difference
between the normal and measured subcooling is approximately zero, even with changing operat-
ing conditions at full charge. Furthermore, the use of a TXV causes the superheat to be relatively
constant. Therefore, as charge is removed or added, the feature defined by Equation 9 should not
change until the receiver becomes either completely empty or completely full. When the system
is undercharged to the point where the receiver is empty, there should be two-phase refrigerant
in the liquid line. This will cause the TXV to fully open and lose control of the superheat.
Undercharging the system further should cause superheat to increase while subcooling should
remain approximately equal to zero. At the other extreme, when the receiver is completely full,
refrigerant will back up in the condenser and additional charge will cause an increase in the con-
denser subcooling. However, the superheat should remain relatively constant for overcharging.
Since superheat only changes for low-refrigerant charge and subcooling only varies for severe
overcharging, the simple feature of Equation 6 was chosen for characterizing these faults. How-
ever, practically it is not necessary to consider refrigerant overcharging for the walk-in cooler
and freezer because both systems had to be charged to at least 200% of nominal charge before
the receiver completely filled. Below this level of refrigerant charge there was essentially no
impact on system performance.
EVALUATION OF VIRTUAL SENSORS
Pressure
A virtual pressure sensor uses a thermocouple mounted to a return bend of the condenser or
evaporator to infer the high-side or low-side pressure. In both the walk-in cooler and walk-in
freezer experiments, eight thermocouples were soldered to the return bends of both the con-
denser and the evaporator and insulated from the ambient air using two layers of foam tape. The
thermocouples at the inlet to the evaporator and on the fifth return bend of the condenser were
selected to calculate evaporating and condensing pressures because these positions had
two-phase refrigerant for all conditions. The calculated pressures were then compared to the
averages of the measured inlet and outlet pressures for each heat exchanger.
Figures 4 and 5 compare measured and virtual evaporating and condensing pressures for the
walk-in cooler for all of the test conditions. Uncertainty bars for the virtual pressures are
included along with lines that represent an error of 5%. The differences between virtual and
T
sh
T
suc
T
e
=
T
sc
T
c
T
cond out ,
=
T
sh sc
k
sc sh ,
T
sh
T
sh normal ,
( ) T
sc
T
sc normal ,
( ) =
VOLUME 15, NUMBER 1, JANUARY 2009 85
actual pressure outputs were within the uncertainty of the actual pressure sensors except for one
evaporating pressure that had a difference of about 10%. This operating point corresponded to a
liquid-line restriction for the highest fault level at an ambient temperature of 75F. Very similar
results were obtained for the freezer. It should be noted that there is small bias in the results of
Figure 5 in that the virtual condensing pressure is always lower than the actual pressure mea-
surement. The most logical explanation for the bias is that the virtual pressure measurement is
based on the use of a surface temperature measurement, which is always lower than the refriger-
ant temperature.
Refrigerant Mass Flow
Figure 6 compares mass flow rate from the compressor map to measured mass flow rate for
all of the experiments performed on the walk-in cooler. The mass flow was calculated using
virtual pressure sensors as inputs and a correction for superheat. The large uncertainties in the
estimated mass flow rates are primarily due to uncertainties in the evaporating pressure esti-
mates. Overall the predictions agree well with measured mass flow rates. The root mean
square (RMS) errors were 1.06 lb
m
/h for all of the data. Somewhat worse results were
obtained for the freezer.
Figure 4. Comparison of virtual and measured evaporation pressures for the walk-in
