Sei sulla pagina 1di 30

VOLUME 14, NUMBER 5 HVAC&R RESEARCH SEPTEMBER 2008

719
Modeling of a Two-Stage Rotary Compressor
Margaret M. Mathison James E. Braun, PhD, PE Eckhard A. Groll, PhD
Student Member ASHRAE Fellow ASHRAE Fellow ASHRAE
Received January 21, 2008; accepted May 26, 2008
This paper presents the development and validation of a computer model for hermetic two-stage
rotary compressors that can be used for design optimization. The model considers the effects of
refrigerant leakage between the compressor chambers, heat transfer between the cylinder wall
and the refrigerant gas, and heat loss by natural convection to the surroundings. The model
numerically solves mass and energy balances for each of the compressors chambers to predict
the pressure and temperature variations over a crankshaft revolution. The compressor power
input and refrigerant mass flow rate are calculated and used to determine the isentropic and
volumetric efficiencies. The model can operate both with and without intercooling between the
stages to allow for study of the effects of intercooling on compressor performance..
The model results are compared to the results of external performance tests that were con-
ducted using a prototype compressor. The intermediate temperature and pressure, discharge
temperature, power consumption, and mass flow rate were recorded for 17 different operating
conditions. The model predicts the compressor power consumption and mass flow rate within
5% of the experimental results.
INTRODUCTION
Previous researchers have analyzed the compression process of single-stage rolling-piston
compressors. Several have focused on providing a detailed analysis of the compression chamber
geometry and the motion of the rolling piston in the cylinder (Okada and Kuyama 1982; Yanagi-
sawa et al. 1982). The motion analysis includes a force balance on the compressor, which proves
essential for determining the mechanical efficiency based on frictional losses. Other papers
explore the topic of frictional losses in rolling-piston compressors in more detail, but for this proj-
ect a constant mechanical efficiency was assumed, eliminating the need for a force analysis.
Another topic that many researchers have explored is the refrigerant and oil leakage that
occurs in the compressor. Because leakage interactions between the suction and compression
chambers and the shell can have a large impact on the compressor efficiency, it is a very impor-
tant topic. Yanagisawa and Shimizu (1985a, 1985b) focused on the leakage through the radial
clearance between the roller and the cylinder and leakage across the roller face. Lee and Min
(1988) combined a study of leakage losses and frictional losses to better understand sources of
inefficiencies in the compressor. A similar study on optimal compressor design based on mini-
mizing the effects of leakage and friction losses was performed by Costa (1986). Costa et al.
(1990) also experimentally studied the flow patterns through leakage paths to develop a new
leakage model.
Heat transfer from the cylinder to the refrigerant gas is also a significant source of inefficien-
cies in the compressor and, thus, is the focus of many other studies. No new correlations have
been developed to characterize the heat transfer in a rolling-piston chamber, so researchers have
Margaret M. Mathison is a graduate student and James E. Braun and Eckhard A. Groll are professors in the School
of Mechanical Engineering, Purdue University, West Lafayette, IN.
2008, American Society of Heating, Refrigerating and Air-Conditioning Engineers, Inc. (www.ashrae.org). Published in HVAC&R Research,
Vol. 14, No. 5 (September 2008). For personal use only. Additional reproduction, distribution, or transmission in either print or digital form is not
permitted without ASHRAE's prior written permission.
720 HVAC&R RESEARCH
proposed different methods of modeling this process. Shimizu et al. (1980) suggested using Dit-
tus and Boelters formula for the heat transfer coefficient, while Padhy and Dwivedi (1994)
treated the suction chamber as a circular duct and used a correlation for reciprocating compres-
sors in the compression chamber. Ishii et al. (2000a) focused on the heat transfer from the thrust
plates on the top and bottom of the chamber to the gas. The correlation selected for this project,
originally developed for spiral plate tube heat exchangers, was demonstrated for a scroll com-
pressor by Chen et al. (2002a).
Though several researchers have combined the topics of friction, leakage, and heat transfer
losses to develop models for single-stage rolling-piston compressors, analysis of two-stage com-
pressors is limited. Mechanical friction losses have been considered (Jun 2002), but no analysis
pulls together the friction, leakage, and heat transfer losses for a two-stage model. Because of
the potential for energy savings through intercooling or economizing between stages, the devel-
opment of a two-stage model that can consider these different system configurations is impor-
tant. As the demand for energy-efficient air-conditioning and refrigeration equipment increases
and companies seek to incorporate two-stage compressors into systems, engineers will need
models that can be used to develop optimized two-stage compressor designs.
This paper presents a complete model for a hermetic two-stage rotary compressor. Results
from the simulation model are compared to external compressor measurements, which were
conducted using an available compressor load stand (Chen et al. 2002b). The model was also
used to study the performance of the existing compressor in order to understand the relative
importance of different leakage paths and the impact of intercooling on performance.
MODELING EQUATIONS
Volumes of the Chambers
The rolling-piston compressor uses an eccentric roller contained in a cylinder to form the suc-
tion and compression chambers, which are separated by a vane that extends from the cylinder
wall to the roller surface. The geometry of the rolling piston and cylinder is shown in Figure 1,
with the compression chamber shaded. In this diagram, the roller is drawn to follow a counter-
clockwise path, with the suction port located to the left of the vane and the discharge port to the
right of the vane. The crankshaft angle, , is defined as the angle between the vane slot and the
point of contact between the rolling piston and the cylinder wall. The angle is measured across
the suction chamber. Thus, at small crankshaft angles, which will be considered the beginning of
the crankshaft revolution, the volume of the suction chamber is small.
Because the suction port does not have a valve, gas continuously enters the suction chamber as
its volume increases over an entire revolution of the rolling piston. The volume of the shaded com-
pression chamber, located opposite the suction chamber, decreases as the suction volume
increases. At the beginning of the crankshaft revolution, the compression chamber is open to the
suction port and refrigerant can flow out of the compression chamber to the suction pipe. After the
rolling piston rotates past the suction port, the refrigerant mass is sealed in the compression cham-
ber, and the pressure increases until the valve in the discharge port opens. Refrigerant then flows
through the discharge port to a muffler. From the first-stage muffler, the refrigerant is piped out-
side of the shell to enter the suction pipe of the second stage. If the compressor is operating with
intercooling, the refrigerant will also pass through a cooling coil before entering the second-stage
suction pipe. If the compressor is operating with economizing, saturated vapor or two-phase refrig-
erant mixes with the first-stage discharge gas before entering the second-stage suction chamber.
Economizing was not considered in the study presented here. From the second-stage muffler, the
refrigerant enters the high-side shell, flows over the motor, and exits through a discharge pipe at
the top of the shell. For this analysis, the gas inside the shell is separated into two control volumes;

