You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/282635959

A simulation approach for hermetic reciprocating compressors including


electrical motor modeling

Article in International Journal of Refrigeration · July 2015


DOI: 10.1016/j.ijrefrig.2015.07.023

CITATIONS READS

30 1,754

2 authors, including:

Thiago Dutra
Federal University of Santa Catarina
16 PUBLICATIONS 301 CITATIONS

SEE PROFILE

All content following this page was uploaded by Thiago Dutra on 14 May 2019.

The user has requested enhancement of the downloaded file.


international journal of refrigeration 59 (2015) 168–181

Available online at www.sciencedirect.com

ScienceDirect

j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / i j r e f r i g

A simulation approach for hermetic


reciprocating compressors including electrical
motor modeling

Thiago Dutra, Cesar J. Deschamps *


POLO Research Laboratories for Emerging Technologies in Cooling and Thermophysics, Federal University of
Santa Catarina, CEP 88048-300 Florianopolis, SC, Brazil

A R T I C L E I N F O A B S T R A C T

Article history: The design of a high-efficiency reciprocating compressor requires an understanding of the
Received 16 January 2015 interactions between different phenomena occurring inside the compressor. This paper de-
Received in revised form 9 June scribes a comprehensive simulation approach for hermetic reciprocating compressors including
2015 modeling of the electrical motor. The simulation of the compression cycle follows an inte-
Accepted 15 July 2015 gral control volume formulation for mass and energy conservation. A thermal model is adopted
Available online 23 July 2015 with steady-state thermal energy balances applied to the compressor components via global
thermal conductances. The equivalent circuit method is employed to form a steady-state
Keywords: model of a single-phase induction motor. The coupling between the three models pro-
Reciprocating compressor vides the motor slip and mean compressor speed, which are seen to affect the compressor
Electrical model efficiency. The simulation model is validated through comparisons between predictions and
Induction motor measurements of the parameters associated with the compressor efficiency, temperature
Superheating distribution and motor performance. A parametric analysis is carried out to investigate the
dependence of the motor temperature on the input voltage and the results are discussed.
© 2015 Elsevier Ltd and International Institute of Refrigeration. All rights reserved.

Approche par simulation pour des compresseurs à pistons


hermétiques incorporant la modélisation du moteur
électrique
Mots clés : Compresseur à pistons ; Modèle électrique ; Moteur à induction ; Surchauffe

need to be considered. The development of more accurate simu-


1. Introduction lation models is of paramount importance in the prediction
of the compressor performance. Different phenomena take
In the design of refrigeration compressors a number of key place inside compressors, encompassing thermodynamic,
factors, such as efficiency, reliability and manufacturing cost, heat transfer and electromechanical processes, requiring

* Corresponding author. POLO Research Laboratories for Emerging Technologies in Cooling and Thermophysics, Federal University of Santa
Catarina, CEP 88048-300 Florianopolis, SC, Brazil. Tel.: +55 48 3234 5691; Fax: +55 48 3234 5166.
E-mail address: deschamps@polo.ufsc.br (C.J. Deschamps).
http://dx.doi.org/10.1016/j.ijrefrig.2015.07.023
0140-7007/© 2015 Elsevier Ltd and International Institute of Refrigeration. All rights reserved.
international journal of refrigeration 59 (2015) 168–181 169

Nomenclature

A area (m2) τ torque (N m)


cv specific heat at constant volume (J kg−1 K−1) ϕ gas mixing factor (–)
E voltage (V) ω compressor speed (rad s−1)
f (s) function of s
h specific enthalpy (J kg−1) Subscripts
H heat transfer coefficient (W m−2 K−1) 0 reference
I electrical current (A) b bearings
j complex number indicator (–) c condensing
m mass (kg) core stator core
m actual mass flow rate (kg s−1) dc discharge chamber
m bd backflow through the discharge port (kg s−1) dl discharge line
m bs backflow through the suction port (kg s−1) dm discharge muffler
md mass flow rate through the discharge dt discharge tube
port (kg s−1) e evaporating
l
m leakage flow rate through piston–cylinder ee external environment
gap (kg s−1) g gas in the compression chamber
ms mass flow rate through the suction port (kg s−1) h compressor housing
m th ideal mass flow rate (kg s−1) i lumped element index
p pressure (Pa) ie internal environment
P motor number of poles (–) in input
Q convective heat transfer rate (W) m motor; mean; magnetizing
R electrical resistance (Ω) n nominal
s slip ratio (–) out output
t time (s) r required
T temperature (K) rot rotor
UA global thermal conductance (W K−1) s isentropic
V volume (m3) sc suction chamber
W rate of work (W) sh shaft
X electrical reactance (Ω) sl suction line
Z electrical impedance (Ω) sm suction muffler
sta stator
Greek letters u upper
β temperature coefficient (K−1) w cylinder wall
Δ difference (–)
ηv volumetric efficiency (–) Superscripts
ηs isentropic efficiency (–) + due to forward rotating magnetic field
ηele electrical motor efficiency (–) − due to backward rotating magnetic field
ρ specific mass (kg m−3) ′ referred to stator side

multi-physics modeling for comprehensive simulations. In this finite volume method (FVM) has also been employed in com-
regard, different approaches have been developed to simu- pressor thermal simulations. Kara and Oguz (2010) used a
late the compression cycle and to predict the compressor commercial FVM package to solve heat conduction in solid com-
temperature distribution. The simulation of the compression ponents, prescribing temperature measurements and typical
cycle has been carried out through approaches based on poly- heat transfer coefficients as boundary conditions. Raja et al.
tropic processes (Cavallini et al., 1996; Chikurde et al., 2002), (2003) and Birari et al. (2006) adopted FVM to solve governing
lumped-parameter models (Rigola et al., 2000; Todescat et al., equations in both solid and fluid compressor domains, that is,
1992) and distributed-parameter models (Birari et al., 2006). without requiring the prescription of thermal conductances.
The temperature distribution inside the compressor is Recently, hybrid models have been developed by combin-
usually predicted via lumped-parameter methods, which require ing a lumped formulation for the fluid domain and FVM for
the establishment of thermal conductances between ele- heat conduction in solid elements (Ribas, 2007; Sanvezzo and
ments. Some authors (Meyer and Thompson, 1990; Todescat Deschamps, 2012). Hybrid models are less computationally de-
et al., 1992) have determined such conductances using experi- manding than those based solely on FVM and they can better
mental data, whereas others (Cavallini et al., 1996; Ooi, 2003; describe the heat transfer in solid components than lumped-
Rigola et al., 2000) have adopted classical heat transfer parameter models. All of the aforementioned models take into
correlations. In addition to lumped-parameter models, the account the electrical motor via relationships between torque,
170 international journal of refrigeration 59 (2015) 168–181