cooler.
Figure 5. Comparison of virtual and measured condensing pressures for the walk-in
cooler.
86 HVAC&R RESEARCH
A compressor energy balance is used to determine refrigerant mass flow rate instead of the com-
pressor map when the compressor is known to be faulted. Figure 7 compares measured mass flow
rates with flow rates determined from a compressor energy balance for all of the data points. The
terms in the energy balance were evaluated using measured suction and discharge temperatures
with virtual measurements for pressures and power consumption. The accuracy of this approach is
considerably worse than that of the compressor map (RMS error of 3.63 lb
m
/h versus 1.06 lb
m
/h
for the compressor map approach). However, it is adequate for estimating mass flow rate for con-
ditions where the performance of the compressor is known to have degraded.
Compressor Power Based on a Compressor Map
Compressor power is estimated using a compressor map available from the manufacturer. The
map power was calculated using virtual pressure inputs to simulate the implementation of the
diagnostic method. The compressor map power was evaluated without correcting for superheat,
as recommended by Shen (2006). Figure 8 compares the compressor map power to the measured
compressor power for all of the experiments performed on the cooler. The uncertainty in the
compressor map is primarily due to uncertainties in the virtual pressure measurements. The
Figure 6. Comparison of refrigerant flow from a compressor map and measurements
for the walk-in cooler.
Figure 7. Comparison of refrigerant mass flow from a compressor energy balance and
measurements for the walk-in cooler.
VOLUME 15, NUMBER 1, JANUARY 2009 87
RMS error for the compressor map is 22.2 W. Comparisons for the compressor used in the
freezer were somewhat worse. However, the accuracy of these maps is adequate for use in esti-
mating discharge temperatures to perform compressor fault diagnoses.
Compressor Heat Loss
It is necessary to estimate compressor heat loss for use in an energy balance for determining the
normal compressor discharge temperature. For small hermetic compressors, heat loss typically is
between 20% and 75% of the input power consumption. Data collected from the experiments was
used to develop a simple correlation for heat loss. The heat loss is a function of the ambient tem-
perature and the compressor shell temperature. However, the instrumentation setup did not include
any shell temperature measurements, so a pseudo-shell temperature was defined as
. (10)
A heat loss factor was estimated from measurements and correlated with the temperature dif-
ference between the shell and ambient temperatures. The heat loss factor ( ) is the ratio of
the compressor heat loss to the electrical input and was estimated from measurements using an
energy balance on the compressor according to
. (11)
The measurements included everything necessary to evaluate the heat loss factor, including dis-
charge temperature and pressure, suction temperature and pressure, refrigerant mass flow rate,
and compressor power. However, these measurements would generally not be available; there-
fore, virtual measurements of pressures and power consumption and mass flow rate from com-
pressor maps were used to determine the heat loss factor.
The heat loss factor for the walk-in cooler is plotted as a function of the difference between the
shell temperature (T
shell
) and the ambient temperature (T
amb
) in Figure 9. The heat loss rate ranged
from about 46% to 56% of the compressor input power. The data points in Figure 9 include rela-
tively large error bars associated with the heat loss estimates. A best-fit linear relationship was
determined from the data and used for estimating heat loss as a function of temperature difference.
Figure 8. Compressor map power compared to the measured power for the walk-in
cooler.
T
shell
T
di s
T
suc
+
2
--------------------------