VOLUME 14, NUMBER 5, SEPTEMBER 2008 721


the volume below the motor, which includes the volume surrounding the compression cylinders,
will be called the lower cavity, while the volume surrounding and above the motor will be called
the upper cavity. The shell is divided into these control volumes for the purpose of modeling the
heat transfer to gas in the upper cavity due to motor and mechanical inefficiencies.
The volume of the compression chamber can be calculated using the known dimensions of the
compressor and the calculated vane extension:
(1)
where the distance that the vane extends into the cylinder, x, can be calculated as
(2)
and
(3)
The resulting variation of volume with crankshaft angle is shown in Figure 2. For both stages,
the beginning of the crankshaft revolution, when the compression chamber volume is at a maxi-
mum, corresponds to an angle of 0. However, it is important to note that the two stages are 180
out of phase. This must be taken into consideration when linking the two stages in the model.
It is also necessary to know the rate at which the chamber volumes change with respect to
crankshaft angle, which can be determined by taking the derivative of Equation 1:
Figure 1. Compression chamber geometry for calculating volume.
V H
c
R
c
2
R
r
2
( )
H
c
2
------ R
c
2
R
r
2
+ ( ) [ ] =
H
c
2
------e R
r
r
v
+ ( )sin + ( )
H
c
2
------r
v
2
tan
H
c
2
------bx +
x R
c
r
v
R
r
r
v
+ ( )cos ecos + =
sin
1
e
R
r
r
v
+
----------------sin


=
722 HVAC&R RESEARCH
(4)
where
(5)
(6)
(7)
Surface Area of the Chambers
The surface area of the suction and compression chambers must also be calculated for use in the
heat transfer calculations. Figure 3 shows the suction and compression chambers with a new vari-
able, , defined to measure the distance between the cylinder wall and the roller at any angle, .
This distance is both a function of the crankshaft angle, , and the angle at which it is mea-
sured, :
(8)
Figure 2. Variation of chamber volume with crankshaft angle.
dV
d
------ - H
c
1
2
--- R
c
2
R
r
2
R
r
2

+ + ( ) e
R
r
r
v
+
2
----------------


cos + ( ) 1

+ ( ) +

r
v
2
2 cos ( )
2
----------------------

b
2
--- R
r
r
v
+ ( ) sin ( )

ebsin
=
z e
sin
R
r
r
v
+
---------------- =
sin
1
z =

1
1 z
2

------------------ =

R
c
2ecos ( ) 4e
2
cos
2
( ) 4 e
2
R
r
2
( )
2
---------------------------------------------------------------------------------------------------------------- =
VOLUME 14, NUMBER 5, SEPTEMBER 2008 723
The area of the chamber on the top and bottom of the cylinder is then calculated by numeri-
cally integrating the distance between the cylinder wall and the roller across the entire chamber.
This area is added to the area of the vertical surfaces to determine the total surface area of each
chamber:
(9)
(10)
Chamber Conservation of Mass
A mass balance can be written for the gas within the suction and compression chambers, muf-
flers, and upper and lower cavities of the shell as follows:
(11)
By application of the chain rule for differentiation, and assuming that density is a function of
temperature and pressure, the mass balance can be rewritten in terms of the unknowns,
and :
(12)
Figure 3. Compressor geometry for calculating surface area.
A
s
R
r
R
c
+ ( )
i
d ( )
i 0 =

+ =
A
c
2 ( ) R
r
R
c
+ ( )
i
d ( )
i =
2

+ =
d V ( )
dt
--------------- m

in
m

out

=
dP d
dT d
V

P
------
dP
d
------ -

T
------
dT
d
------ +



dV
d
------- m

in
m

out

( )
1

---- + =
724 HVAC&R RESEARCH
Chamber Conservation of Energy
An energy balance for the gas in the suction and compression chambers, mufflers, and upper
and lower cavities of the shell can be written as follows:
(13)
The enthalpy terms in the energy balance account for the energy transfer by gas flow through
the suction pipe, discharge pipe, and leakage paths. Assuming that the compression process is a
quasi-equilibrium process, the changes in kinetic and potential energy are negligible, and the
specific internal energy is a function of pressure and temperature, the energy balance can be
rewritten as follows:
(14)
Chamber Leakage Model
Leakage occurs to and from the suction and compression chambers and the upper and lower
cavities of the compressor at several locations, causing an overall decrease in efficiency due to
the re-expansion of compressed gas. Figure 4 shows a sketch of a single stage of the compressor
with arrows marking the flow paths, including leakage, inlet, and outlet flows. A description of
each flow path is included in Table 1 along with a classification of the type of flow through the
path. The types of flow are classified as either isentropic flow of compressible ideal gas, laminar
viscous flow, or mixed plane Couette and Poiseuille flow.
Isentropic Flow Model. For the case of isentropic flow, the mass flow rate is dependent on
the pressure ratio across the flow path, :
dE
cv
dt
----------- m

i n
h
i n

out
h
out

+ =
uV

P
------ V
u
P
------ +


dP
d
------ - uV

T
------ V
u
T
------ +


dT
d
------ + u P + ( )
dV
d
------ - m

in
h
i n
m

out
h
out
Q

+ ( )
1

---- + =
Figure 4. Schematic of leakage paths in the compression cylinder.
P
r
VOLUME 14, NUMBER 5, SEPTEMBER 2008 725
(15)
However, the pressure ratio must be greater than or equal to the critical pressure ratio that
occurs at choked flow. The critical pressure ratio depends on the specific heat ratio of the gas, :
(16)
The mass flow rate also depends on flow path area, , and the high-side pressure and tempera-
ture, and , respectively:
(17)
Mixed Couette and Poiseuille Flow Model. The assumption of mixed Couette and Poi-
seuille flow is used to model the flow along the sides of the vane in the vane slot because the
effect of the vanes motion on the leakage flow must be considered. The motion of the vane
can increase or decrease the leakage flow rate either by acting in the same direction as the
pressure-driven flow or by opposing the pressure-driven flow. Therefore, the velocity of the
vane must be determined as follows:
(18)
Table 1. Summary of Leakage Paths
1
Path From Flow Type
m
ec
From back of vane to compression chamber through clearance between vane and slot 2
m
rb
From shell to suction chamber through vertical clearance
between rolling piston and cylinder
3
m
rc
From shell to compression chamber through vertical clearance
between rolling piston and cylinder
3
m
vb
From compression to suction chamber through vertical clearance
between vane and cylinder
1
m
vt
From compression to suction chamber through radial clearance
between vane tip and rolling piston
1
m
21
From suction chamber to suction pipe 1
m
31
From compression chamber to suction pipe 1
m
32
From compression to suction chamber through radial clearance
between cylinder wall and rolling piston
1
m
42
From clearance volume to suction chamber 1
m
46
From clearance volume to muffler 1
1
Flow is designated as (1) isentropic flow of compressible ideal gas, (2) mixed plane Couette flow and Poiseuille flow,
or (3) laminar viscous flow into the chamber and isentropic flow out of the chamber.
P
r
P
l
P
h
------ =
k
P
r
P
r
( )
crit ical