efficiency and compressor speed, usually identified from ex-


perimental data.
Electrical models are commonly employed for the analy-
sis of electrical machines, e.g., Totally-Enclosed Fan-Cooled
(TEFC) induction motors. Many papers in the literature (Huai
et al., 2003; Hwang et al., 2000; Mellor et al., 1991; Mezani et al.,
2005) propose the thermal–electrical modeling of electrical
motors, since stator and rotor electrical resistances are de-
pendent on temperature and vice-versa. Some of these
proposals combine lumped-parameter thermal models with
equivalent circuit methods (Eltom and Moharari, 1991; Mellor
et al., 1991) and others adopt the finite element method for
both thermal and electrical modeling (Hwang et al., 2000; Li
et al., 2010).
With regard to the compressor simulation, few research-
ers have included an electrical model for the driving motor.
Peruzzi et al. (1980) employed thermodynamic–thermal–
electrical modeling as a design tool to assess proposals to
improve the efficiency of a reciprocating compressor, but the
authors did not provide details regarding the electrical motor
model. He et al. (2013) developed a simulation model for a
semi-hermetic screw compressor employing the equivalent
circuit method for the thermal and electrical modeling. The
authors validated their model by comparing predictions with
measurements of the temperature along the stator winding
of the three-phase induction motor and carried out an analy-
sis of the temperature distribution as a function of the
geometrical parameters of the motor. Mantri et al. (2014) pro-
posed a multi-physics simulation model for a reciprocating Fig. 1 – Schematic of the compressor chamber.
compressor with a permanent magnet linear motor repre-
sented by the equivalent circuit. Their electrical model was
validated through a comparison between the measured and
predicted motor current. Unfortunately, the studies of He et al. 2. Mathematical model
(2013) and Mantri et al. (2014) provided little information re-
garding the input parameters and the procedure for the solution
The mathematical model is obtained through the combina-
of the motor circuit.
tion of three sub-models: (i) a thermodynamic model for the
This paper reports a simulation model developed to predict
compression cycle; (ii) a thermal model for the prediction of
the performance of hermetic reciprocating compressors. The
the compressor temperature distribution; and (iii) an electri-
sub-models adopted to solve the compression cycle and tem-
cal model for the 2-pole single-phase induction motor. All
perature distribution are based on lumped-parameter methods
models are solved in a coupled manner supplying input data
and are similar to those employed by Link and Deschamps
to each other. The following sections provide details regard-
(2011) and Todescat et al. (1992), respectively. However, instead
ing these models.
of the usual approach in which experimental or theoretical
data are used to express the relationship between the motor
efficiency and the compressor shaft power, a specific sub- 2.1. Thermodynamic model
model based on the motor equivalent circuit is proposed for
the single-phase induction motor. The embedding of a model The compression cycle is solved using the approach de-
for the electrical motor in a compressor simulation model scribed in Link and Deschamps (2011). Basically, this model
has the advantage of allowing the parametric study of the consists of an integral formulation of the unsteady-state mass
motor. The thermal–electrical coupling makes it possible to and energy conservation equations applied to the compres-
evaluate the electrical resistance of a motor as a function of sion chamber of a reciprocating compressor. Therefore,
its temperature. Moreover, the thermodynamic–electrical cou- thermodynamic properties such as pressure and tempera-
pling also allows estimates of the motor slip and mean speed ture are assumed to be time-dependent and spatially uniform
of the compressor to be obtained for different operating inside the chamber. Fig. 1 shows a schematic of the compression
conditions. chamber highlighting the energy fluxes through the control
Results for the compressor performance obtained with the volume boundaries. The mass and energy balances are repre-
new simulation model are expressed in terms of volumetric sented by Eqs. (1) and (2):
and isentropic efficiencies, as well as the temperature distri-
bution and motor parameters, validated via comparisons with dmg
s −m
=m d −m
l −m
 bs + m
 bd (1)
measured data. dt
international journal of refrigeration 59 (2015) 168–181 171