loss

loss
1
m

ref
h
dis
h
suc
( )
W

comp
------------------------------------------ =
88 HVAC&R RESEARCH
A linear relationship was chosen for simplicity and because one would expect a linear relationship
with temperature difference. For the freezer, the heat loss factors were much lower (between about
10% and 20%) because of the significantly lower suction pressures and temperatures. Furthermore,
there was a little more scatter in the data. However, a linear relationship also provided a reasonable
fit to the data.
FAULT FEATURE CHARACTERISTICS
In this section, individual fault features are presented as a function of percent capacity degra-
dation associated with different individual faults. The cooling capacity of the 0th fault level was
taken as a baseline value for determining percent capacity degradation. Cooling capacity degra-
dation was chosen as a general indicator of the severity of fault because it is more sensitive to
fault level than efficiency degradation. Furthermore, for refrigeration equipment, lost cooling
capacity could mean the loss of valuable refrigerated or frozen product, which is more important
than lost efficiency.
The cooling capacity was determined using a refrigerant flow measurement, and enthalpies
were determined with inlet and outlet refrigerant temperatures and pressures. Average uncer-
tainties in capacities for the refrigerator and the freezer were 7.1% and 7.4%, respectively. As
capacity is reduced with increasing fault severity, the absolute value of the uncertainty remains
relatively constant, causing the percent uncertainty to increase. Although these uncertainties
may seem high, cooling capacity measurements were not used to detect faults. They were only
used to provide a common way to indicate the level of each fault.
In this section, only sample results that highlight particular performance characteristics and
problems are presented. The fault features were calculated for two cases: 1) using all of the mea-
sured data and 2) using virtual sensors wherever possible. Table 4 summarizes the measure-
ments used in calculating the fault features when using all real sensors and when employing the
virtual sensors presented in this paper. The left-hand column is a list of all sensors required if all
real sensors (and no virtual sensors) were used in performing the diagnostics. The right-hand
column is a list of temperature sensors required if virtual sensors were used.
Compressor Valve Leakage
Figures 10 and 11 show the effect of capacity degradation due to compressor valve leakage on
discharge temperature residuals for the walk-in cooler and freezer. This fault feature increases
Figure 9. Cooler compressor heat loss factor as a function of the difference between the
pseudo-shell temperature and the ambient temperature (using virtual sensor data).
VOLUME 15, NUMBER 1, JANUARY 2009 89
with fault level for both systems with either real or virtual sensors. However, the discharge
residual is much less sensitive to capacity degradation for the cooler than for the freezer. The
small dependence of discharge temperature residual on capacity degradation for compressor
valve leakage fault with the cooler is problematic because this feature also depends on con-
denser fouling, as shown in the next section. This effect may be due to an inadequate heat loss
model for the compressor. The compressor heat loss is much more significant for the cooler
application than for the freezer application because of higher suction pressures and tempera-
tures. Therefore, errors in this model have a more significant effect on the discharge temperature
for the cooler.
The compressor valve leakage fault also caused the feature for low refrigerant charge to
increase at high fault levels for the walk-in cooler, as demonstrated in Figure 12. As the com-
pressor flow is reduced, the TXV must open to maintain the target superheat. At severe fault lev-
els, the TXV saturates open and the superheat increases. This could potentially lead to a false
alarm for low charge for capacity degradations greater than about 40%. However, this may not
be a problem, because the compressor fault would normally be identified at lower fault levels
before the charge feature is affected. Furthermore, false alarms could be avoided for individual
faults because the diagnostic feature for compressor valve leakage would take precedence. In
other words, if the compressor temperature discharge feature exceeded a threshold for this fault,
Table 4. Measurements for Estimating Diagnostic Features with Real and Virtual Sensors
(See Figure 3 for Nomenclature)
Measurements Required
When Using All Real Sensors
Measurements Required
When Using Virtual Sensors
T
dis
, F T
dis
, F
T
cond,out
, F T
cond,out
, F
T
ll1
, F T
ll1
, F
T
ll2
, F T
ll2
, F
T
exp,in
, F T
exp,in
, F
T
evap,in
, F T
evap,in
, F
T
evap,out
, F T
evap,out
, F
T
suc
, F T
suc
, F
T
eao
, F T
eao
, F
T
eai
, F T
eai
, F
T
cao
, F T
cao
, F
T
cai
, F T
cai
, F
T
cond,5
, F T
cond,5
, F
P
dis
, psia
P
cond,out
, psia
P
evap,in
, psia
P
suc
, psia
, lb
m
/h
, W
m