2
k 1 +
----------- -


k
k 1
-----------
=
A
P
h
T
h
m

AP
h
2k
k 1 ( )RT
h
-------------------------- P
r
2
k
---


P
r
k 1 +
k
------------

vane
esin

+ ( ) =
726 HVAC&R RESEARCH
where
(19)
and
(20)
Then the mass flow rate is a function of the vane velocity and the high- and low-side pres-
sures, and , respectively:
(21)
where is the height of the vane slot and c is the clearance between the vane and the vane slot.
Laminar Viscous Flow Model. When leakage occurs across the top of the rolling piston into
either the suction chamber or the compression chamber, the model of laminar viscous flow is
used. This model is applied because of the high concentration of oil on the top of the rolling pis-
ton. The mass flow rate in this case depends on the vertical clearance between the top of the roll-
ing piston and the cylinder, ; the ratio of the outer and inner radii of the rolling piston, and
, respectively; and the mean viscosity of the high- and low-side fluids, :
(22)
where
(23)
and
(24)
Valve Model
The acceleration of the valve, used to solve for the valve position, is predicted using the pres-
sure difference between the compression chamber and the muffler. The acceleration of the valve
at the point in time depends on its displacement, , and velocity, , at time :
(25)

ecos

R
r
R
b
+ ( ) z
------------------------------ =
z 1
esin
R
r
R
b
+
------------------


2
=
P
h
P
l
m

h
c
3
12
---------
P
h
P
l
( )
l
---------------------- c

vane
2
------------- +


=
h
c R
r
R
r i ,

mean
m

2
h
c
3
P
h
P
l

6
mean
log R
r
R
r i ,
( )
--------------------------------------------------
2
2
---------------


=
sin
1
e
R
r
-----sin


=
+ =
t t + y
n
y
n
t
y
n 1 +
F
f
1.14D
dp
( )
2
4
-------------------------------
P
muffl er
P
compression

m
valve
1 0.5
y
n
y
max
-----------


-------------------------------------------------------- 2C
d
k
val ve
m
val ve
1 0.5
y
n
y
max
-----------


-------------------------------------------------- y
n
=

k
val ve
m
val ve
1 0.5
y
n
y
max
-----------


-------------------------------------------------- y
n

VOLUME 14, NUMBER 5, SEPTEMBER 2008 727


The maximum displacement of the valve, ; the mass of the valve, ; and the diame-
ter of the discharge port, are all known properties of the compressor. The friction factor, ,
and the damping coefficient, , are determined experimentally. The spring constant for the
valve, , depends on the position of the valve:
(26)
where the coefficients , , and must also be determined experimentally. Then, the velocity
of the valve is given as follows:
(27)
(28)
If the displacement of the valve is nonzero, then mass can flow through the discharge port, and
the valve displacement is used to calculate an effective cross-sectional area for the mass flow:
(29)
The isentropic flow model is then applied to calculate the mass flow rate based on the pres-
sure across the discharge port and the effective cross-sectional area.
Heat Transfer Model
Heat transfer between the cylinder wall and the gas in the suction and compression chambers
is calculated by determining an appropriate convection coefficient and applying Newtons law
of cooling. The spiral heat exchanger model that has been applied to scroll compressors (Chen et
al. 2002a; Yi et al. 2004) was selected for modeling the heat exchange in the rolling-piston suc-
tion and compression chambers. The spiral heat exchanger model relates the heat transfer coeffi-
cient to the Reynolds number, Re, and Prandtl number, Pr, by the following expression:
(30)
where is the hydraulic diameter of the chamber and is the average radius of the cham-
ber. The Reynolds number and Prandtl number must be calculated at each crankshaft position
because of their dependence on temperature and pressure. The velocity of the gas, , used in the
Reynolds number can be approximated as the average surface velocity of the cylinder and the
roller, which is constant:
(31)
The hydraulic diameter of the chamber is a function of the volume, , and surface area, , of
the chamber:
y
max
m
val ve
D
dp
F
f
C
d
k
valve
k
valve
a exp b y
n
( ) c + =
a b c
y
n 1 +
y
n
y
n
t + =
y
n 1 +
y
n
y
n
t
y
n
t ( )
2
2
--------------------- + + =
A
val ve
1.14D
dp
( )
2
2
------------------------------- 1.5
1.14D
dp
( )
2
2
-------------------------------
D
dp
y
-------------------------------
2





1 2 /
=
h
c
0.023
k
D
h
------Re
0.8
Pr
0.4
1.0 1.77
D
h
r
aver
-----------


+ =
D
h
r
aver
u
u
1
2
--- 2R
r
rps ( ) =
V A
728 HVAC&R RESEARCH
(32)
Once the heat transfer coefficient is determined, Newtons law of cooling states that the heat
transfer rate will be proportional to the temperature difference between the gas and the surface:
(33)
Newtons law of cooling also applies to the heat transfer from the outer surface of the com-
pression cylinder to the gas in the shell. However, for this case the correlation for forced convec-
tion in an annulus is used to predict the heat transfer coefficient, , and the hydraulic
diameter is defined as follows:
(34)
The Reynolds number is used to determine the regime of flow according to the following
transition points:
(35)
(36)
where
(37)
Then the convection coefficient is determined for the laminar, transition, and turbulent
regimes as follows:
(38)
(39)
(40)
where
(41)
The heat transfer from both the interior and exterior walls of the cylinder depends on the cyl-
inder temperature. For these calculations, the temperature of the cylinder surface, , is
assumed to have a linear distribution around the cylinder, varying from 5 K above the average
cylinder temperature near the discharge port to 5 K below the average cylinder temperature near
the suction port. However, since the average cylinder temperature is initially unknown, solving
D
h
4V
A
------- =
Q

h
c
A T
c
T
g
( ) =
h
c outer ,
D
h out ,
D
shell
2R
c
=
Re
lam
2089.26 686.15R

+ =
Re
turb
2963.02 3343.16R

+ =
R

2R
c
D
shell
-------------- =
Nu
lam
0.186 0.029logR

0.008 logR

( )
2
+ [ ]
1
for Re Re
lam
< =
Nu 0.025Re
0.78
Pr
0.48
R

( )
0.14
for Re Re
t urb
> =
Nu exp log Nu
lam
log Nu
turb
log Nu
lam
( )
log Re log Re
lam