dTg 1 ⎪⎧  Q i − W
 i = (mh
∑  )i,in − ∑ (mh
 )i, out (6)
= ⎨Q w + m s (hsc − hg ) + m bd (hdc − hg )
dt mg c v ⎪⎩
⎡ Tg ∂pg ⎛ dVg dmg ⎞ ⎤ ⎪⎫ where Q i encompasses the rate of heat transfer between a
−⎢ ⎜ ρg − ⎟ ⎥⎬ (2)
⎢⎣ ρg ∂Tg ρ ⎝ dt dt ⎠ ⎥⎦ ⎪⎭ generic lumped element i and the surroundings, as well as the
rate of any heat generation within the element i. The rate of
Data on the mass flow rates through the suction and dis- work associated with the element i is W  i , and the terms due
charge valves ( m s, m
 d, m  bd ) as well as the leakage in the
 bs, m to advective energy transport into and from the element i are
piston–cylinder gap, m  l , are required. The former are modeled denoted by ∑ (mh
 )i,in and ∑ (mh
 )i, out , respectively.
with reference to the isentropic compressible flow condition The reciprocating compressor was divided into eight lumped
in a convergent nozzle combined with data on the effective elements: suction muffler (1), compression chamber (2),
flow area. The dynamics of the suction and discharge valves discharge chamber (3), discharge muffler (4), discharge tube (5),
is solved via one-degree-of-freedom models, considering stiff- electrical motor (6), internal environment (7) and compressor
ness and damping terms combined with the concept of effective housing (8). The internal environment represents the portion
force area. Details regarding the mass flow rates and valve dy- of refrigerant inside the housing. A schematic view of the
namics calculation can be found in Link and Deschamps (2011). lumped elements and the thermal interactions between them
Leakage is calculated according to the method proposed by Lilie is given in Fig. 2, with the different arrows identifying differ-
and Ferreira (1984). The specific enthalpies associated with mass ent mechanisms of energy transfer. The main gas path is located
flow rates, hsc and hdc, are evaluated from temperatures cal- between elements (1) and (5), but gas leakage occurs through
culated in the thermal model using the library REFPROP 7.0 the piston–cylinder clearance between the compression
(Lemmon et al., 2002). Equations (1) and (2) are solved via a chamber (2) and the internal environment (7), along with a small
fourth-order Runge–Kutta method. mass flow through the pressure equalizing orifices in the
The instantaneous heat transfer rate between the gas and suction muffler (1). Equation (6) is applied to all lumped
cylinder walls is Q w : elements identified in Fig. 2, giving rise to a system of non-
linear equations that are iteratively solved via the Newton–
Q w = Hw Aw (Tw − Tg ) (3) Raphson method.
Tables 1 and 2 show the corresponding terms Q , W ,
with Aw being the surface area of the compression chamber ∑ (mh
 )i,in and ∑ (mh
 )i, out , associated with each one of the
and Hw the convective heat transfer coefficient, calculated control volumes defined in the energy balances of the thermal
through the correlation proposed by Annand (1963). The tem- model, identified in Fig. 2. It should be mentioned that the cyl-
peratures Tg and Tw represent, respectively, the instantaneous inder wall temperature, Tw, is calculated from the steady-
gas temperature and the average temperature of the cylinder state energy balance applied to the compression chamber (2).
wall, the latter being obtained from the thermal model. The input data required to run the thermal model are the mass
The thermodynamic model also computes the indicated flow rates in different regions ( m ,m s, m  d, m
 bs, m  l ) , heat gen-
 bd, m
power, W  ind , through the integration of the instantaneous pres- eration rates in different parts of the electrical motor ( Q sta ,
sure with respect to the compression chamber volume: Q rot and Q core ), indicated power ( W  ind ), mechanical losses in
the bearing system ( W  b ), time-averaged specific enthalpy in
 ind = − ω the cylinder ( hg ) and refrigerant gas mixing factor ( ϕ ).
W
2π ∫ p dV
g g (4)
The mass flow rates, compression power and time-averaged
specific enthalpy in the cylinder are obtained from the ther-
where ω is the compressor speed obtained with the electrical
modynamic model. The mechanical losses are estimated from
model and Vg is the instantaneous volume of the compres-
measurements and the motor losses are computed from the
sion chamber, which is calculated as a function of the
electrical model. The refrigerant gas mixing factor accounts for
geometrical parameters of the compressor (An et al., 2002). The
 r , is obtained by the proportion of gas entering through the suction line, which
power required to run the compressor, W
 b , and is mixed in the internal environment of the compressor, as pro-
summing the total mechanical losses in the bearings, W
posed by Meyer and Thompson (1990). In some compressors,
the indicated power:
the suction line is directly connected to the suction muffler
r =W
 ind + W
b (5) and there is no mixing between the gas from the suction line
W
and the gas in the internal environment; i.e. φ = 1. In other com-
Estimates of W  b can be obtained from experimental data pressor designs, the suction line and the entrance of the suction
or theoretically. For the compressor considered in this study, muffler are misaligned and some of the gas entering the com-
W b is estimated to be approximately 9 W and almost insen- pressor flows into the internal environment of the compressor
sitive in relation to the operating conditions. before reaching the suction muffler inlet; hence 0 ≤ φ < 1. In this
study, a compressor with a direct suction system was adopted
2.2. Thermal model (φ = 1).
Heat transfer between the control volumes defined inside
A steady-state integral formulation of the energy equation was the compressor is taken into account by introducing global con-
applied to the compressor components (Todescat et al., 1992), ductances (UAs). These conductances are explicitly obtained
neglecting the kinetic and potential energies. A generic form from energy balances for each control volume in Fig. 2, based
of this equation can be mathematically represented by: on a set of temperature measurements for a single operating
172 international journal of refrigeration 59 (2015) 168–181

Table 1 – Heat transfer rate terms used in the energy Table 2 – Rate of work and advective energy transport
balances of the thermal model. terms used in the energy balances of the thermal
model.
Component Q