ref
W

comp
90 HVAC&R RESEARCH
Figure 10. T
dis,residual
as a function of the percent cooling capacity degradation due to
compressor valve leakage for the walk-in cooler.
Figure 11. T
dis,residual
as a function of the percent cooling capacity degradation due to
compressor valve leakage for the walk-in freezer.
Figure 12. T
shsc
as a function of the percent cooling capacity degradation due to com-
pressor valve leakage for the walk-in cooler.
VOLUME 15, NUMBER 1, JANUARY 2009 91
then the refrigerant charge feature would be ignored. However, under some circumstances it
may not be possible to remotely diagnose both low refrigerant charge and compressor valve
leakage if they are simultaneously present. In this case, a technician might need to visit the unit
and use a sight glass to determine whether the refrigerant charge is low prior to deciding
whether to replace the filter dryer.
The virtual condenser airflow rate determined with virtual sensors was also found to have
some dependence on the compressor valve leakage fault. However, this feature increased with
fault level, which would not lead to a false alarm.
Condenser Fouling
Figure 13 shows the impact of the condenser fouling fault level on the virtual condenser
aiflow rate. This feature is very sensitive to the condenser fault and decreases by a factor of two
for about a 5% degradation in cooling capacity. This feature was also effective in characterizing
condenser fouling for the freezer.
For the cooler, the discharge temperature residual also increased with condenser fouling level
(shown in Figure 14). This will lead to a false alarm for a compressor valve leakage fault in the
presence of condenser fouling. The discharge temperature residual did not have the same level
Figure 13. Virtual condenser airflow rate as a function of percent capacity degradation
due to condenser fouling for the walk-in cooler.
Figure 14. Discharge temperature residual as a function of percent capacity degrada-
tion due to condenser fouling for the walk-in cooler.
92 HVAC&R RESEARCH
of dependence on condenser fouling for the freezer. Therefore, this behavior may be due to an
inadequate heat loss model for the compressor. As previously mentioned, the compressor heat
loss is much more significant for the cooler application than for the freezer application because
of higher suction pressures and temperatures.
Evaporator Fouling
Figure 15 shows the impact of evaporator fouling on the virtual evaporator airflow rate. The
results are very similar to those for condenser fouling. This feature is very sensitive to fouling
and decreases by a factor of two for less than a 5% degradation in capacity. Very similar results
were obtained for the freezer. Furthermore, this feature was insensitive to other faults and would
not trigger any false alarms.
One of the issues for the walk-in freezer was the ability to distinguish between evaporator
fouling and ice accumulation on the heat exchanger prior to a defrost cycle. To characterize the
ice buildup, an ice accumulation test was run. First, the freezer was allowed to reach its normal
steady-state operating conditions. Then, a pot of boiling water was placed inside the freezer on a
heater to promote the accumulation of ice. When a steady state was achieved again, data was
collected for about 9600 seconds. Over time, the temperature difference across the evaporator
(T
ea
) increased as ice built up on the heat exchanger. As a result, the virtual evaporator airflow
rate decreased over time, as shown in Figure 16, and is a good indicator of the buildup of frost.
Although not shown in this figure, the virtual evaporator airflow feature returned to a value of
about 1200 cfm after a defrost cycle.
Evaporator fouling and ice accumulation could easily be distinguished within a fault detection
system. After a defrost cycle and when steady-state conditions are achieved, the evaporator
aiflow rate should return to a normal, unfaulted value. If the evaporator were fouled, then a
defrost cycle would not return the virtual evaporator airflow indicator to normal and a fault
could be flagged.
In fact, the virtual evaporator airflow feature could be used as part of a smart defrosting
scheme to determine the need for defrosting. A threshold could be determined that balances the
trade-offs between improved cycle efficiency and defrost energy. This could lead to reduced
defrost cycles for situations where the evaporator is not exposed to significant moisture (e.g.,
infrequent door openings) and increased defrost cycles for cases with high indoor box moisture
levels.
Figure 15. Virtual evaporator airflow rate as a function of capacity degradation due to
evaporator fouling for the walk-in cooler.
VOLUME 15, NUMBER 1, JANUARY 2009 93
Liquid-Line Restriction
Figure 17 shows the effect of capacity degradation due to liquid-line restriction on the liquid-
line temperature difference for the cooler. This fault feature increases significantly with fault
level for all three ambient conditions. However, the impact is less at lower ambient temperatures
due to a lower pressure ratio for the system. For the same fault level, the pressure drop across the
restriction is larger for higher ambient temperatures, which increases the temperature drop
across the restriction. Very similar results were obtained for the freezer.
The liquid-line restriction also caused the feature for low refrigerant charge to increase for
the walk-in freezer, as demonstrated in Figure 18. Again, this occurs because of an interaction
with the TXV. As the refrigerant flow is reduced with a liquid-line restriction, the TXV must
open to maintain the target superheat. At a sufficient fault level, the TXV saturates open and the