log Re
t urb
log Re
l am

------------------------------------------------------ +


for Re
lam
Re Re
turb
< < =
Nu
turb
0.025Re
t urb
0.78
Pr
0.48
R

( )
0.14
=
T
c
VOLUME 14, NUMBER 5, SEPTEMBER 2008 729
for the heat transfer is an iterative process. An initial average cylinder temperature is assumed
and used to calculate the heat transfer; it is then adjusted between iterations until the heat trans-
fer into the cylinder from the gas in the shell is equal to the heat transfer out of the cylinder to
the gas in the suction and compression chambers.
Intercooling Model
When intercooling is incorporated, the model uses a specified degree of superheat at the
second-stage suction to control the amount of intercooling. The model calculates the satura-
tion temperature corresponding to the pressure in the first-stage muffler and adds the specified
superheat to determine the temperature of the gas available to the second stage. If repre-
sents the temperature of the gas entering the second stage, is the intermediate pressure,
and is the degree of superheat:
(42)
where .
The required capacity of the intercooler to achieve the specified superheat is determined by
multiplying the change in the enthalpy of the fluid across the intercooler by its mass flow rate:
(43)
Shell Energy Balance
A simple shell energy balance is applied to the compressor to estimate the temperature of the
gas in the shell. For these calculations, it is assumed that all of the compressor power lost due to
mechanical and electrical inefficiencies is transferred to the gas in the shell as heat:
(44)
It is also assumed that the outer surface of the compressor is at the same temperature as the
gas in the shell. Therefore, the amount of heat lost to the environment due to natural convection
is dependent on the temperature of the gas in the shell according to Newtons law of cooling:
(45)
The heat transfer coefficient for natural convection from a vertical cylinder can be approxi-
mated as follows:
(46)
where the properties of air are evaluated at the film temperature, which is the average of the
ambient and shell temperatures. The Raleigh number, , is determined according to the fol-
lowing expression:
(47)
T
int
P
i nt
T
sh
T
int
T
sat
T
sh
+ =
T
sat
T P
int
X = 1 , ( ) =
Q

i nt
m

h
int
=
Q

i n
P
overall
1
mech

motor
( ) =
Q

out
h
c
A T
shell
T
amb
( ) =
h
c
k
H
shel l
-------------- 0.825 0.387Ra
L
1 6 /
1
0.492
Pr
-------------


9 16 /
+
8 27 /
+



2
=
Ra
L
Ra
L
g T
shel l
T
amb
( )H
shel l
3
T
f ilm

fil m

fi lm
------------------------------------------------------- =
730 HVAC&R RESEARCH
Because the enthalpy and mass flow rate of the gas entering the shell from the second-stage
compression chamber are known, the energy balance can be solved for the enthalpy of the gas
leaving the shell:
(48)
The mass flow rate of gas leaving the shell is known because it must equal the mass flow rate
of gas entering the shell at steady-state conditions. However, both the enthalpy of the gas leav-
ing the shell and the rate of heat transfer from the shell due to natural convection depend on the
shell temperature. Therefore, determining the shell temperature is an iterative process.
NUMERICAL SOLUTION OF THE MODEL
For each control volume, combining the final mass and energy balance equations in matrix
form yields the following:
(49)
Solving this set of equations using Cramers Rule and replacing the internal energy terms
with the relation results in the following expression for the differential temperature
and pressure, respectively:
(50)
(51)
where
(52)
and
(53)
The differential terms in Equations 50 and 51 can be approximated using the centered-difference
formula. If is the derivative of property with respect to ,
h
out
m

h ( )
in
Q

in
Q

out
+
m

out
------------------------------------------------ =
uV

P
------ V
u
P
------ +


uV

T
------ V
u
T
------ +


V

P
------ V

T
------





dP
d
------ -
dT
d
------





u P + ( )
dV
d
------ - m

i n
h
in
m

out
h
out
Q

+ ( )
1

---- +

dV
d
------ - m

i n

out

( )
1

---- +





=
u h Pv =
dT
d
------

2
h
P
------
2
P
v
P
------ P

P
------ + + +


V

------ v
m

-------



P
------
m
in

----------- h
out
h
i n
( )
1

----Q

+
V
h
P
------ 1



T
------
h
T
------
P
T
------



P
------
---------------------------------------------------------------------------------------------------------------------------------------------------------------------------------- =
dP
d
------ -

2
h
T
------
2
P
v
T
------
P
T
------ P

P
------


V

------ v
m

-------



T
------
m
in

----------- h
out
h
in
( )
1

----Q


V
h
P
------ 1



T
------
h
T
------
P
T
------



P
------
------------------------------------------------------------------------------------------------------------------------------------------------------------------------------------ =
dm
d
-------
m

in
m

out

( )
dt
d
------ =
dm
in
d
-----------
m

i n
dt
d
------

=
f
x
f x
VOLUME 14, NUMBER 5, SEPTEMBER 2008 731
(54)
Before applying the mass and energy balance in the form shown in Equations 50 and 51, it is
necessary to have an initial guess of the variation of temperature and pressure in the chambers
with crankshaft angle for evaluating the mass flow rates through the various leakage paths and
estimating the heat transfer. To obtain the initial guess values, the model assumes constant tem-
peratures and pressures in the suction chambers and mufflers. The first-stage muffler pressure is
estimated as the intermediate pressure for an ideal two-stage compressor:
(55)
Then, the temperature in the first-stage and second-stage mufflers are estimated by assuming
an isentropic process:
and (56)
The model starts by assuming that when the crankshaft angle, , is zero (at the end of the suc-
tion process and the start of the compression process), the first-stage compression chamber is at
suction temperature and pressure. The remaining properties, such as density, enthalpy, and vis-
cosity, can be evaluated using the temperature and pressure. Then, the derivatives of tempera-
ture and pressure with respect to crankshaft angle at the crankshaft angle of zero can be
estimated by neglecting leakage and heat transfer:
(57)
(58)
The derivatives of pressure and temperature with respect to crankshaft angle evaluated at
angle are then used to predict the pressure and temperature at angle using the modified
Euler method. The modified Euler method first calculates a predicted temperature and pressure
at angle :
(59)
(60)
The derivatives of pressure and temperature with respect to crankshaft angle are then reevalu-
ated using the predicted pressure and temperature at angle . The pressure and temperature
at angle are then recalculated using an average of the derivatives:
f
x

n f
n 1 +
f
n 1

2x
---------------------------- =
P
i nt
P
sucti on
P
di sch e arg
=
T
int
T
suction
P
i nt
P
suction
-------------------


k 1 ( ) k
= T
d
T
sucti on
P
d
P
sucti on
-------------------


k 1 ( ) k
=

dT
d
------

2
h
P
------
2
P
v
P
------ P

P
------ + + +


V

------


V
h
P
------ 1



T
------
h
T
------
P
T
------



P
------
-------------------------------------------------------------------------------------- =
dP
d
------ -

2
h
T
------
2
P
v
T
------
P
T
------ P

P
------


V

------


V
h
P
------ 1



T
------
h
T
------
P
T
------



P
------
---------------------------------------------------------------------------------------- - =
+
+
T
P
+ ( ) T ( )
T