1. Suction muffler −UAsm (Tie − Tsm )
Component W 
mh ∑ 
mh
in ∑ out

2. Compression chamber UAw (Tw − Tie ) 1. Suction –  [ϕ hsl + (1 − ϕ ) hie ] + m


m  shsc
3. Discharge chamber UAdc (Tdc − Tie ) muffler +m l hie + m
 bshg
4. Discharge muffler UAdm (Tm,dm − Tie ) 2. Compression  ind + W
W b  shsc + m
m  bdhdc  l hg + m
m  bshg + m
 d hg
5. Discharge tube UAdt (Tm,dt − Tie ) chamber
6. Motor UAm (Tm − Tie ) − (Q sta + Q rot + Q iron ) 3. Discharge –  d hg
m  bdhdc + mh
m  dc
7. Internal environment UAh,ie (Tie − Th ) + UAsm (Tie − Tsm ) chamber
− UAw (Tw − Tie ) − UAdc (Tdc − Tie ) + 4. Discharge –  dc
mh  dm
mh
−UAdm (Tm,dm − Tie ) − UAdt (Tm,dt − Tie ) muffler
− UAm (Tm − Tie ) 5. Discharge –  dm
mh  dl
mh
8. Housing UAh,ee (Th − Tee ) − UAh,ie (Tie − Th ) tube
6. Motor – – –
7. Internal –  sl + m
mh  l hg m [ϕ hsl + (1 − ϕ ) hie ]
condition (Todescat et al., 1992). The computed UAs are then environment +m  l hie
maintained fixed in the solution of the compressor tempera- 8. Housing – – –
ture distribution under different operating conditions.
The temperatures in the suction muffler, discharge muffler (Fitzgerald et al., 2003; Hrabovcova et al., 2010). This tech-
and discharge tube, denoted by Tsm, Tm,dm and Tm,dt, respec- nique consists of representing the electrical motor in an
tively, required to close the system of equations, are defined equivalent circuit considering the following set of variables:
as arithmetic means of the gas temperatures at the inlet and stator impedance (Zsta), rotor impedance (Zrot) and magnetiz-
outlet of each component: ing impedance (Zm). The symbols (+) and (−) represent forward
and backward loops, following the theory of rotating mag-
Tsm = 0.5 (Tsl + Tsc ) ; Tm,dm = 0.5 (Tdc + Tdm ) ; Tm,dt = 0.5 (Tdm + Tdl ) netic fields. The impedances, Z, are given by the combination
(7) of real and imaginary parts that correspond to the electrical
resistances, R, and reactances, X, respectively. Fig. 3 shows the
2.3. Electrical model equivalent circuit of the 2-pole single-phase induction motor
considered herein.
The electrical model of the induction motor was based on a The solution for the equivalent circuit of the motor pro-
classical approach known as the equivalent circuit method vides the electrical currents in all the branches:

Fig. 2 – Lumped elements of the compressor.


international journal of refrigeration 59 (2015) 168–181 173

Fig. 3 – Equivalent circuit of a single-phase induction motor.

Ein Ista Z + Ista Z − 1991). However, the adoption of such models entails consid-
+ −
Ista = ; Irot = ; Irot = (8)
Zeq +
Zrot −
Zrot erable computational cost, especially when coupled with other
models (e.g. thermodynamic and thermal models). In this paper,
a semi-empirical method proposed by Stepina (2003) was em-
where currents in the stator main winding and forward and
+ −
ployed to estimate the stator core loss:
backward rotor branches are represented by Ista , Irot and Irot ,
respectively. The equivalent impedance of the circuit, Zeq , is q c
Q core = ∑ kaBib ⎛⎜ n ⎞⎟ mcore,i
f
computed from: (15)
i=1
⎝ 50 ⎠

Zeq = Zsta + Z + + Z − (9) where q is the number of parts into which the stator core is
+ − divided. In this case, the stator core was divided into teeth and
with Z and Z being the forward and backward equivalent
yoke regions, with the corresponding masses mcore,t and mcore,y.
impedances, respectively, obtained through:
The magnetic flux densities, Bt and By, are calculated for each
−1 −1
of these parts following Stepina (2003) and inserted in Eq. (15).
⎛ 1 1 ⎞ ⎛ 1 1 ⎞
Z+ = ⎜ + + + ⎟ ; Z− = ⎜ − + − ⎟ (10) The empirical coefficients a, b and c are dependent on the ma-
⎝ Zrot Zm ⎠ ⎝ Zrot Zm ⎠
terial, dimensions and manufacturing process used to obtain
the core sheets. According to Stepina (2003), a, b and c assume
Once the electrical currents are determined, the torque sup-
values of 3.25, 2.25 and 1.75, respectively, for a normal cold-
plied by the motor to the compressor shaft, τ sh , as well as the
rolled annealed steel of thickness 0.63 mm. The exact values
Joule-effect losses in the stator, Q sta , and in the rotor wind-
for these coefficients can be determined from manufactur-
ings, Q rot , can be calculated through the following equations:
er’s data, but the values proposed by Stepina (2003) were
adopted in this study.

τ sh = ⎛ ⎞ sh
P W
(11) An additional coefficient k was inserted in Eq. (15) in order
⎝ 2⎠ ω
to provide a close match between the predicted and mea-
sured motor efficiencies at a specific load point, defined as
Q sta = Ista
2
Rsta (12) 0.5 N m at 220 V. This coefficient compensates for inaccura-
cies in the calculation of the motor losses (stator, rotor or core)
Q rot = 0.5 ( Irot
+ 2 − 2
+ Irot ) Rrot
′ (13) and includes the effects of stray-load losses, which are very
difficult to predict (Fitzgerald et al., 2003). The adjustment of
In these equations, the superscript ′ indicates that the rotor the parameter k can be considered to be a calibration under
resistance is referred to the stator side (Fitzgerald et al., 2003) a single operating condition, which enhances the prediction
and P is the number of poles. The motor output power (or shaft of the motor performance for a wide range of operating con-
power), W  sh , is given by: ditions. Herein, a value of 2.35 was adopted for k.
Once the motor output power and electrical losses are cal-
′ ′ ⎞  culated, the compressor input power and motor efficiency can
 sh = (1 − s) ⎛ Irot
W + 2 0.5Rrot − 2 0.5Rrot
− Irot − Q core (14)
⎝ s 2−s ⎠ be obtained from the following equations:

The first term on the right-hand side is the rotating mag-  =W


W  sh + Q sta + Q rot + Q core (16)
netic field power and is given by the difference between the
forward and backward components of the rotating magnetic  sh
W
ηele = 
(17)
fields (Fitzgerald et al., 2003). The second term on the right- W
hand side is the stator core loss, Q core , which is strongly
dependent on the magnetic flux densities in the stator core, Equations (11)–(14) and (16)–(17) can only be solved when
B, as well as the nominal frequency, fn. High accurate predic- the motor slip ratio, s, is available. This parameter defines
tions of the core loss require sophisticated electromagnetic the relation between the compressor speed and the nominal
models, based on the finite-element method (Bertotti et al., speed, ωn:
174 international journal of refrigeration 59 (2015) 168–181

ω = (1 − s) ω n (18)

where ωn is given by:

4π fn
ωn = (19)
P

As can be noted from Eq. (18), the greater the motor


slip, the lower the compressor speed will be. The value as-
signed to the slip ratio is obtained from the coupling between
thermodynamic and electrical models, by applying a
steady-state angular momentum balance to the compressor
crankshaft:

τ sh − τ r = 0 (20)

where τ r is the resistive torque. The steady-state resistive torque


originates from two sources. The first is due to the load on
the piston, associated with the pressure in the cylinder during
the compression process and friction between the piston and
the cylinder. The second source for the resistive torque is due
to friction in the compressor bearings (Link and Deschamps,
2011). By substituting the relation W  = ωτ in Eq. (20), we obtain:

 sh − W
f (s) = W r =0 (21)

Equations (5) and (18) are inserted in Eq. (21), which is solved
for the slip ratio via the Brent method (Brent, 1973). This it-
erative procedure requires the user to define a pair of slip ratio
values to initially bracket the solution. Since slip ratios lower Fig. 4 – Electrical model solution procedure.
than 0.001 and higher than 0.2 (0.1% and 20% of nominal speed,
respectively) are rare, the initial intervals were set at 0.001 and
0.2. Brent’s algorithm acts by reducing the bracketed interval
data for the geometrical and electrical parameters, as well as
that contains the solution until a specified tolerance is achieved.
the operating condition and an initial estimate of the com-
The method is guaranteed to achieve convergence, unless the
pressor temperature distribution, Tinit. The electrical resistances
initial interval defined by the user does not bracket the solu-
are then determined and the interval that initially brackets the
tion, in which case the specified motor is not able to provide
slip ratio solution is set: sl and su. The thermodynamic and elec-
the necessary torque to run the compressor under the corre-
trical models, along with Eq. (21), are evaluated using the lower
sponding steady-state condition.
limit of the slip ratio, sl. In the next step, this procedure is re-
An overview of the electrical model solution procedure is
peated using the upper limit of the slip ratio, su. If the sign of
given in Fig. 4. The input data are the input voltage, compres-
f(sl) is equal to the sign of f(su), the bracketing interval does
sor speed, and electrical resistances and reactances. Reference
values for the latter two are estimated from the geometrical pa-
rameters and materials of the motor, through analytical relations
available in Stepina (2003). Table 3 shows some of the motor pa-
Table 3 – Materials and geometrical parameters of the
rameters used in this study. The electrical resistances are motor.
continuously updated during the simulation with information
Stator Rotor
on the motor temperature, using the following expression:
Wire material Copper Conductor Aluminum
R = R0 + β R0 (T − T0 ) (22) bar material
Wire diameter 0.63 mm Number of bars 28
where subscript 0 refers to the reference state defined for a Number of slots 24 Bar cross- 21.5 × 10−6 mm2
motor temperature of 25 °C and β is the temperature coeffi- sectional area
Number of (79; 108; 64; End ring 45 mm
cient, which is a thermal property of the material. Once the
conductors/slota 56; 19; 0) diameter
input data are available, the calculation procedure of the elec-
Active length 48 mm End ring cross- 180 × 10−6 mm2
trical model is straightforward. sectional area
Coil radius 40 mm Air gap between 0.25 mm
2.4. Coupled solution procedure rotor and stator
a
The number of conductors per slot is provided for six slots only
The solution algorithm of the simulation model is shown in
due to the stator 4-quadrant symmetry.
Fig. 5. The procedure starts with the specification of the input
international journal of refrigeration 59 (2015) 168–181 175

Fig. 5 – Coupled solution procedure.

not contain a function zero, which means that the motor does data (mass flow rates, compression power and motor losses)
not supply the torque required to run the compressor, and are supplied to the thermal model, which is then used to cal-
therefore the simulation is aborted. On the other hand, if this culate the temperature of the compressor components, such
is not the case, Brent’s algorithm is activated to calculate a as the suction chamber, cylinder wall and electrical motor. If
reduced interval to bracket the solution. This procedure is re- the maximum residual found in the energy balances is greater
peated from point “**” until the bracketing interval is lower than than a specified tolerance, εt, the temperatures are updated and
a specified tolerance, εth-e. Once convergence is achieved, output calculations start over from point “*”. Otherwise, the simulation
176 international journal of refrigeration 59 (2015) 168–181

is considered to be converged and the solution procedure is an electrical heater (EH) and a thermocouple (TC), the com-
finished. pressor suction line temperature is adjusted to the required
superheated condition (point 1), completing the operating cycle.
Some operating conditions require a heat exchanger (HX) con-
nected in series with the electrical heater (EH) in order to
3. Experimental setup
provide better tuning of the suction line temperature. By ad-
justing the refrigerant charge in the system and the settings
A calorimeter facility was employed to establish the different of the control valves, heat exchanger and heater, the thermo-
operating conditions in which the compressor was tested. As dynamic conditions at points 1 and 2 are set to any required
can be seen in the schematic representation of the calorim- test condition.
eter (Fig. 6a), the experimental setup is composed of pipelines, The first step in the experimental procedure is to submit
control valves (CVs), a mass flow meter (FM), a heat ex- the compressor and the pipeline to an appropriate vacuum con-
changer (HX), an electrical heater (EH), a thermocouple (TC) and dition, in order to remove the air, humidity and any other
pressure transducers (PTs). The calorimeter is designed in such contaminants in the system. After the system receives a charge
a way that the refrigerant fluid can flow through the high and of refrigerant, the compressor is switched on. A period of ap-
low pressure lines in the superheated state, as indicated in the proximately 4 h is needed to establish steady-state operating
pressure–enthalpy diagram of Fig. 6b. conditions due to the thermal inertia of the compressor. During
According to this arrangement, refrigerant enters the com- this period, the control valves in the high and low-pressure lines
pressor (C) at point 1 and is compressed at point 2. The mass have to be continuously adjusted to provide the specified
flow rate is measured with a Coriolis flow meter (FM) and then suction and discharge pressure conditions. The compressor is
the refrigerant releases heat through the pipelines until it considered to have reached steady-state operating condition
reaches point 3, when it is adiabatically throttled to an inter- when the temperatures at several locations of the compres-
mediate pressure level (point 4) by means of a control valve sor vary by less than 1 °C over a period of 1 h. When these
(CV1). The refrigerant is cooled again until it reaches point 5 conditions are satisfied, the local temperatures, mass flow rate
when it is throttled via another control valve (CV2) to give the and compressor input power measurements are acquired over
evaporation pressure (point 6). Finally, with the assistance of a period of 15 min. The main geometrical parameters of the
compressor considered in this study are the piston diameter,
Dp = 21.0 mm, shaft eccentricity, e = 8.0 mm, and connecting rod
length, Lcr = 38.5 mm, which give a volume displacement of
Vsw = 5.5 cm3. This compressor model is adopted for house-
hold or light commercial applications and it operates with R290
as the refrigerant fluid. Nine operating conditions were defined
through the combination of three evaporating temperatures
(−10 °C, −23.3 °C and −35 °C) and three condensing tempera-
tures (45 °C, 54.4 °C and 70 °C), each one being tested four times
to assess the measurement repeatability. The external envi-
ronment temperature, Tee, and suction line temperature, Tsl, were
set to (32 ± 1) °C. The input voltage and nominal frequency were
(228 ± 1.2) V and 50 Hz, respectively, for most operating con-
ditions. Measurement uncertainties are within ±1% for data on
the mass flow rate and compressor input power and ±2 °C for
the temperatures, considering a 95% confidence interval.

