Analysis of Vehicle Exhaust Waste Heat Recovery Potential Using A Rankine Cycle

You might also like

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 56

Analysis of Vehicle Exhaust Waste Heat Recovery Potential Using a

Rankine Cycle

Antnio Domingues(1), Helder Santos(2), Mrio Costa(1)*

(1)

Mechanical Engineering Department, Instituto Superior Tcnico, Technical University

of Lisbon, Avenida Rovisco Pais, 1049-001 Lisbon, Portugal.

(2)

School of Technology and Management, Polytechnic Institute of Leiria, Morro do

Lena Alto Vieiro Apt. 4163, 2411-901 Leiria, Portugal.

*Corresponding author:
Mrio Costa
Mechanical Engineering Department
Instituto Superior Tcnico
Av. Rovisco Pais, 1049-001 Lisboa
Portugal
Tel.: +351218417372; E-mail: mcosta@ist.utl.pt

Submitted to Energy
Type of paper: Full length article
October, 2012

Abstract
This study evaluates the vehicle exhaust waste heat recovery (WHR) potential using a
Rankine cycle (RC). To this end, both a RC thermodynamic model and a heat exchanger
model have been developed. Both models use as input, experimental data obtained from
a vehicle tested on a chassis dynamometer. The thermodynamic analysis was performed
for water, R123 and R245fa and revealed the advantage of using water as the working
fluid in applications of thermal recovery from exhaust gases of vehicles equipped with a
spark-ignition engine. Moreover, the heat exchanger effectiveness for the organic
working fluids R123 and R245fa is higher than that for the water and, consequently,
they can also be considered appropriate for use in vehicle WHR applications through
RCs when the exhaust gas temperatures are relatively low. For an ideal heat exchanger,
the simulations revealed increases in the internal combustion engine thermal and vehicle
mechanical efficiencies of 1.4%-3.52% and 10.16%-15.95%, respectively, while for a
shell and tube heat exchanger, the simulations showed an increase of 0.85%-1.2% in the
thermal efficiency and an increase of 2.64%-6.96% in the mechanical efficiency for an
evaporating pressure of 2 MPa. The results confirm the advantages of using the thermal
energy contained in the vehicle exhaust gases through RCs. Furthermore, the present
analysis demonstrates that improved evaporator designs and appropriate expander
devices allowing for higher evaporating pressures are required to obtain the maximum
WHR potential from vehicle RC systems.

Keywords: waste heat recovery; Rankine cycle; working fluid; thermodynamic


efficiency; heat exchanger.

Nomenclature

A
A/F
b
cp
d
Dh

E
f
F
h
I
k
L
LHV
m

N
Nt
Nu
p
P
Pr

Q
Rd
Re
T
U
v
V

area [m2]
air/fuel ratio
long side of a rectangular cross section [m]
heat capacity [kJ kg-1 K-1]
diameter [m]
hydraulic diameter [m]
exergy flow rate [kW]
Darcy friction factor
imposed load [N]
specific enthalpy [kJ kg-1]; heat transfer coefficient [W m-2 K-1]
exergy destruction rate [kW]
thermal conductivity coefficient [W m-1 K-1]
evaporator tube length [m]
low heating value [MJ kg-1]
mass flow rate [kg s-1]
engine speed [rpm]
tubes number
Nusselt number
pressure [Pa]
vehicle effective power [kW]
Prandtl number
heat rate [kW]
fouling factors [m2 K W-1]
Reynolds number
temperature [K]
overall heat transfer coefficient [W m-2 K-1]
specific volume [m3 kg-1]
vehicle speed [km h-1]
power [kW]

Greek symbols

aspect ratio of rectangular ducts, ratio of a small to large side length


surface area density [m2 m-3]
distance between tubes [m]
pressure drop [Pa]
heat exchanger effectiveness
efficiency
dynamic viscosity [N s m-2]
density [kg m-3]

Subscripts
o

initial
3

1,2,3,4
amb
c
cond
crit
e
evap
exp
ext
f
g
h
i
in
m
out
p
pp
pump
s
t
w

Rankine cycle process


ambient
thermodynamic cycle
condensation
critical
effective
evaporating
expansion
external
working fluid
exhaust gases
hydraulic
internal
inlet
material
outlet
pump
pinch-point
pumping
isentropic
turbine
wall

Superscripts
m

viscosity ratio exponent

Abbreviations
BMEP
EGR
ETC
HCFC
HFC
ICE
MTC
NTU
ORC
PPTD
RC
TEG
TIGERS
TWC
WHR

break mean effective pressure


exhaust gas recirculation
electrical turbo-compounding
hydrochlorofluorocarbons
hydrofluorocarbons
internal combustion engine
mechanical turbo-compounding
number of transfer units
organic Rankine cycle
pinch-point temperature difference
Rankine cycle
thermo-electric generator
turbo-generator integrated gas energy recovery system
three way catalyst
waste heat recovery

1. Introduction
Internal combustion engines (ICEs) are the major source of motive power in the
world, and this is expected to continue for some decades. Greenhouse effects and
depleted petroleum supplies are crucial issues that the developed worlds economies are
facing. Because of this, governments in industrialized countries have introduced strict
regulations for ICE emissions and fuel economy standards. In the last two decades,
manufacturers have improved significantly ICE efficiencies by applying a number of
new technologies [1]. In recognition of the need to further reduce vehicle exhaust
pollutant emissions (CO, NOx, hydrocarbons and particulate matter) and, more recently,
also CO2 emissions, there has been a lot of interest in the development of cleaner and
more efficient vehicle powertrains [2].
In ICEs only about 1/3 of the fuel combustion energy is converted into useful
work to drive the vehicle and its accessory loads. The remainder is engine waste heat
dissipated by the engine exhaust system, coolant system, and convection as well as
radiation from the engine block [3]. Nearly 40% of the heat energy is wasted with the
engine exhaust gases [4]. If the waste heat of an ICE can be recovered, the engine
efficiency will be improved [3]. Furthermore, global warming will be decreased [5].
To increase the ICEs thermal efficiency and to reduce CO2 emissions, different
waste heat recovery (WHR) techniques were recently proposed [1-10]. Among the
existing WHR techniques, the most important are the electrical turbo-compounding
(ETC), the mechanical turbo-compounding (MTC), the turbo-generator integrated gas
energy recovery system (TIGERS), the thermo-electric generator (TEG) and the
Rankine cycle (RC) [6, 7].
A number of studies [e.g., 1-3, 5-6] demonstrate that the RC or the organic
Rankine cycle (ORC) have a high WHR potential in automotive applications. The RC is

based on the steam generation in a secondary circuit, which represents an indirect


method of WHR. This technique has advantages compared with the so-called direct
WHR techniques (e.g., ETC, MTC and TIGERS) that use a power turbine fitted to the
vehicle exhaust, which has a much higher impact on the engine pumping losses. In
addition, a RC allows for high waste energy utilization and it is cheaper than other
WHR techniques such as thermo-electric generators [5].
The choice of the working fluid to be used in the RC depends on a number of
factors, namely, thermodynamic, environmental, safety, process-related and economic
issues. In particular, when implementing such a system on a moving vehicle with live
occupants, the choice must consider worse case scenarios like leakages or crashes. On
that event the fluid must be harmless to the vehicle occupants. For vehicle applications,
the low flammability level is a major concern. Hence, alcohols and hydrocarbons are
arguably not the best candidates, in spite of their good thermodynamic efficiencies.
Instead, refrigerants, already used in air conditioning systems, are usually good
candidates. The refrigerants (e.g., R245fa) are widely used in ORC applications because
of their good heat transfer properties, excellent thermal stability and low viscosity. They
are generally safe (non-flammable) and compatible with most materials. Under typical
low temperature ambient conditions they do not freeze, which is a major concern with
water. However, the current generation of refrigerants called hydroflurocarbons (HFCs)
has a high global warming potential (100-year time horizon) 950-1070 for R245fa ,
which means that their use could be limited or banned in the near future.
Various working fluids have been proposed for RC WHR applications. Saleh et al.
[11] investigated 31 pure component working fluids for low-temperature ORCs. Lai et
al. [12] investigated several pure working fluids (alkanes, aromates and linear siloxanes)
for high-temperature ORCs. Yamamoto et al. [13] evaluated the optimum operating