------ ( ) + =
P
P
+ ( ) P ( )
P

------ ( ) + =
+
+
732 HVAC&R RESEARCH
(61)
(62)
Using the modified Euler method, it is possible to start with the initial guess of temperature
and pressure at the crankshaft angle of zero in the compression chamber and step through the
crankshaft revolution to predict the temperature and pressure at each angle. The valve subrou-
tine is also called at each angle to determine when the discharge process begins; it uses the pre-
dicted temperature and pressure in the compression chamber along with the muffler pressure and
temperature to determine if the valve will open. As soon as the valve opens, it is assumed that a
constant pressure discharge process occurs. The process is repeated for the second stage.
After the initial guesses of pressure and temperature with crankshaft angle are obtained, the
model starts again with the first-stage suction chamber, this time applying the mass and energy
balance in Equations 50 and 51 instead of assuming constant temperatures and pressures in the
suction chambers and mufflers. At each angle, the leakage flow rates are evaluated using the
pressure and temperature in the chamber and the surrounding control volumes. The instanta-
neous heat transfer rate is also evaluated and then substituted into the calculations for the deriv-
atives of pressure and temperature with respect to crankshaft angle.
The suction processes no longer occur at a constant temperature or pressure. The mass and
energy balance is applied to the first-stage suction chamber and then the first-stage compression
chamber, assuming a constant temperature and pressure in the first-stage muffler. The mass flow
rate into the first stage is then compared to the mass flow rate exiting the first stage, and the tem-
perature in the muffler is adjusted. Without changing the first-stage muffler pressure, the model
continues to loop through the first-stage calculations until the first-stage inlet and outlet mass
flow rates agree within a specified tolerance. This fixes the first-stage muffler temperature.
The model then repeats the process for the second stage using the constant temperature and
pressure in the first-stage muffler as an input. The second-stage muffler pressure is held constant
at the discharge pressure while the second-stage muffler temperature is varied. The model loops
through the second-stage calculations until the muffler temperature is such that the inlet and out-
let mass flow rates for the second stage agree within a specified tolerance.
The temperature of the gas in the shell is assumed to be equal to the discharge temperature from
the second stage. Because leakage occurs from the shell to the first stage, and the heat transfer to
the first stage depends on the shell temperature, the calculations must now be repeated with the
updated discharge temperature. The muffler pressure is not updated until the discharge tempera-
ture does not change significantly between iterations.
When the discharge temperature converges, the muffler pressure is updated to attempt to sat-
isfy an overall mass balance. The first-stage muffler pressure is adjusted until the mass flow rate
through the two stages agrees within the specified tolerance. As mentioned previously, the tem-
perature of the gas in the shell must also be determined in an iterative process. Therefore, the
model will only converge when the change in the shell temperature between iterations is less
than a specified tolerance. Figure 5 provides a flowchart of the program showing the iterative
process.
After the program has converged, the calculated pressure and temperature variations are used
to determine the power consumption of each stage. The suction and compression chambers can
be approximated as closed systems with constant pressures over the very small time steps used
T + ( ) T ( )

2
-------
T

------ ( )
T

------ + ( ) + + =
P + ( ) P ( )

2
-------
P

------ ( )
P

------ + ( ) + + =
VOLUME 14, NUMBER 5, SEPTEMBER 2008 733
Figure 5. Flowchart of two-stage rolling-piston compressor model.
734 HVAC&R RESEARCH
in the program. Therefore, the work required for each stage is the sum of the boundary work for
each time step over one revolution:
(63)
where the total efficiency, , is the product of the motor and mechanical efficiencies,
and , respectively. Then, the isentropic efficiency can be calculated by com-
paring the actual power input to each stage to the power required for an isentropic compression
process between the same pressures:
(64)
The volumetric efficiency, , can also be calculated from the model results. The actual mass
flow rate predicted by the model is compared to the maximum theoretical mass flow rate, which
is based on the maximum suction volume of the stage, ; the density of the gas entering the
stage, ; and the number of revolutions per second made by the crankshaft, :
(65)
RESULTS AND DISCUSSION
Model Results
The model was tuned to accurately predict discharge temperature, mass flow rate, and power
consumption based on the external measurements taken at test conditions similar to those
encountered in air-conditioning applications. Test Condition 1 will be used to reference an evap-
orating temperature of 10.1C, a suction gas temperature of 33.5C, and a condensing tempera-
ture of 55.3C for the remainder of this paper. The parameters that were tuned include the
damping coefficients of the valve model, the combined motor and mechanical efficiency, the
friction coefficient for the largest leakage path, and the heat transfer coefficient between the
compressor and the surroundings.
Without tuning, the model predicted mass flow rates about 3% lower than those measured.
To address this discrepancy, the mass flow rate through the most significant leakage path, the
predicted flank leakage from the compression chamber to the suction chamber, was multiplied
by a correction factor of 0.43. This value was selected to achieve agreement between the exper-
imental and modeled mass flow rates through the compressor at Test Condition 1. Although
this is a significant reduction in the leakage flow rate, several variables contribute to inaccu-
racy in the leakage model, including the assumption that the effect of oil is insignificant in
most of the leakage paths. Only the calculations for leakage over the top of the rolling piston
and leakage from the back of the vane slot consider the effect of oil on the density and viscosity
of the fluid in the flow path, but the presence of oil will affect the flank leakage as well. There-
fore, the tuning factor approximates the reduction in leakage achieved when oil seals the con-
tact between the rolling piston and cylinder. The damping coefficients in the valve model also
influence the models ability to predict mass flow rate. These coefficients were selected so that
the valve opened at the angle expected by the compressor designers and exhibited the damping
that would be expected in the valves motion.
W
actual
W
s
W
c
+

t ot al
--------------------
P
s
i ( ) V
s
i 1 + ( ) V
s
i ( ) [ ] P
c
i ( ) V
c
i 1 + ( ) V
c
i ( ) [ ] +

motor

mechani cal
---------------------------------------------------------------------------------------------------------------------------
i