4. Results

Before running the simulations, the thermodynamic model


must be calibrated. Basically, this tuning consists of provid-
ing slight variations in the dead volume of the compression
chamber in order to obtain a good match between the predic-
tions and measurements of the mass flow rate and compressor
input power. This calibration was carried out for a single op-
erating condition (ASHRAE LBP rating condition), represented
by the pair of evaporating and condensing temperatures
(−23.3 °C and 54.4 °C). Therefore, after the dead volume is ad-
justed, its value is maintained fixed for the remaining
simulations under a wide range of operating conditions.
Fig. 6 – (a) Schematic of the calorimeter facility and (b) p–h As previously discussed, virtually all compressor simula-
chart of the calorimeter test. tion models use the values for the compressor mean speed as
international journal of refrigeration 59 (2015) 168–181 177

Fig. 7 – Influence of operating condition on volumetric Fig. 8 – Influence of operating condition on isentropic
efficiency: (■) Te = −10 °C; (●) Te = −23.3 °C; (▲) Te = −35 °C. efficiency: (■) Te = −10 °C; (●) Te = −23.3 °C; (▲) Te = −35 °C.

input data or consider the relationship between the compres- measurements when the pressure ratio differs considerably
sor speed and shaft torque. In models in which the motor slip from that of the calibration condition. At high-pressure ratios
is unknown, the nominal speed is adopted. Since the com- given by the operating condition (−35 °C/70 °C), the deviation
pressor speed is lower than its nominal value, the adjusted dead between the predictions and measurements of the volumet-
volume in such models is higher than the value required in ric and isentropic efficiencies, respectively, are 10.2% and 7.8%
models that include the motor slip. for the standard model and 5.3% and 5.6% for the present
Figs. 7 and 8 show the plots for the volumetric and isen- model. However, at low-pressure ratios (Te = −10 °C), no rel-
tropic efficiencies as a function of the evaporating and evant differences between the predictions of the two models
condensing temperatures, obtained with a standard model are observed. The calibration of the dead volume in the stan-
(without motor slip) and with the proposed model (with motor dard model is not very representative because it also includes
slip). The motor efficiency in the standard model is expressed a correction for the motor slip, which varies with the pres-
as a function of the mean shaft power of the compressor, W  sh . sure ratio. Hence, the evaluation of the compressor speed in
The experimental data in these figures (and in other figures the present model allows a more accurate calibration and better
which appear later in the text) are expressed with uncer- predictions for the compressor performance at different pres-
tainty bars corresponding to a 95% confidence interval. The sure ratios.
volumetric efficiency, ην , is defined as the ratio between the The predictions and measurements of the temperature in
actual mass flow rate, m  , and the ideal mass flow rate, m th , the suction chamber, cylinder wall and motor can be ob-
as shown in Eq. (23). The ideal mass flow rate would be ob- served in Figs. 9–11. The first two temperatures considerably
tained in the absence of a cylinder clearance volume, gas affect the gas superheating, which decreases both the volu-
leakages, flow restriction and backflow in the valves, and gas metric efficiency, due to a reduction in the gas density, and the
superheating in the suction process. isentropic efficiency, since the compression work per unit mass
increases with the gas temperature (Kremer et al., 2012). On
m the other hand, the motor temperature is required in order to
ηv = (23) verify the compressor reliability.
 th
m
The predictions of the two models are in good agreement
The isentropic efficiency, ηs , is expressed herein as the ratio with the experimental data for the operating condition in
between the compression power for an isentropic process, which they were calibrated (−23.3 °C/54.4 °C). The predictions
m (h2,s − h1 ) , and the input power of the compressor. become shifted in relation to the measurements when the
evaporating temperature is varied. This occurs because the
 (h2,s − h1 )
m thermal conductances are kept fixed in the model, regardless
ηs =  (24)
W the mass flow rate, which is strongly affected by the evapo-
rating temperature. The maximum values for the difference
The results in Figs. 7 and 8 show similar trends for the two between the prediction of the proposed model and the ex-
models, but the predictions obtained with the model pro- perimental data reach 9.5 °C and 7.7 °C for the suction chamber
posed herein are in better agreement with experimental data and cylinder wall, respectively. The two models provide almost
for most of the operating conditions. Predictions deviate from the same prediction results and the greatest difference between
178 international journal of refrigeration 59 (2015) 168–181

Fig. 9 – Suction chamber temperature. (a) Te = −10 °C. Fig. 10 – Cylinder wall temperature. (a) Te = −10 °C.
(b) Te = −23.3 °C. (c) Te = −35 °C. (b) Te = −23.3 °C. (c) Te = −35 °C.
international journal of refrigeration 59 (2015) 168–181 179