superheat increases. Although the refrigerant charge feature is sensitive to liquid-line restric-
tion, false alarms can be avoided for individual faults because the diagnostic feature for liq-
uid-line restriction would take precedence. In other words, if the liquid-line feature exceeded a
threshold for this fault, then the refrigerant charge feature would be ignored. However, it may
Figure 16. Virtual evaporator over the course of the ice accumulation test.
Figure 17. Liquid-line temperature difference as a function of percent capacity degra-
dation due to a liquid-line restriction for the walk-in cooler.
94 HVAC&R RESEARCH
not be possible to remotely diagnose both low refrigerant charge and a liquid-line restriction if
they are simultaneously present. In this case, a technician would need to visit the unit and use a
sight glass to determine whether the refrigerant charge is low.
A liquid-line restriction did not cause changes in other fault features that would lead to false
alarms.
Refrigerant Charge
Figure 19 shows the effect of refrigerant charge level on capacity degradation for the walk-in
cooler. Overcharging the system had very little effect on performance because of the liquid-line
receiver. Very high charge levels had a small impact at low ambient temperatures at the point
where the receiver became full of refrigerant. Reduced charge did have an impact on cooling
capacity once the charge was less than about 80% of the nominal charge. At this point, the liq-
uid- line receiver was empty and the TXV saturated at a fully open position. This led to an
increasing superheat and fault feature with decreasing charge, as demonstrated in Figure 20.
Similar results were obtained for the freezer.
Figure 18. T
shsc
as a function of percent capacity degradation due to a liquid-line
restriction for the walk-in freezer.
Figure 19. Percent capacity degradation as a function of charge for the walk-in cooler.
VOLUME 15, NUMBER 1, JANUARY 2009 95
Low or high refrigerant charge did not have a significant effect on other fault features and
would not lead to false alarms. Although it is not possible to detect system overcharging for
these systems, this is not a problem because overcharging does not impact system performance.
DIAGNOSTIC PERFORMANCE
Li (2004) defined a normalized fault indicator as
, (12)
where IND
faultname
is the fault indicator for a specific fault, fv
current
is the current fault feature
value, and fv
predefined
is the predefined fault feature value.
For this study, the fv
predefined
was chosen as the feature value that represented a 10% cooling
capacity loss. All of the thresholds for IND
faultname
were then set to be 0.5, which represents a
5% capacity loss. Tables 5 and 6 show the fv
predefined
and fault thresholds set for the walk-in
cooler and freezer, respectively. The compressor valve leakage fault thresholds are extremely
different between the cooler and freezer because of different impacts on cooling capacity.
Fault features determined using data from the experiments were used along with the fault
thresholds to determine the levels at which faults could be detected and to identify false alarms.
Virtual sensors were used for this evaluation to represent actual implementation of the decou-
pled diagnostic method. Tables 7 and 8 give results for fault detection sensitivity. For the single-
fault tests on both the walk-in cooler and walk-in freezer there were no missed faults. The diag-
nostic method was able to detect compressor valve leak, condenser fouling, and liquid-line
restriction at the first fault level implemented on the cooler. Liquid-line restriction was detected
at the first level on the freezer as well.
Although the diagnostic method did not miss any faults, in some cases two faults were identi-
fied when only a single fault was present. Table 9 shows the implemented faults and fault levels
that resulted in secondary fault indications. As previously described, both compressor valve
leakage and liquid-line restriction can influence the feature used to determine low refrigerant
charge if they are severe enough to cause the TXV to saturate in a fully open position. For the
compressor valve leakage fault with the cooler, this occurred at a significantly higher fault level
Figure 20. Difference in T
sh
and T
sc
as a function of percent nominal charge for the
walk-in cooler.
IND
f aul tname
fv
current
fv
predef ined
---------------------------- =
96 HVAC&R RESEARCH
Table 5. Fault Thresholds and Predefined Fault Feature Values for the Walk-In Cooler
Fault
Compressor
Valve Leak
Condenser
Fouling
Evaporator
Fouling
Liquid-Line
Restriction
System
Undercharge
Feature T
dis
T
ll
T
sh-sc
fv
predefined
1.5F 652 ft
3
/min 594 ft
3
/min 4.7F 4.7F
Threshold 0.5 0.5 0.5 0.5 0.5
Table 6. Fault Thresholds and Predefined Fault Feature Values for the Walk-In Freezer
Fault
Compressor
Valve Leak
Condenser
Fouling
Evaporator
Fouling
Liquid-Line
Restriction
System
Undercharge
Feature T
dis
T
ll
T
sh-sc
fv
predefined
41.7F 600 ft
3
/min 700 ft
3
/min 2.2F 4.2F
Threshold 0.5 0.5 0.5 0.5 0.5
Table 7. Fault Levels Where the Algorithm Could First Diagnose Individual Faults
for the Walk-In Cooler
Fault
Ambient
Temperature
Fault
Level
Degraded
Capacity
Compressor valve leak 55F First 21.0%
Compressor valve leak 75F First 16.0%
Compressor valve leak 95F First 17.5%
Condenser fouling 55F First 1.2%
Condenser fouling 75F First 1.2%
Condenser fouling 95F First 2.6%
Evaporator fouling 55F Second 1.9%
Evaporator fouling 75F Third 5.2%
Evaporator fouling 95F Third 6.1%
Liquid-line restriction 55F First 19.4%
Liquid-line restriction 75F First 7.5%
Liquid-line restriction 95F First 8.5%
System undercharge 55F Fourth 7.9%
System undercharge 75F Third 10.5%
System undercharge 95F Third 6.1%
V