conditions of an ORC comparing hydrochlorofluorocarbon (HCFC)-123 and water as


working fluids. Ringler et al. [14] examined two kinds of RCs for WHR from gasoline
ICEs using water and ethanol as working fluids. Chammas and Clodic [15] compared
the performance of water, R-245ca and isopentane as RC working fluids in hybrid
vehicles.
Since RCs generate additional power without requiring extra fuel, both the
specific fuel consumption and the pollutant emissions of the vehicle are reduced
[8, 9]. The performance analysis of various system configurations based on RCs is
promising. According with Boretti [6] improvements in fuel economy up to 17% may
be possible, which can serve as a reference basis for the assessment of the current
potential of this technology, whose major downfalls are the increase of weight, the
packaging complexity, the transient operation and the costs. Of course, the
implementation of a RC on a vehicle requires detailed investigations of all these issues.
To recover the exhaust waste heat, the RC needs to utilize a heat exchanger to
extract energy from the exhaust gases. A heat exchanger used in such an application has
to be able to provide an adequate surface area in order to achieve high exchange
efficiency, while using a small-size and lightweight arrangement. Furthermore, the
pressure drop has to be minimized to avoid excessive pumping losses that will have a
negative impact on the ICE efficiency [16]. Mavidrou et al. [16] examined the exhaust
gas heat exchanger design problem, focusing on the usage of different heat exchanger
configurations and different types of heat transfer surfaces.
This article presents an analysis of a combined RC (power cycle) and heat
exchanger for vehicle WHR applications, which, to the best of our knowledge, is
currently lacking in the literature. Furthermore, the present study provides a systematic
analysis for three working fluids (water, R123 and R245fa). Initially, this work

evaluates experimentally the exhaust thermal energy contained in a vehicle equipped


with a spark ignition ICE. Subsequently, the data obtained are used as input in two
developed models: a RC thermodynamic model, which includes both energy and exergy
analyses, and a heat exchanger model. Among other features, the RC thermodynamic
model allows assessing the RC efficiency and the net power as function of the
evaporating pressure of the three working fluids. The heat transfer model permits to
perform heat exchanger sizing calculations and to assess the different heat exchanger
efficiencies and pressure drop. Finally, both the thermodynamic RC and the heat
exchanger models were used together to evaluate the vehicle exhaust WHR potential
using different RCs.
The evaporator and the expander are the most critical components of a RC system.
The present study considers available components (evaporator and expander) that allow
building a short term RC prototype for WHR in vehicle applications. Based on the
measured values in the chassis dynamometer experiments and on the simulation results,
the potential power output of the proposed RC prototype is assessed and compared for
different vehicle operating conditions.

2. Experimental approach
This section describes the experimental approach followed to gather the data used
as input in the RC thermodynamic and heat exchanger models. It also quantifies the
exhaust waste heat as a function of the vehicle operating conditions.
Chassis dynamometer measurements were carried out on a vehicle equipped with
a 2.8 liter VR6 spark ignition engine in order to measure the exhaust gases mass flow
rate and temperature for several steady state operating conditions (i.e., after engine
warm-up). For each engine speed (2000, 3000 and 4000 rpm) tests were made for

various loads. Table 1 presents the vehicle test conditions considered in this work. In the
table, N is the imposed engine speed, F is the imposed load, BMEP is the break mean
effective pressure, V is the vehicle speed, Pe is the vehicle effective power (or brake
power) measured at the chassis dynamometer,

m
g

is the vehicle exhaust mass flow

rate, Tg,in is the exhaust gases temperature measured downstream of the three way
catalyst (TWC), and

available
Q

is the heat available in the exhaust gases. The test

conditions originated exhaust gases mass flow rates and temperatures ranging from 12.8
g/s to 59.7 g/s and 730.9 K to 1052.3 K.
All measurements were obtained after both the engine and the TWC reached their
steady states. To validate the steady states the coolant temperature was controlled to 95
C 10 C. In addition, controlled temperatures had to remain stable within 5 C for,
at least, 1 minute. In the present study, the recorded measurements were always the
average of the readings over a period of time of, at least, 2 minutes.
In order to evaluate the repeatability of the torque and engine speed
measurements, six tests were performed for each steady state operating condition. In the
torque measurements, the combined uncertainty ranged from 2.01% to 2.97% and
no relationship between the uncertainty and the torque magnitude was identified. In the
engine speed measurements, it was found that the combined uncertainty decreases
monotonically as the engine speed increases. Consequently, the maximum uncertainty
occurred at idle ( 2.24%) and the minimum uncertainty at 4000 rpm ( 1.41%). The
BMEP depends on the torque measurement. As a result, giving that the displacement
volume is known, the experimental uncertainties associated with the BMEP calculations
are those of the torque measurements.
In the present study the exhaust gases mass flow rate was calculated based on the
inlet air mass flow rate and stoichiometric air fuel ratio. The inlet air mass flow rate was

measured using the engine air mass flow rate sensor. Repeatability tests yielded
maximum uncertainties of 3.2%. The uncertainties associated with the temperature of
the exhaust gases were evaluated to be within 10 K of the mean value.
The exhaust gases properties have been calculated using the equations shown in
Table 2, which were derived using the software Refprop 9.0 [17]. The composition
(mass fractions) of the exhaust gases was assumed to be 20.4% CO 2, 7.8% H2O and
71.8% N2 (minor components have been neglected).
In the present study, the heat available in the exhaust gases was calculated through
the following equation:

g ,T amb
T

Qavailable=m
g c pg

where

m
g

(1
)

is the exhaust gases mass flow rate,

g ,
T

is the exhaust gases

temperature before the RC heat exchanger (i.e., the exhaust gases temperature after the
TWC) and T amb is the ambient temperature, taken equal to 25 C in the present study.
Table 1 includes the values of

available
Q

for all the operating conditions. It is seen that

the available exhaust waste heat changes significantly from low loads and speeds to
high loads and speeds. Furthermore, regardless of the operating condition, the available
waste heat at the vehicle exhaust is higher than the vehicle effective power.

3. RC thermodynamic model
The RC is a vapor power cycle used in numerous applications to generate
electrical power. Figure 1 shows a schematic of a simple RC. It is composed by four
main components: a pump, a heat exchanger, a turbine/generator and a condenser. The

10

ideal thermodynamic cycle includes the following processes: an isentropic compression


process in a pump (1-2), an isobaric heat transfer process in a heat exchanger (2-3), an
isentropic expansion process through a turbine (or other expansion machine) (3-4), and
an isobaric heat transfer process in a condenser (4-1).
The pump supplies the working fluid to the heat exchanger, where the working
fluid is heated and vaporized, removing heat from the exhaust gases. The working fluid
leaves the heat exchanger in saturated or superheated state. The high enthalpy vapor is
then expanded in the expander (usually a turbine), which is coupled to a generator that
deliveries the RC power output. After the expander, the working fluid enters the
condenser where it condensates.
The mathematical model of the simple RC uses the thermodynamic energy
conservations equations [18]. The model considers a steady state operation with
negligible kinetic and potential energy effects. Taken these considerations into account,
the pump power is given by:

(2

p= m
f ( h2h1 )
W

)
The heat absorbed from the exhaust gases by the working fluid in the heat exchanger is
given by:

=m
f ( h3h2 )
Q

(3)

The turbine power is calculated by:

t =m
f ( h3 h4 )
W

(4

11

)
and the heat rejected from the condenser is given by:

out = m
f ( h4 h1 )
Q

(5)

The RC efficiency can be defined as the net power produced referred to the heat
received at the heat exchanger as follows:

c =

t W
p
W

(6)

Exergy is a useful concept for evaluating the performance of various energy


systems. Exergy is the maximum amount of work that can be done by a process as it
approaches the thermodynamic equilibrium with its surroundings by a sequence of
reversible processes [5, 18, 19]. The exergy of a subsystem is a measure of its distance
from equilibrium and, thus, it can classify the energy quality of the subsystem. Exergy
destruction rate labels the loss of exergy during the process [5]. The exergy destruction
is due to irreversibilities occurring inside the system or the components of the system, a
control mass for the system or, as in this work, a control volume for each component
and it can be caused by internal or external factors [20]. As in the previous studies [e.g.,
20-22], in this work, the contributions of the internal and external irreversibilities are
not recognized separately, being calculated as a whole.
The exergy destruction rate for each process in the cycle (evaporation, expansion,
condensation and pumping) can be expressed as follows:

12

T g ,
T g ,out

m
f ( s3s 2) + m
g c pg ln ( ]
I evap =T amb

(7a)

I exp =T amb m
f ( s 4s3 )

(7b)

I cond+discharged =T amb m
f s1s 4

h1 h4
T amb

(7c)

I pump =T amb m
f ( s 2s1 )

(7d)

Note that at the condenser the exergy destruction rate expresses the sum of the
irreversibility of the condenser and the exergy discharged with the cooling air that flows
across the condenser.
Apart from the simple RC, there are other RC configurations that permit to
increase the recovered thermal energy. For example, the thermal efficiency of a RC can
be augmented by adding a preheater or a regenerator [7]. Mago et al. [21] presented an
analysis of regenerative ORCs using dry fluids, where they demonstrated that
regenerative ORCs not only have higher first and second law efficiencies than basic
ORCs, but they also have lower irreversibilities and lower heat requirements to produce
the same power. Recently, Wang et al. [22] studied the characteristics of a dual loop
ORC system which recovers the waste heat from the exhaust and the coolant of a spark
ignition ICE. The high temperature (HT) loop recovers the exhaust waste heat using
R245fa as the working fluid and the low temperature (LT) loop recovers the coolant
waste heat and the residual heat from the HT loop using R134a as the working fluid.
The results revealed that with this configuration the net power of the LT loop is higher
than that of the HT loop.