Fig. 12 – Output torque of the motor: (■) 220 V; (●) 255 V.

them (2.2 °C) is observed in relation to the motor tempera-


ture (Fig. 11) under the high-pressure ratio condition
(−35 °C/70 °C). Under this condition, the result for the motor
temperature of the present model is in closer agreement
with the measured value.
In order to validate the electrical model for different input
voltages, measurements were obtained for 220 V and 255 V in
a dynamometer, with the motor temperature controlled at
(81 ± 2) °C. Fig. 12 shows the predictions and the experimen-
tal data for the motor output torque as a function of the
compressor speed. As expected, the results show that the higher
the resistive torque (or motor output torque) the lower the com-
pressor speed due to motor slip will be. Moreover, the good
agreement between the measured and predicted values for the
two input voltages is an indication of the good prediction of
the motor slip.
Fig. 13 shows the results for the motor efficiency as a func-
tion of the torque. The predictions are found to be in good
agreement with experimental data, with the maximum de-
viation being approximately 1.5% for an input voltage of 220 V
and low torque conditions. Fig. 13 also indicates that the motor
efficiency is slightly overestimated under high torque condi-
tions and with an input voltage of 255 V.
Finally, a parametric analysis was carried out to evaluate
the motor temperature as a function of three input voltages
(198 V, 228 V and 255 V) and two operating conditions
(−10 °C/54.4 °C and −35 °C/45 °C). The experimental data for 198 V
and 255 V were obtained from single measurements, and there-
fore uncertainty bars are only provided for 228 V. As can be seen
in Fig. 14, the trends for the predictions are similar to those
observed for the measurements, with the results being almost
the same for the high torque condition (−10 °C/54.4 °C). In ad-
dition, the temperature of the motor remains almost the same
when the voltage is varied from 228 V to 198 V. In this situa-
tion, the model predicts a decrease in the motor efficiency from
78.2% to 77.5%.
Fig. 11 – Motor temperature. (a) Te = −10 °C. (b) Te = −23.3 °C.
On the other hand, the model underpredicts the motor tem-
(c) Te = −35 °C.
perature under the low torque operating condition (−35 °C/45 °C),
180 international journal of refrigeration 59 (2015) 168–181

5. Conclusions

A new approach has been proposed herein to simulate her-


metic reciprocating compressors with a specific model for the
electrical motor, providing predictions for the compression cycle,
temperature distribution and electrical motor performance. An
important feature of the proposed model is the calculation of
motor slip, and therefore the compressor speed. Results
obtained using the model were compared to predictions pro-
vided by a standard model (without motor slip) and with
experimental data. The predictions obtained with the model
proposed herein for volumetric and isentropic efficiencies were
in better agreement (compared with the standard model) with
the measurements for most of the operating conditions. Under
a high-pressure ratio condition (Te = −35 °C; Tc = 70 °C) the de-
viations between the predicted and measured values for the
volumetric and isentropic efficiencies were 10.2% and 7.8% for
the standard model and 5.3% and 5.6% for the proposed model,
Fig. 13 – Motor efficiency: (■) 220 V; (●) 255 V. respectively. In relation to the compressor temperature dis-
tribution, the two models provided close results.
The predictions provided by the model presented herein for
the electrical motor performance parameters, such as torque
especially at 255 V. The experimental data indicate an in-
and motor efficiency, can be considered to be acceptable. The
crease of 5.7 °C in the motor temperature when the input
predicted values for the motor efficiency at different input volt-
voltage is varied from 228 V to 255 V, whereas the model pre-
ages, for instance, were found to deviate from the measured
dicts an increase of 3.5 °C associated with a decrease in the
values only by up to 1.5%. Finally, a parametric study of the
efficiency from 74.5% to 71.1%.
motor temperature as a function of the input voltage and shaft
The parametric analysis of the compressor adopted in this
torque was carried out as an example of procedure to estab-
study indicates that the increase in the input voltage under
lish the application envelope for a reliable motor. For the motor
the low torque condition has a more detrimental effect on the
under study, when input voltage was decreased from 228 V to
motor temperature than the decrease in the input voltage under
198 V under a high torque condition (Te = −10 °C; Tc = 54.4 °C),
the high torque condition. This is an important aspect that has
both the measurements and predictions indicated negligible
to be considered in the motor design and it is a consequence
variation in the motor temperature. However, when the input
of the motor efficiency curves. Therefore, the simulation model
voltage was increased from 228 V to 255 V under a low torque
proposed in this paper can be used to establish the applica-
condition (Te = −35 °C; Tc = 45 °C), the measurements and pre-
tion envelope for a reliable motor.
dictions indicated increases in the motor temperature of 5.7 °C
and 3.5 °C, respectively.

Acknowledgments

This study was developed as part of a technical-scientific co-


operation program between the Federal University of Santa
Catarina and EMBRACO. The authors are also grateful to the
Brazilian governmental agency Conselho Nacional de
Desenvolvimento Científico e Tecnológico (National Council of
Research) for the grant 573581/2008-8 (National Institute of
Science and Technology in Refrigeration and Thermophysics).

REFERENCES

An, K.H., Lee, J.H., Lee, I.W., Lee, I.S., 2002. Performance prediction
of reciprocating compressor. In: Proceedings International
Compressor Engineering Conference at Purdue. West
Lafayette, USA, paper C7-4.
Annand, W.J., 1963. Heat transfer in the cylinders of reciprocating
Fig. 14 – Influence of input voltage on the motor internal combustion engines. Proc. Inst. Mech. Eng. 177, 973–
temperature: (■) −10 °C/54.4 °C; (●) −35 °C/45 °C. 996.
international journal of refrigeration 59 (2015) 168–181 181