= =

total

motor

mechanical

s
m

h
2s
h
1
( )
W

actual
---------------------------- =

v
V
max

s
rps

v
m

actual
V
max

s
rps
----------------------------------- =
VOLUME 14, NUMBER 5, SEPTEMBER 2008 735
Before tuning, the model predicted power consumptions approximately 2% lower than the
measured values. Modifying the model to increase the mass flow rate improved the models
power consumption calculations. To further improve the models accuracy at Test Condition 1,
the combined mechanical and motor efficiency of the compressor was modified from 76.5% to
78.2%. These two modifications also improved the models ability to predict the discharge tem-
perature, which was over-predicted by 4 K before tuning.
While tuning the model improved the accuracy of its predictions of discharge temperature, mass
flow rate, and power consumption, it is interesting to note that the important parameters of isen-
tropic and volumetric efficiency are less sensitive to the tuning factors. Figures 6 through 8 investi-
gate the effect of motor efficiency, the flank leakage coefficient, and the shell heat transfer
coefficient on isentropic efficiency, volumetric efficiency, and compressor discharge temperature.
The tuning factors were varied over different ranges based on the expected accuracy of the param-
eter as well as the sensitivity of the model to the parameter. For example, because the combined
motor and mechanical efficiency can be estimated with reasonable accuracy and is expected to
have a large impact on the results, it was only varied by 15%, whereas the heat transfer coeffi-
cient, which can be predicted with less confidence, was varied by 50%. Therefore, the horizontal
axis uses a variable multiplier that represents the percentage of the range over which the tuning
factor is varied. For example, the results for multiplying the combined efficiency by 0.85 are
shown as 100% of the variable multiplier, as are the results for multiplying the heat transfer coef-
ficient by 0.50.
As expected, Figure 6 shows that increasing the flank leakage coefficient results in lower volu-
metric efficiencies. However, varying the flank leakage by 75% only results in a 4% variation
in volumetric efficiency. In addition, the effects of combined motor and mechanical efficiency
and heat transfer on volumetric efficiency are negligible. Therefore, the models prediction of
Figure 6. Variation of volumetric efficiency with 15% variation in combined motor and
mechanical efficiency, 75% variation in flank leakage coefficient, and 50% variation in
heat transfer coefficient.
736 HVAC&R RESEARCH
Figure 7. Variation of isentropic efficiency with 15% variation in combined motor and
mechanical efficiency, 75% variation in flank leakage coefficient, and 50% variation in
heat transfer coefficient.
Figure 8. Variation of the exit temperature with 15% variation in combined motor and
mechanical efficiency, 75% variation in flank leakage coefficient, and 50% variation in
heat transfer coefficient.
VOLUME 14, NUMBER 5, SEPTEMBER 2008 737
volumetric efficiency can be stated with confidence. As seen in Figure 7, the isentropic efficiency
is more sensitive to tuning factors, showing a direct dependence on motor and mechanical effi-
ciency. To a much lesser degree, the isentropic efficiency also depends on flank leakage, decreas-
ing by about 4% as the flank leakage multiplier is increased from 0 to 1.18. Figure 8 shows that
the variation in the motor and mechanical efficiency has the greatest impact on the compressor
exit temperature, causing the temperature to decrease by 18 K as the efficiency increases from
15% to +15% of its nominal value. In addition, the exit temperature decreases with increasing
heat transfer and increases with increasing flank leakage, but these variations are less than 2 K.
The effect of the convergence criteria on the model results was also studied. For the model to
converge, the percent difference between the mass flow rates into and out of the compressor
shell has to be less than the specified percent tolerance. While the model results in this paper
were obtained using a 0.1% tolerance, raising the tolerance to 5% does not change any of the
output parameters by more than 0.5%. Thus, the results of the model can be used with reason-
able confidence.
The variation of pressure with crankshaft angle in the suction and compression chambers pre-
dicted by the tuned model is shown in Figure 9. The results are plotted with the beginning of the
compression process for each stage shown as 0. However, it is important to remember that the
two stages operate 180 out of phase. The variation of temperature with crankshaft angle for the
same compressor at the same operating conditions is shown in Figure 10. The graph shows that
the gas in both the first- and second-stage compression chambers experiences a significant rise
in temperature at the beginning of the suction process (at a crankshaft angle of approximately
10) due to heat transfer to the small amount of gas that is present in the suction chamber before
the suction port is open to the chamber. The temperature in the chamber drops quickly as soon as
mass begins to flow in through the suction pipe. Leakage from the high-pressure shell also con-
tributes to the rise in temperature.
The heat transfer to the cylinder gas is a function of the heat transfer coefficient, the tempera-
ture difference, and the surface area. The instantaneous heat transfer rate is plotted as a function
of crankshaft angle in Figure 11, and the average heat transfer rate, obtained by integrating the
instantaneous rate, is summarized in Table 2. The heat transfer rate to the gas in the suction
chambers increases with crankshaft angle as expected due to the increase in surface area. Simi-
larly, the heat transfer rate to the gas in the compression chambers decreases with crankshaft
angle due to both decreased surface area and a decrease in the temperature difference driving the
heat transfer. The heat transfer even becomes negative at the end of the second-stage compres-
sion process, indicating that heat is transferred from the gas to the cylinder wall, due to the high
gas temperature.
Table 2. Average Heat Transfer Rate to Refrigerant over One Revolution
Predicted for Prototype Compressor at Test Condition 1
Chamber Average Heat Transfer Rate to Refrigerant, W
First Stage
Suction 9.95
Compression 8.74
Second Stage
Suction 20.3
Compression 19.8
738 HVAC&R RESEARCH
Figure 9. Variation of pressure with crankshaft angle predicted for prototype compressor
at Test Condition 1.
Figure 10. Variation of temperature with crankshaft angle predicted for prototype
compressor at Test Condition 1.
VOLUME 14, NUMBER 5, SEPTEMBER 2008 739
Leakage also has a significant impact on compressor performance. Figure 12 provides a com-
parison of the total mass flow rate through each path in the first stage over an entire revolution.
It can be seen that the most significant source of leakage is through the contact between the
roller and cylinder wall, m
32
. The leakages across the top and bottom of the vane, m
vb
, are also
significant. By summing the mass flow rate through all of the leakage paths in Figure 12, it can
be seen that the leakage accounts for about 3.95% of the total mass flow rate. The most signifi-
cant leakage path in the compression chamber is flank leakage at the point of contact between
the roller and cylinder wall.
When intercooling is incorporated, the mass flow rate through the first stage does not change
significantly because the first-stage suction state has not changed. Therefore, the mass flow rate
through the second stage and, thus, the density of the gas entering the second stage should not
change significantly when intercooling is incorporated. To achieve the same gas density at the
second-stage suction when the intermediate temperature is lower due to intercooling, the inter-
mediate pressure becomes lower. Figure 13 shows that the intermediate pressure predicted by
the model when intercooling is used to maintain a 10 K superheat is lower than the case without
intercooling, as expected. The temperature profiles with and without intercooling are compared
in Figure 14.
The main benefit of intercooling is reduced power consumption, which is summarized in
Table 3 by comparing the cases with and without intercooling. There is almost a 10% reduction
in power requirement for intercooling at these conditions. The primary improvement results
from increased density of the suction gas for the second stage that leads to lower isentropic work
along with an improvement in the volumetric efficiencies. The isentropic efficiencies actually
decrease slightly for the intercooled case, a trend that can be explained by differences in leakage
flow rates between the standard and intercooled cases. However, the lower isentropic efficien-
cies are more than offset by reduced isentropic work.
Figure 11. Variation of heat transfer rate with crankshaft angle predicted for prototype
compressor at Test Condition 1.
740 HVAC&R RESEARCH
Comparison of Model and Experimental Results
The model was validated by testing a prototype compressor on a load stand at 17 different
operating conditions. A detailed description of the load stand can be found in the paper by Chen
et al. (2002b) that documents experimental testing of a scroll compressor. Table 4 compares the