13

However, RCs with regeneration, preheating processes or dual loop systems


requires more piping (increased pressure losses) and more complex devices [7].
Furthermore, compactness is an essential issue in RCs for vehicle applications.
Therefore, the present study analyzes only the simple RC for different working fluids.
The working fluids can be classified into three categories according to the slope of
the saturation vapor curve in a T-s diagram. Dry fluids have positive slopes, wet fluids
have negative slopes, and isentropic fluids have infinitely large slopes. A disadvantage
of using water (or other wet fluid) as the working fluid is the need to superheat the
steam to prevent condensation during the expansion, a problem that results directly from
the water thermophysical properties. The occurrence of condensation is a problem
because it can lead to erosion of the turbine blades [7]. The choice of appropriate
working fluids can circumvent the superheating requirement, especially in applications
that require operation at low temperatures, or that have a heat source with relatively low
temperature. In these situations, an organic working fluid offers advantages since, in
many cases, superheating is unnecessary. These cycles are called ORCs. In the case of
HCFCs and HFCs the saturation vapor line is almost vertical. This implies that, without
initial superheat, end points lie in the dry vapor region [23].
Fluid selection is one of the most important contributors to the overall RC
performance. There are several general criteria that the working fluids should ideally
satisfy. Stability, non-fouling, non-corrosiveness, non-toxicity and non-flammability are
a few preferable physical and chemical characteristics [18]. Several organic fluids for
use in RCs have been proposed in the literature [11-15, 20-22, 24-27]. Among them, the
organic fluids R123 and R245fa appear to be the most promising ones for the operating
conditions used in this work, mainly due to its non-flammable behavior and
thermodynamic performance. Therefore, the organic fluids R123 and R245fa have been

14

selected for the present study. In addition, water has been also considered as working
fluid in this work because it is non-toxic and due to its abundant availability. Table 3
presents the main thermophysical properties of the working fluids studied in this work
(R123, R245fa and water).
Figure 2 shows a schematic of a typical RC waste heat recovery system from ICE
exhaust gases. Considering the operational conditions of a simple RC assembled on the
ICE exhaust system of a vehicle, the present model constraints are as follows:
pcond , and

(i)

evaporating pressure varying between condensation pressure,

(ii)
(iii)
(iv)
(v)
(vi)

critical pressure, pcrit ;


dry expansion for all fluids;
isentropic expander efficiency, t =0.7 [7];
isentropic pump efficiency, p=0.8 [7];
negligible pressure losses in the heat exchangers and pipes;
for both organic fluids (R123 and R245fa) the condensation temperature is
Tcond = 323 K, which corresponds to the condensation pressures given in Table
3; for water, the condensation temperature is Tcond = 373 K, which corresponds

(vii)

to a condensation pressure of 1 bar [28];


the superheating temperature is set as the minimum to guarantee a dry
expansion.

A sensitivity assessment was performed for the values of the expander and pump
isentropic efficiencies, water evaporating and condensation pressures and organic fluids
evaporating and condensation temperatures. Table 4 summarizes the results obtained. It
is seen that the expander efficiency affects significantly the RC net power output.
Moreover, it is interesting to note that the RC net power output is more sensitive to the
condensation conditions of the working fluids than to the evaporating conditions.
Figure 3 shows T-s process diagrams for water, R123 and R245fa. The solid line
in each figure represents the engine exhaust gases temperature for the vehicle operating
condition 9 (see Table 1). Note that

T g ,out

15

depends on the RC heat exchanger

characteristics. In the present study, we used a

T g ,out

= 200 C in all RC

thermodynamics calculations performed. It is seen that only in the case of water the
temperature difference between the exhaust gases and the working fluid at the beginning
of the evaporation process (the so-called pinch-point temperature difference PPTD)
may be problematic.

Figure 4 shows a schematic of the T- Q

diagram used for the pinch-point

analysis in the RC heat exchanger. The PPTD is the smallest temperature difference in
the RC heat exchanger, establishing the maximum allowable evaporation pressure and,
thus, limiting the RC efficiency. It is necessary to guarantee a minimum PPTD of 30 C
[8], which is particularly challenging at low loads and low engine speeds.
In the present study, a RC thermodynamic model, which includes both energy and
exergy analyses, was developed to evaluate the vehicle exhaust WHR potential using a
RC. The required thermodynamic and transport properties for the water and the organic
working fluids were calculated with the aid of the Refprop 9.0 [17]. Figure 5 shows the
algorithm for the RC thermodynamic model, which was implemented in Matlab. The
thermodynamic properties of the working fluids studied were calculated for each
evaporating pressure from the condensing pressure to the critical pressure of the organic
working fluids (3.64 MPa for R245fa and 3.66 MPa for R123). In the case of the water
the calculation was performed up to 4.4 MPa, which corresponds to a working fluid
temperature close to that of the exhaust gases at the heat exchanger outlet.

16

4. Heat exchanger model


The heat exchanger (evaporator) is an essential component in vehicle WHR
applications. The following characteristics are desirable in an heat exchanger for such
applications:
i)
ii)

high heat exchanger effectiveness;


low pressure drop trough the heat exchanger, which minimizes the negative

iii)

impact of the exhaust back pressure on the ICE;


compactness.
Improvements in the heat exchanger effectiveness can be obtained by increasing

the heat transfer area or the heat transfer coefficient from the exhaust gases side. In a
tubular heat exchanger, the heat transfer area is usually increased by increasing the
number of tubes and/or by using fins inside the tubes. It is well known that the exhaust
gases heat transfer coefficient is much smaller than the RC working fluid heat transfer
coefficient. As a result, finned surfaces are usually placed on the exhaust gases side to
increase the heat transfer area [16].
A literature survey [e.g., 16, 29] revealed that shell and tube heat exchangers are
the most appropriate to be used as evaporator in RCs for vehicle exhaust applications.
The heat exchanger model developed in the present study allows accessing the thermal
and hydraulic characteristics of various duct geometries (square, rectangular and
circular) as function of the number of tubes of the heat exchanger.
The present heat exchanger model is based on the effectiveness-NTU (-NTU)
method. The overall heat transfer coefficient U is calculated according to the following
equation [30]:

U=

1
d ext d ext Rd ,i d ext
d ext
1
+
+
ln
+ Rd ,ext +
di hg
di
2 km
di
hf

( )

17

(8
)

In the present study the tube was assumed to be made of Aluminum, with
k m =225 W m1 K1 [30]. For this type of applications, it is extremely difficult to find
accurate data for the fouling resistances (Rd) appearing in eq. (8), but typical values
have been assumed to partially account for the fouling effects; specifically, Rd,ext =
8.810-5 m2 K W-1 for all areas in contact with the working fluid and Rd,i = 1.810-4 m2 K
W-1 for all areas in contact with the exhaust gases [31]. The overall heat transfer
coefficient U is little affected by the fouling resistances [16]. The fouling resistance on
the gas side is higher than that on the fluid side. The particulate matter present in the
exhaust gases are the principal responsible for the fouling resistance on the gas side.
Note that the present study examines an evaporator placed after a TWC (see Figure 2)
where the particulate matter is rather low. Moreover, we found that considering values
of the fouling resistances ten times higher than those above, the RC net power output
decreases by 6.7%. If the values of the fouling resistances are not considered the RC net
power output increases by 1.1%.
The heat transfer coefficient h is calculated as follows:

h=k g

(9

Nu d
Dh

For Red < 2100, the Nusselt number is calculated through the Sieder and Tate
correlation [30]:

1
3

( )] ( / )

Dh
Nud=1.86 d Pr
L

(10
d <2100
)

18

For d >2100 , the Nusselt number is evaluated with the aid of the Gnielinski
correlation [30]:

( f /8 ) ( d 1000 ) ( Pr )

( ) ] ( / )

D
Nud=
1+ h
1/ 2
2/ 3
L
1+12.7 ( f /8 ) ( Pr 1 )

2
3

d >2100

(11)