ca
V

ea

ca
V

ea

VOLUME 15, NUMBER 1, JANUARY 2009 97


(third level in Table 9) than necessary to detect the compressor valve leakage fault (first level in
Table 7) and would not be a problem. However, the fault indicators for both liquid-line restric-
tion and low refrigerant charge fault exceeded their thresholds at the lowest level of liquid-line
restriction. If it were only necessary to diagnose a single fault, then this would not be a problem
because the indicator for liquid-line restriction would take precedence since it is independent of
other faults. However, the method cant reliably diagnosis combinations of liquid-line restric-
tion/low refrigerant charge or compressor valve leakage/low refrigerant charge. In order to elim-
inate low refrigerant charge as a possible fault in situations where these fault combinations are
indicated, it would be necessary to visit the site and observe the refrigerant level using a receiver
sight glass.
Table 9 also shows a secondary fault indication for compressor valve leakage when condenser
fouling was implemented. Both faults were indicated at the lowest level of condenser fouling, so
the condenser fouling indicator would take precedence because its feature is independent of all
faults. As previously discussed, it is believed that diagnosis of a compressor valve leakage for
this case is due to inaccuracies in the estimate of the normal compressor discharge temperature
due to an inadequate heat loss model.
Besides the secondary fault indications, there were no false alarms encountered for either the
cooler or freezer at any of the fault levels.
Diagnostic performance was also examined for the multiple fault tests described in Tables 2
and 3. For these tests, the method applied to the cooler yielded no false alarms. Also, the method
could correctly diagnosis low refrigerant charge combined with either condenser or evaporator
fouling. However, as expected, the method could not detect combinations of refrigerant under-
charge with either a liquid-line restriction or compressor valve leakage. For the freezer, the diag-
nostic method correctly detected evaporator fouling combined with either compressor valve
leakage or refrigerant undercharge. However, the method detected refrigerant undercharge in addi-
tion to evaporator fouling and liquid-line restriction when those two faults were implemented.
Table 8. Fault Levels Where the Algorithm Could First Diagnose the Individual Faults
for the Walk-In Freezer
Fault
Ambient
Temperature
Fault
Level
Degraded
Capacity
Compressor valve leak 75F Third 3.5%
Evaporator fouling 75F Third 3.0%
Liquid-line restriction 75F First 22.5%
System undercharge 75F Fourth 11.9%
Table 9. Implemented Faults Causing Secondary Fault Indications
System
Implemented
Fault
First Fault Level
Indicating
Secondary Fault
Falsely Indicated
Fault
Cooler Compressor valve leak Third System undercharge
Cooler Condenser fouling First Compressor valve leak
Cooler Liquid-line restriction First System undercharge
Freezer Liquid-line restriction First System undercharge
98 HVAC&R RESEARCH
CONCLUSIONS
In general, the decoupling features developed for compressor valve leakage, condenser foul-
ing, evaporator fouling, liquid-line restriction faults, and refrigerant charge should work well for
diagnostics applied to vapor-compression systems used in walk-in coolers and freezers. However,
it was found that the refrigerant charge feature is not completely decoupled from liquid-line
restriction and compressor valve leakage faults at high-fault levels. Systems that utilize a liq-
uid-line receiver and TXV are insensitive to charge until the system becomes starved and the
TXV saturates open. A saturated TXV leads to an increase in superheat, which is used to indicate
low refrigerant charge. However, the TXV can also be saturated open when refrigerant flow is
reduced significantly due to a high level of compressor valve leakage or a severe liquid-line
restriction.
Although the refrigerant charge feature is sensitive to liquid-line restriction and compressor
valve leakage faults, false alarms can be avoided for individual faults because the diagnostic