model results at Test Condition 1 without intercooling to the experimental results. Because the
Figure 12. Comparison of first-stage leakage mass flow rates.
Figure 13. Variation of pressure with crankshaft angle predicted for prototype compressor
at Test Condition 1 with and without intercooling to maintain a 10C superheat.
VOLUME 14, NUMBER 5, SEPTEMBER 2008 741
Table 3. Summary of Model Results for Prototype Compressor at Test Condition 1
with and without Intercooling to Maintain a 10C Superheat
Without Intercooling With Intercooling
Suction pressure, kPa 1087 1087
Suction temperature, C 33.2 33.2
Intermediate pressure, kPa 1745 1463
Intermediate temperature, C 69.2 69.4
Discharge pressure, kPa 3453 3453
Discharge temperature, C 112.8 89.7
Power consumption, kW 2.08 1.89
Mass flow rate, kg/h 133.4 133.9
Isentropic efficiency
First stage 64.8% 59.4%
Second stage 69.3% 68.2%
Volumetric efficiency
First stage 95.4% 96.2%
Second stage 94.6% 94.9%
Figure 14. Variation of temperature with crankshaft angle predicted for prototype
compressor at Test Condition 1 with and without intercooling to maintain a 10C
superheat.
742 HVAC&R RESEARCH
power required for each stage of compression was not measured experimentally, the actual
power consumption term in the isentropic efficiency calculation (Equation 64) had to be esti-
mated as the product of mass flow rate and enthalpy change. This resulted in a slightly different
modeled isentropic efficiency compared to the values in Table 3. However, the tuned model pre-
dictions agree well with the experimental results.
Figures 15 and 16 compare the model predictions to the experimental data for intermediate
pressure and temperature, respectively. Error bars are used in the figures to indicate that the
experimental temperatures are measured with an accuracy of 2.2 K, while the pressure mea-
surements have an accuracy of 1%. While the intermediate temperatures show a greater dis-
crepancy between the modeled and experimental results than the intermediate pressures, the
majority of the results are within 10 K. The temperatures predicted by the model tend to be
lower than the measured temperatures. This could be related to the models tendency to underes-
timate the intermediate pressure. If the model were to correctly predict the density of refrigerant
entering the second stage to satisfy the mass balance but predict a low intermediate pressure, it
would also have to predict a low intermediate temperature to achieve this density. Comparing
the model results to internal measurements will help to identify the source of this error.
For the majority of the cases, the discharge temperatures are predicted within 10C of the mea-
sured values, as shown in Figure 17. The error in the discharge temperature would partially result
from the error in the intermediate temperature predicted by the model. However, uncertainty in the
calculation of heat loss by natural convection from the shell would also contribute to this error.
Figure 18 compares the predicted and measured power consumptions of the compressor. The
power meter used for the experimental measurements has an accuracy of 0.2%, so the error
bars for the experimental points are not visible in this case. All of the model results agree with
the experimental results within 5%, but the model seems to be most accurate for cases with
high power consumption. This could indicate that the constant combined mechanical and motor
Table 4. Summary of Model Results for Prototype Compressor at Test Condition 1
Compared to Experimental Results
Modeled Experimental
Uncertainty in
Experimental
Measurement
Error in Modeled
Result Compared to
Experimental Result
Suction pressure, kPa 1087 1087 1%
Suction temperature, C 33.2 33.2 2.2C
Intermediate pressure, kPa 1745 1780 1% 2.0%
Intermediate temperature, C 69.2 70.9 2.2C 1.7C
Discharge pressure, kPa 3453 3453 1%
Discharge temperature, C 112.8 111.2 2.2C 1.6C
Power consumption, kW 2.08 2.08 0.2% 0.02%
Mass flow rate, kg/h 133.4 133.2 0.7% 0.18%
Isentropic efficiency
First stage 51.4% 51.0% 6.7% 0.8%
Second stage 76.8% 84.2% 12.5% 9.6%
Volumetric efficiency
First stage 95.4% 95.0% 1.8% 0.4%
Second stage 94.6% 92.9% 1.4% 1.8%
VOLUME 14, NUMBER 5, SEPTEMBER 2008 743
efficiency of 78.3% that the model assumes is more accurate for the higher power consumption
cases. The motor may operate less efficiently in the cases corresponding to lower torques.
The mass flow rate through the compressor predicted by the model is also predicted within 5%
of the experimental values, as shown in Figure 19. The accuracy of the mass flow measurement
depends on the mass flow rate, but the accuracy is at least 0.9% for all of the cases shown. The
modeled flow rate is generally overpredicted at lower mass flow rates, which correspond to cases
Figure 15. Comparison of predicted and measured intermediate pressures.
Figure 16. Comparison of predicted and measured intermediate temperatures.
744 HVAC&R RESEARCH
with the largest ratio between discharge and suction pressure. This indicates that the leakage is
slightly underpredicted by the model for these cases.
Finally, Figure 20 shows that there is good agreement between the energy efficiency ratio
(EER) calculated from the model and experimental results. The EER was calculated assuming
that a vapor-compression cycle operates with 5 K of subcooling exiting the condenser. Employ-
ing an uncertainty analysis proposed by Bege et al. (2002), the accuracy of the experimental
EER is within 0.8%.
Figure 17. Comparison of predicted and measured discharge temperatures.
Figure 18. Comparison of predicted and measured power consumptions.
VOLUME 14, NUMBER 5, SEPTEMBER 2008 745
In summary, the model accurately predicts the mass flow rate through the compressor and its
power consumption. However, there is a wider range of error in the predicted intermediate and
discharge temperatures and in the intermediate pressure. This is to be expected as a result of the
tuning to one particular set of experimental results at one test condition.
CONCLUSION
The agreement between external measurements and model results suggests that the model
provides a good estimate of compressor performance (within 5% for mass flow rate and power).
Figure 19. Comparison of predicted and measured mass flow rates.
Figure 20. Comparison of EER for experimental and modeled results.
746 HVAC&R RESEARCH
The model not only predicts the pressure and temperature distributions in the suction and com-
pression chambers but also provides estimates of the mass flow rates through the various leak-
age paths and the heat transfer from the cylinder walls. Therefore, the effect of leakage and heat
transfer on the performance of the compressor can be investigated. However, the leakage analy-
sis is limited by the assumption of a constant flank leakage coefficient to account for the pres-
ence of lubrication. The assumption of a leakage coefficient also limits the models ability to
calculate exit temperatures, which were predicted within 10 K of the experimental measure-
ments. The incorporation of a lubrication analysis is an area for future work. However, the
model does provide a tool for comparing the performance of an existing compressor prototype
with and without intercooling. For a single case, intercooling resulted in an approximately 10%
improvement in compressor EER.
ACKNOWLEDGMENTS
The authors would like to acknowledge the support of Dr. Kwan-Shik Cho, vice president and
director of the Digital Appliance Company (DAC) Laboratory of LG Electronics. The authors
would also like to acknowledge the contributions of Seung-Jun Lee and Young-Ju Bae of the
DAC Research Laboratory Compressor Group.
NOMENCLATURE
= compression chamber surface
area, m
2
= suction chamber surface area, m
2
= vane thickness, m
= leakage path clearance, m
= constant volume specific heat,
J/kgK
= hydraulic diameter, m
= eccentricity of roller, m
= total energy in the control
volume, J
= specific enthalpy, J/kg
= convective heat transfer coeffi-
cient, W/m
2
K
= cylinder height, m
= change in specific enthalpy of the
refrigerant across the intercooler,
J/kg
= specific heat ratio (isentropic
compression calculations)
= thermal conductivity, W/mK
(heat transfer calculations)
= mass flow rate into the control
volume, kg/s
= mass flow rate out of the control
volume, kg/s
= pressure, Pa
= pressure of gas leaving shell, Pa
= high-side pressure, Pa
= first-stage muffler pressure, Pa
= low-side pressure, Pa
= shaft power, W
= ratio of low- to high-side pressure
= pressure of gas in accumulator, Pa
= Prandtl number
= heat transfer rate, W
= heat transfer rate from refrigerant
in the intercooler, W
= gas constant, J/kgK
= average radius of curvature of
chamber, m
= cylinder radius, m
= roller radius, m
= inner radius of roller, m
= vane tip radius, m
= Reynolds number
= rotational speed of crankshaft,
revolutions per second
= time, s
= temperature, C
= ambient temperature, C
= temperature of cylinder wall, C
= temperature of gas exiting the
shell, C
= temperature of gas in control
volume, C
= temperature of gas entering
second stage, C
= saturation temperature at speci-
fied pressure, C
= temperature of the gas in the
compressor shell, C
= temperature of gas entering first
stage, C
= degree of superheat at inlet to
second stage, C
A
c
A
s
b
c
c
v
D
h
e
E
cv
h
h
c
H
c
h
i nt
k
k
m