In eqs. (10) and (11), is the fluid viscosity at the bulk fluid temperature, w is the fluid
viscosity at the heat transfer boundary surface temperature, m = 0.14 for
Re < 8000 and m = 0.25 for Re > 8000. In eq. (11), f is a logarithmic function of the
Reynolds number:

(12

f =( 0.79 ln ( d ) 1.64)2

The heat exchanger effectiveness, , is calculated from:

NTU

=1e

=1e

hg A
m g c pg

(13)

considering only the exhaust gases flow. Note that this relation is valid only for
condensers and evaporators.
The surface area density,

, for square, rectangular and circular cross flow

geometries is calculated through the following equations [32]:

4b
( b+ )2

(14a)

2 ( +1 ) b
( b + ) ( b+ )

(14b)

19

2 b
2
3 ( b+ )

(14c)

The pressure drop through shell and tube heat exchangers consisting of square,
rectangular and circular cross flow geometries is calculated as follows [32]:

A0)
4 fL ( m/
p=
Nt
2 g D h

(15)

The heat exchanger considered in the present study is a shell and tube counter
flow type, with the hot fluid (exhaust gases) in tubes and the cold fluid (RC working
fluid) in the shell. In order to study the effect of the important thermal and hydraulic
characteristics of the heat exchanger as a function of the number of the tubes for
different cross flow geometries, first we had to define the main dimensions of the heat
exchanger. Hussain and Brigham [29] demonstrated that the heat exchanger
effectiveness decreases for larger shell diameters, which depend on the cross flow area
of the heat exchanger. On the other hand, it is well known that the exhaust gases
pressure drop decreases for larger cross flow areas of the heat exchanger tubes.
Considering this trade-off and the limited space available for the installation of the heat
exchanger, the cross flow area of the heat exchanger tubes in the present study was set
equal to the vehicle exhaust duct cross flow area. The vehicle exhaust duct under study
has a cross flow area equal to
of each tube,

A0

A 0=2.56110 m

. To calculate the cross flow area

was divided by the number of tubes in the evaporator. In regard to

the heat exchanger length, it was assumed tube lengths of 0.5 m owing to the
dimensions of vehicle under study. In this case we have also performed a sensitivity
analysis by considering variations in the heat exchanger length of 0.1 m. The
20

calculations indicated variations in the RC net power output of 15%, revealing the
significant impact of this parameter in the present study.
Figure 6 presents the thermal and hydraulic characteristics (Re, Nu, h, , and p)
of the exhaust gases as a function of the number of tubes in the evaporator for different
cross flow geometries. This figure was constructed based on the exhaust gases mass
flow rate and temperature of operating condition 9. It is seen that both the Re and Nu
decrease when the number of tubes increases, see Figures 6a and 6b The tubes hydraulic
diameter also decreases as the number of tubes increases and, as a result, the heat
transfer coefficient increases, as seen in Figure 6c. Figure 6d reveals that the evaporator
effectiveness increases as the number of tubes increases. This is because both the heat
transfer coefficient and the surface area density increase as the number of tubes
increases, see Figures 6c and 6e.
Figure 6 also reveals that the pressure drop increases as the number of tubes
increases and that both the heat exchanger effectiveness and the pressure drop are
higher for the rectangular section, followed by the square section and by the circular
cross section. It is known that low pressure drops are associated with a low heat transfer
coefficients. In the present application, the heat exchanger is to be installed in the
exhaust duct of the vehicle. In such an application, the minimization of the exhaust
gases back pressure on the ICE is of critical importance. The heat exchanger with
circular tubes provides the lowest exhaust gases back pressure (see Figure 6f).
Furthermore, this is the simplest geometry for construction, which allows minimizing
the heat exchanger costs. As a result, in spite of the heat exchanger effectiveness
penalty, the configuration with circular tubes was selected for this study.
As already pointed out, Figure 6f indicates that the exhaust gases pressure drop
increases markedly with the number of tubes, while Figure 6d shows that the increase in

21

the heat exchanger effectiveness for a large number of tubes (40 to 100) is
comparatively small. These results establish the maximum for the number of tubes in
the heat exchanger: 43 tubes in the present study. Considering the cross flow area and
the hexagonal shell and tube heat exchanger arrangement, 43 tubes corresponds to a
tube diameter of 10 mm, with a distance between the tubes of 4 mm, and a shell inside
radius of 94 mm.
The heat exchanger was divided into three zones for modeling purposes, as shown
in Figure 7, namely, a preheating zone, an evaporating zone and a superheating zone.
These zones were considered as individual heat exchangers with the appropriate
boundary conditions for temperature and mass flow rate. The amounts of exchanged
heat for preheating, evaporating and superheating of the RC working fluid were
estimated from the heat transfer relations and the -NTU method [29, 31]. The -NTU
method was applied to the three zones of the heat exchanger. The enthalpy ratio was
used as a first approximation for each zone.
Figure 8 shows the algorithm for the heat exchanger model, which was also
implemented in Matlab. As input data, this model uses the results of the RC
thermodynamic model, discussed earlier. The required thermodynamic and transport
properties for the exhaust gases were calculated with the aid of the Refprop 9.0 [17]. A
parametric iterative method was employed for selecting the optimum working fluid and
to obtain the working fluid mass flow rate for the investigated heat exchangers. The
calculations were performed for different evaporating pressures and working fluids until
reaching convergence between the working fluid mass flows in the various subcomponents of the heat exchanger.

5. Results and discussion

22

5.1. RC thermodynamic analysis


This section presents a combined first and second law analysis. The main
objective of this section is to assess the RC efficiency,
expansion ratio, v4/v3, the working fluid mass flow rate,

, the turbine outlet/inlet

m
f , and RC net power,

net , as function of the evaporating pressure for the working fluids studied. To this
W
end, an adiabatic heat exchanger was considered so that the maximum vehicle exhaust
WHR potential, assuming a

T g ,out

= 200 C (see section 3) was assessed. The

thermodynamic analysis reported below has been performed for the exhaust conditions
corresponding to the vehicle operating conditions 3, 9 and 13 (see Table 1), which
represent low, intermediate and high engine speed and load steady state vehicle
operating conditions, respectively. This allows the assessment of the performance of the
RC for different exhaust mass flow rates and temperatures.
Figure 9 shows the RC efficiency as a function of the evaporating pressure for the
working fluids studied. The figure reveals that the RC efficiency (first law efficiency) is
very sensitive to the evaporating pressure (or evaporating temperature) and working
fluid. In regard to the working fluid, the RC efficiency at the evaporating pressure of,
for example, 2 MPa is 14.29% for water, 12.03% for R123 and 9.53% for R245fa.
Figure 10 shows the turbine outlet/inlet expansion ratio (v4/v3) as a function of the
evaporating pressure for the working fluids studied. The working fluid R245fa presents
the lower turbine outlet/inlet expansion ratio (v4/v3), regardless of the evaporating
pressure. This is mainly due to the higher R245fa condenser pressure (4.012 bar) as
compared to that of the water (1 bar). The turbine outlet/inlet expansion ratio (v4/v3) is
an important parameter as it indicates how much the fluid volume increases through the
expansion process. It should be noted that the expansion ratio (v4/v3) can change
significantly according with the characteristics of the working fluid. The expansion ratio

23

is also very important for the expander selection. When the expansion ratio (v4/v3) is
smaller than 50, expansion efficiencies higher than 0.8 can be achieved using a single
stage axial turbine as expander [7].
Figures 11 and 12 show the working fluid mass flow rate and the RC net power
output as a function of the evaporating pressure for the working fluids studied for
operating conditions 3, 9 and 13. For a given operating condition, Figure 11 reveals that
the working fluid R123 requires the highest mass flow rate, being the mass flow rate
required by the working fluid R245fa slightly lower. It can also be seen in Figure 11
that, for a given operating condition, the water mass flow rate is one order of magnitude
lower than the organic fluids mass flow rate as a consequence of the highest enthalpy
increase in the evaporation process. This indicates that the pump power required in the
ORC is much higher than the pump power required in the water RC. Thus the pump
should be carefully selected to minimize the power costs under high evaporating
pressures. Furthermore, the selected pump needs to be equipped with a frequency
converter to control its rotation speed and, hence, the evaporating pressure and the mass
flow rate of the RC. For a given operating condition, Figure 12 reveals that the RC net
power output is higher for water and lower for R245fa, regardless of the evaporating
pressure.
Figure 11 shows that the working fluid mass flow rate increases for the engine
operating condition 13 as compared to the operating conditions 9 and 3 due to the
increase in the waste heat flow. Figure 12 reveals that when the waste heat flow
increases (from the operating condition 3 to the operating condition 13), the RC net
power output increases significantly.
It is important to point out that the condensation pressures (and temperatures) for
the organic fluids and water are different. As mentioned in the section 3, we have