features for these other faults work well and would identify the appropriate fault in the absence
of low charge. However, it is not possible to identify multiple fault combinations of low refrig-
erant charge and either compressor valve leakage or liquid-line restriction. Furthermore, it is
not possible to determine overcharging of refrigerant with this feature, unless the charge is
severely overcharged to the point where the receiver is full. For systems with liquid-line
receivers and TXVs, it may make sense to employ a low-cost level sensor in the receiver. The
normal charge level could be correlated with operating conditions and used to decouple charge
from other faults.
There is also a need to develop a better model for estimating heat loss in order to improve
estimates of the normal compressor discharge temperature under all conditions. This should
eliminate an indication of compressor valve leakage when condenser fouling is present.
Evaporating frosting tests were also implemented, and the evaporator fouling feature was
found to be a reliable indicator. This feature could be used as part of a smart defrosting scheme
to determine the need for defrosting. A threshold could be determined that balances trade-offs
between improved cycle efficiency and defrost energy. This could lead to reduced defrost cycles
for situations where the evaporator is not exposed to significant moisture (e.g., infrequent door
openings) and increased defrost cycles for cases with high indoor box moisture levels.
ACKNOWLEDGEMENTS
This work was supported by the U.S. Department of Energy. We appreciate the donation of
equipment provided by Manitowoc, Inc., and coordinated by Daryl Erbs.
REFERENCES
Breuker, M.S., and J.E. Braun. 1998. Evaluating the performance of a fault detection and diagnostic system
for vapor compression equipment. HVAC&R Research 4(4):40123.
Cowan, A. 2004. Review of recent commercial rooftop unit field studies in the Pacific Northwest and Cali-
fornia. Northwest Power and Conservation Council and Regional Technical Forum, October 8, Port-
land, OR.
Grimmelius, H.T., J.K. Woud, and G. Been. 1995. Online failure diagnosis for compression refrigeration
plants. International Journal of Refrigeration 18(1):3141.
Li, H., and J.E. Braun. 2003. An improved method for fault detection and diagnosis applied to packaged air
conditioners. ASHRAE Transactions 109(2):68392.
Li, H. 2004. A decoupling-based unified fault detection and diagnosis approach for packaged air condition-
ers. PhD thesis, School of Mechanical Engineering, Purdue University, West Lafayette, IN.
Li, H., and J.E. Braun. 2006. Evaluation of a decoupling-based fault detection and diagnostic technique
Part II: Field evaluation and application. Journal of Harbin Institute of Technology 13:16471.
VOLUME 15, NUMBER 1, JANUARY 2009 99
Li, H., and J.E. Braun. 2007a. A methodology for diagnosing multiple-simultaneous faults in vapor com-
pression air conditioners. HVAC&R Research 13(2):36995.
Li, H., and J.E. Braun. 2007b. Decoupling features and virtual sensors for diagnosis of faults in vapor com-
pression air conditioners. International Journal of Refrigeration 30(3):54664.
Pak, B.C., E.A. Groll, and J.E. Braun. 2005. Impact of fouling and cleaning on plate fin and spine fin heat
exchanger performance. ASHRAE Transactions 111(1):496504.
Proctor, J., and T. Downey. 1995. Heat pump and air conditioner performance. Affordable Comfort Con-
ference, March 2631, Pittsburgh, PA.
Rossi, T.M., and J.E. Braun. 1997. A statistical, rule-based fault detection and diagnostic method for vapor
compression air conditioners. HVAC&R Research 3(1):1937.
Shen, B. 2006. Improvement and validation of unitary air conditioner and heat pump simulation models at
off-design conditions. PhD thesis, School of Mechanical Engineering, Purdue University, West Lafay-
ette, IN.
Stylianou, M., and Y.S. Lau. 1996. Performance monitoring, fault detection, and diagnosis of reciprocating
chillers. ASHRAE Transactions 102(1):61527.
Yang, L., J.E. Braun, and E.A. Groll. 2007. The impact of fouling on the performance of filter-evaporator
combinations. International Journal of Refrigeration 30(3):48998.

Вам также может понравиться