i n
m

out
P
P
disch e arg
P
h
P
i nt
P
l
P
overall
P
r
P
suction
Pr
Q

i nt
R
r
aver
R
c
R
r
R
r i ,
r
v
Re
rps
t
T
T
amb
T
c
T
disch e arg
T
g
T
int
T
sat
T
shell
T
suction
T
sh
VOLUME 14, NUMBER 5, SEPTEMBER 2008 747
= specific internal energy, J/kg
(energy balance calculations)
= average velocity of gas in control
volume, m/s (Reynolds number
calculations)
= total internal energy of the control
volume, J
= specific volume, m
3
/kg
= volume, m
3
= vane velocity, m/s
= power, W
= vane extension, m
Greek Symbols
= angle between vane slot and line
connecting vane tip center to
roller center (geometry calcula-
tions)
= thermal diffusivity, m
2
/s (heat
transfer calculations)
= distance between the cylinder
wall and roller, m
= mechanical efficiency of the
compressor
= motor efficiency
= isentropic efficiency
= volumetric efficiency
= crankshaft angle
= dynamic viscosity, Ns/m
2
= density, kg/m
3
= angle at which distance is
measured
= kinematic viscosity, m
2
/s
= rotational speed of crankshaft,
degrees/s
REFERENCES
Bege, G., J. Drnovek, I. Punik, and J. Bojkovski. 2002. Calculation and proper presentation of the mea-
surement uncertainty in testing. Proceedings of the 19th IEEE Instrumentation and Measurement
Technology Conference, Anchorage, AK, pp. 145760.
Chen, Y., N.P. Halm, E.A. Groll, and J.E. Braun. 2002a. Mathematical modeling of scroll compressors
Part I: Compression process modeling. International Journal of Refrigeration 25(6):73150.
Chen, Y., N.P. Halm, J.E. Braun, and E.A. Groll. 2002b. Mathematical modeling of scroll compressors
Part II: Overall scroll compressor modeling. International Journal of Refrigeration 25(6):75164.
Costa, C.M.F.N. 1986. Use of a simulation model for theoretical optimization analysis of a rolling-piston
type rotary compressor. Proceedings of the International Compressor Engineering Conference at Pur-
due University, West Lafayette, IN, pp. 82439.
Costa, C.M.F.N., R.T.S. Ferreira, and A.T. Prata. 1990. Considerations about the leakage through the min-
imal clearance in a rolling piston compressor. Proceedings of the International Compressor Engineer-
ing Conference at Purdue University, West Lafayette, IN, pp. 85362.
Ishii, N., N. Morita, M. Kurimoto, S. Yamamoto, S. Kiyoshi, and K. Sawai. 2000a. Calculations for compres-
sion efficiency caused by heat transfer in compact rotary compressors. Proceedings of the 15th Interna-
tional Compressor Engineering Conference at Purdue University, West Lafayette, IN, pp. 46774.
Ishii, N., M. Noriyuki, M. Ono, O. Aiba, K. Sano, and K. Sawai. 2000b. Net efficiency simulations of com-
pact rotary compressors for its optimal performance. Proceedings of the International Compressor
Engineering Conference at Purdue University, West Lafayette, IN, pp. 47582.
Jun, Y. 2002. Mechanical loss analysis of inverter controlled two cylinders type rotary compressor. Pro-
ceedings of the International Compressor Engineering Conference at Purdue University, West Lafay-
ette, IN, pp. C56.
Lee, J., and T.S. Min. 1988. Performance analysis of rolling piston type rotary compressor. Proceedings
of the International Compressor Engineering Conference at Purdue University, West Lafayette, IN,
pp. 15462.
Okada, K., and K. Kuyama. 1982. Motion of rolling piston in rotary compressor. Proceedings of the Inter-
national Compressor Engineering Conference at Purdue University, West Lafayette, IN, pp. 17884.
Padhy, S.K., and S.N. Dwivedi. 1994. Heat transfer analysis of a rolling-piston rotary compressor. Interna-
tional Journal of Refrigeration 17(6):40010.
u
u
U
cv
v
V
v
vane
W

mech

motor

748 HVAC&R RESEARCH


Shimizu, T., M. Kobayashi, and T. Yanagisawa. 1980. Volumetric efficiency and experimental errors of
rotary compressors. International Journal of Refrigeration 3(4):21925.
Yanagisawa, T., T. Shimizu, I. Chu, and K. Ishijima. 1982. Motion analysis of rolling piston in rotary com-
pressor. Proceedings of the International Compressor Engineering Conference at Purdue University,
West Lafayette, IN, pp. 18592.
Yanagisawa, T., and T. Shimizu. 1985a. Leakage losses with a rolling piston type rotary compressor:
Part IRadical clearance on the rolling piston. International Journal of Refrigeration 8(2):7584.
Yanagisawa, T., and T. Shimizu. 1985b. Leakage losses with a rolling piston type rotary compressor.
Part IILeakage losses through clearances on rolling piston faces. International Journal of Refrig-
eration 8(3):15258.
Yi, F., E.A. Groll, and J.E. Braun. 2004. Modeling and testing of an automobile AC scroll compressor: Part
IModel development. Proceedings of the International Compressor Engineering Conference at Pur-
due University, West Lafayette, IN, pp. C082.

Potrebbero piacerti anche