24

chosen a condensation temperature of 323 K for the organic fluids (R123 and R245fa)
and 373 K for water. Despite the higher water condensation temperature, Figure 12
reveals that water yields higher RC net power output than the organic fluids. The higher
water condensation temperature also allows the use of smaller size condensers as
compared to the organic fluids. Note that when the temperature difference between the
condenser and the ambient conditions decreases the condenser size has to be increased.
This is a critical issue when using organic fluids in vehicle exhaust WHR applications
that needs to be carefully considered in future ORC studies.
The higher temperature difference between the exhaust gases and the working
fluid in the heat exchanger for R123 and R245fa (see Figure 3) increases the evaporator
irreversibilities that are the main cause for the lower RC efficiencies for the organic
working fluids as compared to water.
As pointed out earlier, Mago et al. [21] presented an analysis of the performance
of different ORC configurations using dry organic fluids (R113, R123 and R245ca).
These authors showed that superheating organic fluids increase the irreversibilities and
decrease the second law efficiency. Note that in the present study the evaporating
pressure of the organic fluids varied between the condensation pressure and the critical
pressure (see Table 3). In order to reduce the evaporator irreversibilities at the
evaporator output the organic fluids operated at saturated conditions (see Figure 3);
closer to the critical pressure the superheating temperature was set as the minimum to
guarantee a dry expansion.
Figure 13 shows the RC evaporator exergy destruction rate as a function of the
evaporating pressure for the working fluids studied for operating condition 9. High
evaporating pressures reduce the temperature difference in the evaporator, so the exergy
destruction rate decreases. Increasing the evaporating pressure is a good method to

25

avoid second law losses. High evaporating pressures also increase the RC first law
efficiency, as shown in Figure 9. As a result, increasing the evaporating pressure is a
better method to increase the RC performance than increasing the superheating
temperature.
Previous studies [e.g., 2] demonstrate that binary mixture working fluids can
diminish significantly the evaporator entropy generation (exergy destruction) because
these fluids can reduce the temperature difference in the evaporator. Research on binary
mixture working fluids is still very limited and more work is necessary to obtain a better
understanding of their influence on the performance of RCs.
The present results (not shown here) revealed also that the entropy generation rate
in the condenser is significantly lower than that in the evaporator mainly because of the
smaller temperature differences that occur in the condenser. The entropy generation rate
in the expander and pump are related to the isentropic efficiency of these devices. Our
results indicated that the expander entropy generation rate is much higher than that in
the pump, but considerably lower than those in the evaporator and condenser.
Consistently with previous studies [e.g., 20], the present analysis demonstrates that the
evaporator makes the biggest contribution to the overall entropy generation rate in the
RC system.
To complement the present second law analysis, the evaporator exergy efficiency
was also determined. To this end, the dead state was specified as

T amb =25C

and

pamb=1 atm .
The exergy flow rate of the exhaust gases (assumed to be an ideal gas) entering
and leaving the RC can be calculated as:

26

( ( ) ( )) ]

E j=m g c pg ( T g , jT amb) c pg ln

T g, j
pj
R g ln
T amb
pamb

(16)

The exergy flow rate of the working fluid at any point of the RC can be
determined as:

E i= m f [ ( hihamb )T amb ( s is amb ) ]

(17)

The evaporator exergy efficiency can be expressed as:

E g , E
E
E
E
exergy ,evap = useful = 3 2
Eavailable
g ,out

where

useful
E

and

available
E

(18)

are the actual exergy used and the theoretically

available exergy at the evaporator,

2
E

and

3
E

are the exergy flow rates of the

working fluid entering and leaving the evaporator, respectively, and


g ,out
E

g ,
E

and

are the exergy flow rates of the exhaust gases entering and leaving the

evaporator, respectively.
Figure 14 depicts the variation of the RC evaporator exergy efficiency as a
function of the evaporating pressure for the working fluids studied for operating
condition 9. The results presented in this figure agree well with the results presented in
Figure 13, since a decrease in the evaporator irreversibilities represent an increment in
the evaporator exergy efficiency. Figure 14 also reveals that the evaporator exergy
efficiency is rather modest regardless of the working fluid. Nevertheless it is important
to observe that the evaporator exergy efficiency is about two times higher for water as

27

compared to the organic fluids. The results from the combined first and second law
analysis show that water yields higher thermodynamic efficiency as compared to the
organic fluids.
The thermodynamic analysis demonstrates that the water is an adequate working
fluid for use in an exhaust heat recovery system with a RC for the following reasons: as
compared to the organic fluids, water (i) yields higher thermodynamic efficiency and
net power output; (ii) requires lower quantities of working fluid in the RC (less weight);
(iii) condenses easily at atmospheric pressure; (iv) has lower price and higher
abundance; and (v) presents no environmental risks. It should be stressed, however, that
the use of water as the working fluid in cold regions may be problematic. In case of
frost, water expands at freezing and this can destroy the equipment in a single cold
night.

5.2. Heat exchanger analysis


This section extends the RC thermodynamic analysis presented in the previous
section, by considering also the heat exchanger model. It examines two shell and tube
counter flow type heat exchangers with the exhaust gases passing through the tubes,
referred hereafter as evaporator 1 and evaporator 2. Evaporator 1 maintains the cross
flow area of the vehicle exhaust duct under study with tube diameters of 1 cm. Specific
heat exchangers for WHR applications are currently not available, so a shell and tube
counter flow type exhaust gas recirculation (EGR) cooler from a MAN diesel truck was
selected as evaporator 2. A similar approach was considered in previous studies [e.g.,
27]. Table 5 summarizes the main characteristics of evaporators 1 and 2. The main
objective of this section is to assess the influence of the heat exchanger on the RC net

28

power output,

net , as a function of the evaporating pressure for the working fluids


W

studied.
Figure 15 shows the working fluid mass flow rate as a function of the evaporating
pressure for the working fluids studied using evaporator 1 for the operating conditions
3, 9 and 13. It is seen that the working fluid mass flow rate decreases (approximately
50%) compared to the working fluid mass flow rate obtained for the optimal heat
transfer conditions, as presented in Figure 11, regardless of the vehicle operating
condition. This is mainly due to the high thermal resistance of evaporator 1. Figure 15
reveals an increase in the organic fluids mass flow rate for evaporating pressures higher
than 2.25 MPa for the operating condition 13, which is due to the flow regime transition
from laminar to turbulent in the heat exchanger. It should be stressed that this occurs
only for operating condition 13, which represents a high engine speed and load steady
state vehicle operating condition.
Figure 16 shows the heat exchanger effectiveness as a function of the evaporating
pressure for the working fluids studied using evaporator 1 for the operating conditions
3, 9 and 13. This figure reveals that the heat exchanger effectiveness for water is lower
than those for the organic working fluids R123 and R245fa, regardless of the vehicle
operating condition. Furthermore, it is seen that the heat exchanger effectiveness for
water decreases monotonically as the evaporating pressure increases. Figure 16 also
reveals that the heat exchanger effectiveness for water is similar for both operating
conditions 9 and 13, but decreases significantly for the operating condition 3. As for the
organic fluids, Figure 16 reveals yet that the heat exchanger effectiveness decreases
from operating condition 3 to operating condition 13. This is because the organic fluids
allow using lower temperature heat sources and yield higher heat exchanger
effectivenesss when the temperatures of the exhaust gases are relatively low.

29

It should be noted that when evaporator 1 is considered in the simulations, the


temperature of the exhaust gases decreases only by about 200 C. In this case the value
of T g ,out

was well above that (200 C) considered in the RC thermodynamic analysis

in the previous section.


Figure 17 shows the RC net power output as a function of the evaporating
pressure for the working fluids studied using evaporator 1 for the operating conditions
3, 9 and 13. For the operating conditions 3 and 9, the figure reveals that the RC net
power output is higher for the two organic fluids than for water for evaporating
pressures > 1 MPa. This is due to the lower heat exchanger effectiveness for water. A
comparison of these results with those in Figure 12 reveals that the tubular heat
exchanger causes a significantly decrease in the net power extracted from the RC using
water as the working fluid. These results demonstrate that the low net power produced
by the RC when using water as the working fluid is mainly due to the low heat
exchanger effectiveness (less than 40%). As the vehicle load and speed increase
(operating condition 13) the heat exchanger effectiveness tends to be similar for the
three working fluids considered, see Figure 16. Consistently, for the operating condition
13, Figure 17 reveals that for evaporating pressures < 2.25 MPa the RC net power
output is (slightly) higher for water than for the organic fluids. In contrast, for
evaporating pressures > 2.25 MPa the RC net power output is higher for the organic
fluids than for water owing to the increase in the organic fluids mass flow rates, as seen
in Figure 15.
Table 6 shows the heat exchanger characteristics for evaporators 1 and 2 for the
operating condition 13 (Table 1). It is seen that the overall heat transfer coefficient is
higher for evaporator 2 than for evaporator 1, mainly due to the reduced cross flow area
of the heat exchanger (2.5 times lower as compared to the normal cross flow area of the

30

existing exhaust duct in the vehicle), which increases the flow velocity of the exhaust
gases.
In regard to the heat transfer characteristics of the working fluids, Table 6
demonstrates that the overall heat transfer coefficient is higher for water followed by the
organic fluids R245fa and R123. Note that Table 6 examines the operating condition 13,
which corresponds to a high vehicle load and high engine speed (see Table 1). This
means that the poor water evaporator effectiveness observed for the operating condition
3 (see Figure 16) is reversed for the operating condition 13. Hence, it can be concluded
that water is the best working fluid for use in vehicle WHR applications through RCs
when the exhaust gases temperatures are relatively high.