Bertotti, G., Boglietti, A., Chiampi, M., Chiarabaglio, D., Fiorillo, F., Lilie, D.E., Ferreira, R.T., 1984. Evaluation of the leakage through
Lazzari, M., 1991. An improved estimation of iron losses in the clearance between piston and cylinder in hermetic
rotating electrical machines. IEEE Trans. Magn. 27, 5007–5009. compressors. In: Proceedings International Compressor
Birari, Y.V., Gosavi, S.S., Jorwekar, P.P., 2006. Use of CFD in design Engineering Conference at Purdue. West Lafayette, USA, pp.
and development of R404a reciprocating compressor. In: 1–6.
Proceedings International Compressor Engineering Link, R., Deschamps, C.J., 2011. Numerical modeling of startup
Conference at Purdue. West Lafayette, USA, paper C072. transients in reciprocating compressors. Int. J. Refrigeration
Brent, R.P., 1973. Algorithms for Minimization Without 34, 1398–1414.
Derivatives. Prentice-Hall, Inc., Englewood Cliffs, NJ. Mantri, P., Bhakta, A., Mallampalli, S., Hahn, G., Kusumba, S., 2014.
Cavallini, A., Doretti, L., Longo, G.A., Rosseto, L., Bella, B., Development and validation of integrated design framework
Zannerio, A., 1996. Thermal analysis of a hermetic for compressor system model. In: Proceedings International
reciprocating compressor. In: Proc. Int. Compress. Eng. Compressor Engineering Conference at Purdue. West
Conference at Purdue, West Lafayette, USA, pp. 535–540. Lafayette, USA, 1153.
Chikurde, R.C., Loganathan, E., Dandekar, D.P., Manivasagam, S., Mellor, P.H., Roberts, D., Turner, D.R., 1991. Lumped parameter
2002. Thermal mapping of a hermetically sealed compressors thermal model for electrical machines of TEFC design. IEE
using computational fluid dynamics technique. In: Proc. B 138, 205–218.
Proceedings International Compressor Engineering Meyer, W.A., Thompson, H.D., 1990. An analytical model of heat
Conference at Purdue. West Lafayette, USA, paper C6-4. transfer to the suction gas in a low-side hermetic
Eltom, A.H., Moharari, N.S., 1991. Motor temperature estimation refrigeration compressor. In: Proceedings International
incorporating dynamic rotor impedance. IEEE Trans. Energy Compressor Engineering Conference at Purdue. West
Convers. 6, 107–113. Lafayette, USA, pp. 898–907.
Fitzgerald, A.E., Kingsley, C.K., Umans, S.D., 2003. Electric Mezani, S., Takorabet, N., Laporte, B., 2005. A combined
Machinery, sixth ed. McGraw-Hill Companies, Inc. electromagnetic and thermal analysis of induction motors.
He, Z., Xing, Z., Chen, W., Wang, X., 2013. Thermal and hydraulic IEEE Trans. Magn. 41, 1572–1575.
analysis on the flow around the motor in semi-hermetic twin Ooi, K.T., 2003. Heat transfer study of a hermetic refrigeration
screw refrigeration compressors. Appl. Therm. Eng. 58, 114– compressor. Appl. Therm. Eng. 23, 1931–1945.
124. Peruzzi, F., Bacci, V., Scandurra, G., 1980. Improvement on a
Hrabovcova, V., Kalamen, L., Sekerak, P., Rafajdus, P., 2010. reciprocating hermetic compressor. In: Proceedings
Determination of single phase induction motor parameters. International Compressor Engineering Conference at Purdue.
In: 2010 International Symposium on Power Electronics West Lafayette, USA, pp. 1–7.
Electrical Drives Automation and Motion (SPEEDAM). pp. 287– Raja, B., Sekhar, S.J., Lal, D.M., Kalanidhi, A., 2003. A numerical
292. model for thermal mapping in a hermetically sealed
Huai, Y., Melnik, R.V., Thogersen, B., 2003. Computational analysis reciprocating refrigerant compressor. Int. J. Refrigeration 26,
of temperature rise phenomena in electric induction motors. 652–658.
Appl. Therm. Eng. 23, 779–795. Ribas, F.A., Jr., 2007. Thermal analysis of reciprocating
Hwang, C.C., Wu, S.S., Jiang, Y.H., 2000. Novel approach to the compressors. Int. Conference on Compress. and their Syst.
solution of temperature distribution in the stator of an London, UK, pp. 277–287.
induction motor. IEEE Trans. Energy Convers. 15, 401–406. Rigola, J., Pérez-Segarra, C.D., Oliva, A., 2000. Advanced numerical
Kara, S., Oguz, E., 2010. Thermal analysis of a small hermetic simulation model of hermetic reciprocating compressors.
reciprocating compressor. In: Proceedings International Parametric study and detailed experimental validation. In:
Compressor Engineering Conference at Purdue. West Proceedings International Compressor Engineering
Lafayette, USA, paper 1307. Conference at Purdue. West Lafayette, USA, pp. 23–30.
Kremer, R., Barbosa, J.R., Jr., Deschamps, C.J., 2012. Cooling of a Sanvezzo, J., Jr., Deschamps, C.J., 2012. A heat transfer model
reciprocating compressor through oil atomization in the combining differential and integral formulations for thermal
cylinder. HVAC&R Res. 18, 481–499. analysis of reciprocating compressors. In: Proceedings
Lemmon, E.W., Huber, M.L., McLinden, M.O., 2002. NIST Standard International Compressor Engineering Conference at Purdue.
Reference Database 23: Reference Fluid Thermodynamic and West Lafayette, USA, pp. 1–11.
Transport Properties-REFPROP, Version 8.0. National Institute Stepina, J., 2003. Single-Phase Induction Motors: Construction,
of Standards and Technology, Standard Reference Data Theory and Calculation. Motorsoft, Inc.
Program, Gaithersburg, USA. Todescat, M.L., Fagotti, F., Prata, A.T., Ferreira, R.T., 1992. Thermal
Li, W., Cao, J., Zhang, X., 2010. Electrothermal analysis of energy analysis in reciprocating hermetic compressors. In:
induction motor with compound cage rotor used for PHEV. Proceedings International Compressor Engineering
IEEE Trans. Ind. Electron. 57, 660–668. Conference at Purdue. West Lafayette, USA, pp. 1419–1428.

View publication stats

You might also like