5.3. Influence of the RC on the ICE thermal and vehicle mechanical efficiencies
This section evaluates the influence of the RC on the ICE thermal and vehicle
mechanical efficiencies. The main objective of this section is to perform a systematic
assessment of the RC vehicle exhaust WHR potential.
The results from the RC thermodynamic and heat exchanger models were used to
calculate improvements in efficiencies. The ICE thermal efficiency was calculated
through the following equation:

th =

net
net
W
W
=
m
fuel LHV m
air
LHV
A
F

(19)

where LHV was taken equal to 44 MJ/kg [33].


The vehicle mechanical efficiency was defined as the ratio of the useful work
net , and the effective power produced by the ICE, Pe:
produced by the RC, W

31

m =

net
W
Pe

(20)

The calculations were performed for three operating conditions; specifically,


operating conditions 3, 9 and 13 in Table 1. In addition, three distinct RCs were
evaluated:
i) RC1 that considers an adiabatic heat exchanger and an evaporating pressure
of 2 MPa;
ii) RC2 that considers evaporator 1 (see Table 5) and an evaporating pressure of
2 MPa;
iii) RC3 that considers evaporator 2 (see Table 5) and an evaporating pressure of
0.52 MPa. In this case a commercial available expander (Green Turbine TM)
that works only with water as a working fluid is considered.
Note that the RC3 case evaluates the performance of a RC prototype system that
uses existing components: evaporator 2 and the Green Turbine TM, which is currently, to
the best of our knowledge, the most appropriate expander for a RC vehicle application.
The main characteristics of the Green TurbineTM are inlet steam maximum pressure of
5.2 bar, inlet steam maximum temperature of 200 C, power of 2.5 kW, weight of 7 kg,
length of 25 cm and diameter of 19 cm [34]. The reduced mass and dimensions makes
this turbine suitable for vehicle exhaust WHR applications.
The control of the RC in a vehicle application is particularly complex due to the
(often) transient regime of the heat source. As a result, optimizing the RC control is
crucial to maximize the performance of the system. For instance, to maintain the
evaporator pressure at 5 bar and to guarantee an inlet turbine temperature constant at
200 C it is needed to control the water flow rate through the evaporator by varying the
pump speed.

32

Table 7 presents the increase in the ICE thermal and vehicle mechanical
efficiencies for the three RCs studied. Water is the working fluid with the greatest
potential to be used in WHR from exhaust gases in vehicles through RC1 as compared
to the organic working fluids R123 and R245fa. Specifically, at an evaporating pressure
of 2 MPa, for operating condition 9, the increase in the vehicle mechanical efficiency
using the RC1 is 15.95%, 13.43% and 10.64% for water, R123 and R245fa,
respectively.
In the case of RC2, Table 7 shows that the organic fluid R123 is more suitable to
be used in WHR from exhaust gases in vehicles through RCs, especially at lower
exhaust gases temperatures. However, for higher temperatures and higher exhaust gas
flows, RC2 presents higher thermal and mechanical efficiencies if water is used as the
working fluid.
Table 7 reveals that the RC3 presents thermal and mechanical efficiencies of
0.3%-0.85% and 2.17%-3.87%, respectively. As compared with the RC1 (see Table 7)
these values are rather modest. However, note that the RC3 accounts for the practical
constrains introduced by the existing RC components.
The RC3 allows assessing the thermal and mechanical efficiencies of a short term
RC prototype. The results found demonstrate that future RC prototypes require
improved evaporator designs and expander devices allowing for higher evaporating
pressures.
The operation of the evaporator at higher pressure allows increasing the RC first
and second law efficiencies and permits reducing the temperature differences across the
heat exchanger surfaces, which minimize the film boiling effect in which rates of heat
transfer fall sharply [35].

33

The open literature [e.g., 7-9, 15] generally analysis the maximum WHR potential
of a RC fitted to the vehicle exhaust. In the present study the maximum WHR potential
was evaluated through the RC1. The ICE thermal efficiencies obtained in this study for
the RC1 are in agreement with those presented by Yamada and Mohamad [9], who
reported increases in thermal efficiencies of 2.9%-3.7%, as compared with 1.4%-3.52%
in this study (see Table 7). In regard to the increase in vehicle mechanical efficiency
(fuel economy), Vaja and Gambarotta [7] reported values around 12%, as compared
with 10.16%-15.95% in this study (see Table 7). Srinivasan et al. [8] examined the
exhaust waste heat recovery potential of a high efficiency, low emissions, dual-fuel, low
temperature combustion engine using an ORC. Their results showed that vehicle
mechanical efficiency improved by an average of 7% for all injection timings and loads
with hot EGR. Finally, Chammas and Clodic [15] reported vehicle mechanical
efficiency of 24%, but for higher evaporating pressures, with the thermal energy being
recovered also from the vehicle cooling system.

6. Conclusions
The vehicle exhaust WHR potential using a RC has been evaluated with the aid of
both a RC thermodynamic model and a heat exchanger model. Both models used as
input, experimental data obtained in a vehicle tested on a chassis dynamometer. The
thermodynamic analysis was performed for water, R123 and R245fa and revealed the
advantage of using water as the working fluid in applications of thermal recovery from
exhaust gases of vehicles equipped with a spark-ignition engine. Moreover, the heat
exchanger effectiveness for the organic working fluids R123 and R245fa is higher than
that for the water and, consequently, they can also be considered appropriate for use in

34

vehicle WHR applications through RCs when the exhaust gas temperatures are
relatively low.
For an ideal heat exchanger the simulations revealed increases in ICE thermal
efficiency and vehicle mechanical efficiency of 1.4%-3.52% and 10.16%-15.95%,
respectively, while for a shell and tube heat exchanger, the simulations showed an
increase of 0.85%-1.2% in the thermal efficiency and an increase of 2.64%-6.96% in the
mechanical efficiency for an evaporating pressure of 2 MPa. However, it is important to
note that the thermal and mechanical efficiencies can be enhanced with the increase in
the evaporating pressure of the working fluid.
The present analysis confirms that RCs have high potential for vehicle exhaust
waste heat recovery. However, improved evaporator designs and appropriate expander
devices allowing for higher evaporating pressures are required to obtain the maximum
WHR potential from vehicle RC systems. Considering increasing fuel prices and
environmental issues, this technology will permit to achieve further reductions in engine
specific fuel consumption and CO2 specific emissions.

35

7. References
[1] Katsanos CO, Hountalas DT, Pariotis EG. Thermodynamics analysis of a Rankine
cycle applied on a diesel truck engine using steam and organic medium. Energy
Conversion and Management 2012;60;68-76.
[2] Wang T, Zhang Y, Peng Z, Shu G. A review of researches on thermal exhaust heat
recovery with Rankine cycle. Renewable and Sustainable Energy Reviews
2011;15:2862-71.
[3] He M, Zhang X, Zeng K, Gao K. A combined thermodynamic cycle used for waste
heat recovery of internal combustion engine. Energy 2011;36:6821-9.
[4] Yu C, Chau KT. Thermoelectric automotive waste heat energy recovery using
maximum

power

point

tracking.

Energy

Conversion

and

Management

2009;50:1506-12.
[5] Wang EH, Zhang HG, Fan BY, Ouyang MG, Zhao Y, Mu QH. Study of working
fluid selection of organic Rankine cycle (ORC) for engine waste heat recovery.
Energy 2011;36:3406-18.
[6] Boretti A. Stoichiometric H2ICE with water injection and exhaust and coolant heat
recovery through organic Rankine cycles. International Journal of Hydrogen
Energy 2011;36:12591-600.
[7] Vaja I, Gambarotta A. Internal combustion engine (ICE) bottoming with organic
Rankine cycles (ORCs). Energy 2010;35:1084-93.
[8] Srinivasan KK, Mago PJ, Krishnan SR. Analysis of exhaust waste heat recovery
from a dual fuel low temperature combustion engine using an organic Rankine
cycle. Energy 2010;35:2387-99.

36

[9] Yamada N, Mohamad MNA. Efficiency of hydrogen internal combustion engine


combined with open steam Rankine cycle recovering water and waste heat.
International Journal of Hydrogen Energy 2010;35:1430-42.
[10] Miller EW, Hendricks TJ, Peterson RB. Modeling energy recovery using
thermoelectric conversion integrated with an organic Rankine bottoming cycle.
Journal of Electronic Materials 2009;38:1206-13.
[11] Saleh B, Koglbauer G, Wendland M, Fischer J. Working fluids for low-temperature
organic Rankine cycles. Energy 2007;32:1210-21.
[12] Lai N, Wendland M, Fischer J. Working fluids for high-temperature organic
Rankine cycles. Energy 2011;36:199-211.
[13] Yamamoto T, Furuhata T, Arai N, Mori K. Design and testing of the organic
Rankine cycle. Energy 2001;26:239-51.
[14] Ringler J, Serfert M, Guyotot V, Hubner W. Rankine cycle for waste heat recovery
of IC engines. SAE paper 2009-01-0174; 2009.
[15] Chammas RE, Clodic D. Combined cycle for hybrid vehicles. SAE paper 2005-011171; 2005.
[16] Mavridou S, Mavropoulos GC, Bouris D, Hountalas DT, Bergeles G. Comparative
design study of a diesel exhaust gas heat exchanger for truck applications with
conventional and state of the art heat transfer enhancements. Applied Thermal
Engineering 2010;30:935-47.
[17] Lemmon EW, McLinden MO, Huber ML. NIST reference fluid thermodynamic
and transport properties - REFPROP version 9.0. Boulder: National Institute of
Standard Technology; 2010.
[18] Moran MJ, Shappiro HN. Fundamentals of engineering thermodynamics. New
York: John Wiley and Sons; 2005.

37

[19] Cengel YA, Boles MA. Thermodynamics an engineering approach. 6th ed.
London: McGraw-Hill; 2008.
[20] Wei D, Lu X, LU Z, Gu J. Performance analysis and optimization of organic
Rankine cycle (ORC) for waste heat recovery. Energy Conversion and Management
2007;48:1113-1119.
[21] Mago PJ, Chamra LM, Srinivasan K, Somayaji C. An examination of regenerative
organic Rankine cycles using dry fluids. Applied Thermal Engineering
2008;28:998-1007.
[22] Wang EH, Zhang HG, Zhao Y, Fan BY, Wu YT, Mu QH. Performance analysis of a
novel system combining a dual loop organic Rankine cycle (ORC) with a gasoline
engine. Energy 2012;43:385-395.
[23] Hettiarachchi HDM, Golubovic M, Worek WM, Ikegami Y. Optimum design
criteria for an organic Rankine cycle using low-temperature geothermal heat
sources. Energy 2007;32:1698-706.
[24] Maizza V, Maizza A. Unconventional working fluids in organic Rankine-cycles for
waste energy recovery systems. Applied Thermal Engineering 2001;21:381-90.
[25] Liu BT, Chien KH, Wang CC. Effect of working fluids on organic Rankine cycle
for waste heat recovery. Energy 2004;29:1207-17.
[26] Drescher U, Brggemann D. Fluid selection for the Organic Rankine Cycle (ORC)
in biomass power and heat plants. Applied Thermal Engineering, 2007;27:223-8.
[27] Hung TC, Wang SK, Kuo CH, Pei BS, Tsai KF. A study of organic working fluids
on system efficiency of an ORC using low-grade energy sources. Energy
2010;35:1403-11.

38

[28] Liming F, Wenzhi G, Hao Q, Bixian X. Heat recovery from internal combustion
engine with Rankine cycle; 2010. Proceedings of Power and Energy Engineering
Conference, pp. 1-4, March 28-31, 2010, Chengdu, China.
[29] Hussain Q, Brigham D. Organic Rankine cycle for light duty passenger vehicles.
Directions in Engine-Efficiency and Emissions Research (DEER) Conference;
2011.

Presentation

avalaible

at

http://www1.eere.energy.gov/vehiclesandfuels/pdfs/deer_2011/wednesday/presentat
ions/deer11_hussain.pdf, October 3-6, 2011, Detroit, Michigan, USA.
[30] Incropera FP, DeWitt DP, Bergman TL, Lavine AS. Fundamentals of heat and mass
transfer. New York: John Wiley & Sons; 2007.
[31] Kakac S and Liu H, Heat exchangers, selection, rating and thermal Design. New
York: CRC Press.; 2002.
[32] Shah RK, Dusan PS. Fundamentals of heat exchanger design. New York: John
Wiley and Sons; 2003.
[33] Graboski MS, McCormick RL. Combustion of fat and vegetable oil derived fuels
in diesel engines. Progress in Energy and Combustion Science 1998;24:125-164.
[34] Green TurbineTM. Available from: www.greenturbine.eu [accessed on 31/03/ 2012].
[35] Stobart RK. An availability approach to thermal energy recovery in vehicles.
Proceedings of the Institution of the Mechanical Engineers Part D 2007;221:11071124.

39

Figure captions
Figure 1. Schematic of a simple RC.
Figure 2. Schematic of a typical RC waste heat recovery system from ICE exhaust
gases.
Figure 3. T-s process diagrams. a) Water, b) R123 and c) R245fa.

Figure 4. Schematic of the T- Q

diagram used for the pinch-point analysis in the RC

heat exchanger.
Figure 5. Algorithm for the RC thermodynamic model.
Figure 6. Thermal and hydraulic characteristics of the exhaust gases as a function of the
number of tubes in the evaporator for different cross flow geometries. a) Re,
b) Nu, c) h, d) , e) , f) p.
Figure 7. The three zones of the heat exchanger considered for modeling purposes.
Figure 8. Algorithm for the heat exchanger model.
Figure 9. RC efficiency as a function of the evaporating pressure for the working fluids
studied.
Figure 10. Turbine outlet/inlet expansion ratio (v4/v3) as a function of the evaporating
pressure for the working fluids studied.
Figure 11. Working fluid mass flow rate as a function of the evaporating pressure for
the working fluids studied.
Figure 12. RC net power output as a function of the evaporating pressure for the
working fluids studied.
Figure 13. RC evaporator exergy destruction rate as a function of the evaporating
pressure for the working fluids studied.
Figure 14. RC evaporator exergy efficiency as a function of the evaporating pressure
for the working fluids studied.

40

Figure 15. Working fluid mass flow rate as a function of the evaporating pressure for
the working fluids studied using evaporator 1.
Figure 16. Heat exchanger effectiveness as a function of the evaporating pressure for
the working fluids studied using evaporator 1.
Figure 17. RC net power output as a function of the evaporating pressure for the
working fluids studied using evaporator 1.

41

Table 1. Vehicle test conditions.


Operating
condition

N
[rpm]

F
[N ]

BMEP
V

[km/h]

Pe
[kW ]

m
g
[g /s ]

g ,
T
[K ]

available
Q
[kW ]

2000

12.8

730.9

6.23

2000

500

0.91

31.7

4.26

17.0

790.0

9.47

2000

1000

1.75

30.1

8.18

21.0

829.7

12.71

2000

1500

2.35

26.6

10.96

23.9

850.7

15.07

2000

2000

2.78

23.5

12.96

25.9

868.2

16.89

3000

17.3

807.3

10.00

3000

500

0.98

50.2

6.88

25.8

897.9

17.77

3000

1000

1.95

49.7

13.67

31.5

939.6

23.34

3000

1500

2.85

48.2

19.97

37.9

968.7

29.47

10

3000

2000

3.77

47.2

26.39

43.0

989.8

34.59

11

4000

25.4

869.4

16.60

12

4000

1000

1.98

67.0

18.45

43.0

1001.8

35.25

13

4000

2000

3.98

67.0

37.17

59.7

1052.3

52.81

Table 2. Equations used to calculate the exhaust gas properties(a).


Specific heat capacity
[J kg-1 K-1]
Dynamic viscosity
[N s m-2]
Prandtl number
Thermal conductivity
[W m-1 K-1]
Density [kg m-3]

c pg =956.0+ 0.3386 T g2.476 105 T g2


g =106 ( 3.807+ 4.731 102 T g9.945 106 T g2 )
Pr=0.774+1.387 104 T g +1.863 107 T g2+ 7.695 1011 T g3
k g=103 ( 4.643+6.493 102 T g )

g =1.6652.404 103 T g +1.121 106 T g2


(a)
The equations are valid for 400 T g 1200 K .

42

Table 3. Main thermophysical properties of the working fluids studied (R123, R245fa
and water).
Working
fluid

Category

Water
R123
R245fa

HCFC
HFC

pcr
[MPa ]
22.06
3.66
3.64

T cr
[K ]

Normal boiling
temperature
[ ]

pcond, T = 323 K

Slope of the
saturation vapor line

746.95
456.68
427.15

100
27.8
14.9

0.123
2.125
4.012

Negative
Positive
Positive

Table 4. Sensitivity assessment of parameters used in the RC thermodynamic model.


Parameter
Expander efficiency, t
Pump efficiency, p
Water evaporating
pressure, pevap
Water condensation
pressure, pcond
Organic fluids evaporating
temperature, Tevap
Organic fluids
condensation temperature,
Tcond

Variation considered in
the parameter
+/- 10%
+/- 10%

Resulting variation in the RC net


power output (%)
+/- 16
+/- 0.1

+/- 0.1 bar

+/- 1

+/- 0.1 bar

-/+ 4

+/- 10 C

+/- 5

+/- 10 C

-/+ 9

Table 5. Main characteristics of evaporators 1 and 2.


Characteristic
Number of tubes, N t
Tubes diameter, d i (m)
Distance between tubes, (m)
(a)

Tubes length, L (m)


Thickness, e (mm)
External diameter, d ext (m)
(a)

Evaporator 1
43
0.01

Evaporator 2
38
0.006

0.004

0.002

0.5

0.7
0.5

0.11

Tubes in an equidistant hexagonal arrangement.

43

Table 6. Heat exchanger characteristics for evaporators 1 and 2 (operating condition 13,
see Table 1).
Evaporator

Working
fluid

Thermal resistance,
1
[K W-1]
UA

Water

0.027

R123

0.039

R245fa

0.034

Water

0.024

Area
[m2]

Overall heat transfer


coefficient,
U [W m-2 .K-1]
55.50

0.675

38.04
43.69

0.501

75.17

Table 7. Increase in the ICE thermal and vehicle mechanical efficiencies for the three
RCs studied.
RC
1

Working
fluid
Water
R123
R245fa

Increase of thermal efficiency [%] Increase of mechanical efficiency [%]


(3)
(9)
(13)
(3)
(9)
(13)
2.11
2.98
3.52
15.24
15.95
15.94
1.77
2.51
2.97
12.83
13.43
13.43
1.40
1.99
2.35
10.16
10.64
10.63

Water
R123
R245fa

0.36
0.96
0.85

0.96
1.15
1.03

1.20
1.15
1.06

2.64
6.96
6.18

5.14
6.15
5.53

5.41
5.23
4.79

Water

0.30

0.72

0.85

2.17

3.83

3.87

44

Qin

(3)

Heat exchanger
Wt

Turbine
(2)
Wp

Pump

Condenser

(4)

Qout

Figure 1. Schematic of a simple RC.

Air

Exhaust gases
TWC
2

Heat exchanger

3
Generator

Pump

Turbine

4
Condenser
Air + fuel

Figure 2. Schematic of a typical RC waste heat recovery system from ICE exhaust
gases.

45

1000
Ideal
Real

a) Water

900
800

T (K)

700
600
500
400
300
200

1000 2000 3000 4000 5000 6000 7000 8000 9000

800

b) R123
700

T (K)

600
500
400
300
200
600

800

1000

1200

1400

1600

1800

800

c) R245fa
700

T (K)

600
500
400
300
200
600

800

1000

1200

1400

1600

1800

2000

s (J/kgK)

Figure 3. T-s process diagrams. a) Water, b) R123 and c) R245fa.

46

T (K)

Tg,in

T3

Tg,pp
Tg,out

Tpp

Working fluid

T2x

T2
Q (kW)

diagram used for the pinch-point analysis in the RC


Figure 4. Schematic of the T- Q
heat exchanger.

47

Working fluid,
Tcond, pcrit

Input variables

Thermodynamic
conditions

Database
interface
Refprop 9.0

Calculate
efficiency and
thermophysics
properties

For each
evaporating
pressure

Flow balance in
evaporator
For each
experimental
condiction
Calculate power
output and
working fluid flow

Show results

Figure 5. Algorithm for the RC thermodynamic model.

48

30000

70

a)

60

25000

Nu

Re

15000
10000

30

10
0

10 20 30 40 50 60 70 80 90 100

100

120

c)

90

110

10 20 30 40 50 60 70 80 90 100

d)

80
70

100

(%)

h (W m-2 K-1)

40

20

5000

90

60
50
40

80

30

70
60

Square
Rectangular

50

20000

b)

20
10 20 30 40 50 60 70 80 90 100

300

10

10 20 30 40 50 60 70 80 90 100

1.5

e)

f)
(x 105 Pa)

(m2 m-3)

250
200
150

0.5

100
50

10 20 30 40 50 60 70 80 90 100

Nt

10 20 30 40 50 60 70 80 90 100

Nt

Figure 6. Thermal and hydraulic characteristics of the exhaust gases as a function of the
number of tubes in the evaporator for different cross flow geometries. a) Re, b) Nu, c) h,
d) , e) , f) p.

49

Tg,out

Tg,pp
Preheating
zone

Tg,pp2
Evaporating
zone

T2

Tg,in
Superheating
zone

T3x

T3

Figure 7. The three zones of the heat exchanger considered for modeling purposes.

50

Results of the
thermodynamic
model

Define
geometrical
characteristics

Initial conditions:
temperature and
lenght
for each sub-component

Interface with
Refprop 9.0

Exhaust gas
characteristics
Repeat
calculations
for each
sub-component
-NUT method

Calculate
flow balance
for each
sub-component

Calculate
exhaust gases
temperature and
sub-component
lenght

Calculate
heat exchanger
power output
and e fficiency

Figure 8. Algorithm for the heat exchanger model.

51

Run
until
convergence

0.2
0.18
0.16
0.14
0.12
0.1
0.08
0.06
Water
R123
R245fa

0.04
0.02
0

2
3
Evaporating pressure (MPa)

Figure 9. RC efficiency as a function of the evaporating pressure for the working fluids
studied.

30
Water
R123
R245fa

25

v4 / v3

20
15
10
5
0

2
3
4
Evaporating pressure (MPa)

Figure 10. Turbine outlet/inlet expansion ratio (v4/v3) as a function of the evaporating
pressure for the working fluids studied.

52

Working fluid flow (kg/s)

0.25

0.2
Op. condition 13
Op. condition 9
Op. condition 3

0.15

Water
R123
R245fa

0.1

0.05

2
3
4
Evaporating pressure (MPa)

Figure 11. Working fluid mass flow rate as a function of the evaporating pressure for
the working fluids studied.

8
Op. condition 13
Op. condition 9
Op. condition 3

Net power output (kW)

Water
R123
R245fa

6
5
4
3
2
1
0

2
3
Evaporating pressure (MPa)

Figure 12. RC net power output as a function of the evaporating pressure for the
working fluids studied.

53

11
Water, op. condition 9
R123, op. condition 9
R245fa, op. condition 9

Exergy destruction rate (kW)

10
9
8
7
6
5
4
3

2
3
4
Evaporating pressure (MPa)

Figure 13. RC evaporator exergy destruction rate as a function of the evaporating


pressure for the working fluids studied.

Evaporator exergy efficiency ()

0.45
0.4
0.35
0.3
0.25
0.2
0.15

Water, op. condition 9


R123, op. condition 9
R245fa, op. condition 9

0.1
0.05

2
3
Evaporating pressure (MPa)

Figure 14. RC evaporator exergy efficiency as a function of the evaporating pressure


for the working fluids studied.

54

0.14
Op. condition 13
Op. condition 9
Op. condition 3

Working fluid flow (kg/s)

0.12

Water
R123
R245fa

0.1
0.08
0.06
0.04
0.02
0
0.5

1.5
2
2.5
Evaporating pressure (MPa)

3.5

Figure 15. Working fluid mass flow rate as a function of the evaporating pressure for
the working fluids studied using evaporator 1.
100
Op. condition 13
Op. condition 9
Op. condition 3

90
80

Water
R123
R245fa

70
(%)

60
50
40
30
20
10
0
0.5

1.5
2
2.5
Evaporating pressure (MPa)

3.5

Figure 16. Heat exchanger effectiveness as a function of the evaporating pressure for
the working fluids studied using evaporator 1.

55

Net power output (kW)

2.5

Op. condition 13
Op. condition 9
Op. condition 3

Water
R123
R245fa

2
1.5
1
0.5
0
0.5

1.5
2
2.5
Evaporating pressure (MPa)

3.5

Figure 17. RC net power output as a function of the evaporating pressure for the
working fluids studied using evaporator 1.

56

You might also like