You are on page 1of 9

Energy and Buildings 42 (2010) 16011609

Contents lists available at ScienceDirect


Energy and Buildings
j our nal homepage: www. el sevi er . com/ l ocat e/ enbui l d
Development of system concepts for improving the performance
of a waste heat district heating network with exergy analysis
Herena Toro

, Dietrich Schmidt
Fraunhofer Institute for Building Physics, Department of Energy Systems, Gottschalkstrasse 28a, 34127 Kassel, Germany
a r t i c l e i n f o
Article history:
Received 26 January 2010
Received in revised form 31 March 2010
Accepted 8 April 2010
Keywords:
District heat
Waste heat
Exergy efciency
Energy efciency
Buildings
Improved performance
a b s t r a c t
The building sector is responsible for a great share of the nal energy demand and national CO
2
emis-
sions in countries like Germany. Nowadays, lowquality thermal energy demands in buildings are mainly
satised with high-quality sources (e.g. natural gas red in condensing boilers). Exergy analysis, pur-
suing a matching in the quality level of energy supplied and demanded, pinpoints the great necessity
of substituting high-quality fossil fuels by other low quality energy ows, such as waste heat. In this
paper a small district heating system in Kassel (Germany) is taken as a case study. Results from prelim-
inary steady-state and dynamic energy and exergy analysis of the system are presented and strategies
for improving the performance of waste-heat based district heating systems are derived. Results show
that lowering supply temperatures from95 to 57.7

C increases the nal exergy efciency of the systems


from 32% to 39.3%. Similarly, reducing return temperatures to the district heating network from 40.8 to
37.7

C increases the exergy performance in 3.7%. In turn, the energy performance of all systems studied
is nearly the same. This paper shows clearly the added value of exergy analysis for characterising and
improving the performance of district heating systems.
2010 Elsevier B.V. All rights reserved.
1. Introduction
The built environment is responsible for around 40% of the nal
energy use in Germany [1]. Particularly space heating and cool-
ing and domestic hot water supply represent the biggest share
of energy demands in residential buildings. For these demands,
mainly fossil fuels are used (e.g. condensing boilers), causing great
CO
2
emissions and thereby making a more efcient use of energy
in this sector absolutely necessary. Exergy analysis allows for the
detection and quantication of the improving potential of complex
energy systems [2,3] and has beenwidely used for the optimization
of thermodynamic systems (e.g. power plants) since the middle
of the last century [4]. Similarly, applying the exergy concept to
the building sector delivers more complete information on the use
of the energy ows and opens up room for further insight and
improvements within this eld [5,6].
Exergy is the maximum theoretical work obtainable from the
interaction of a system with its environment, until a state of equi-
librium is reached between them [7]. Consequently, exergy is a
measure of the potential of a given energy ow to be transformed
into high-quality energy. Exergy demands for space heating (SH)
and domestic hot water (DHW) production in buildings are low,
Abbreviations: DH, district heating; DHW, domestic hot water; SH, spaceheating.

Corresponding author.
E-mail address: herena.torio@ibp.fraunhofer.de (H. Toro).
due to the lowtemperature level demanded for these applications.
In most cases, however, this demand is met by high grade energy
sources, such as fossil fuels or electricity which could instead be
used for other higher-grade applications (e.g. power production).
Therefore, besides the well known issue of energy saving, a wide
margin for exergy saving exists within the built environment. Sub-
stituting high-quality fossil fuels used to supply energy demands
in buildings strongly reduces this margin, i.e. increases greatly the
exergy performance of the built environment. Thereby the exergy
approach allows reducing CO
2
emissions caused by the building
sector.
Waste heat available e.g. from combined heat and power pro-
duction (CHP) plants is a low quality energy ow suitable for
supplying the requested energy at an appropriate quality level. The
use of waste heat with lowexergy content allows a good matching
between the exergy level of the demand and supply sides and rep-
resents, thus, a very efcient manner of supplying thermal energy
demands in buildings.
Several authors have performed steady-state exergy analysis
of district heating systems [812]. In [11] the effect of different
reference temperatures on the performance of a district heating
system is studied. Variable reference temperatures between 0
and 25

C are chosen for the reference environment. Values of the


energy and exergy efciency for these conditions vary between
3849% and 4547% respectively. Exergy efciency varies only 2%
being less sensitive than energy efciency. In [12] the energy and
exergy performance of a geothermal district heating systemin four
0378-7788/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.enbuild.2010.04.002
1602 H. Toro, D. Schmidt / Energy and Buildings 42 (2010) 16011609
Nomenclature
Symbols
.
Q heat transfer rate (W)

Ex exergy rate (W)


m mass ow rate (kg/s)
c specic heat capacity (kJ/kgK)
Ex exergy (J)
P electric energy (J)
Q heat transfer (J)
T absolute temperature (K)
temperature (

C)
exergy efciency
energy efciency
H

T
specic heat transfer coefcient of building
(W/m
2
K)
U heat transfer coefcient (W/m
2
K)
Indexes
0 reference
ave average
dem demand
DH district heating
DHW domestic hot water
el electric
FH oor heating
n nal
heater heater
HP heating period
in inlet
p constant pressure
prim primary
pumps pumps
ret return
sec secondary
SH space heating
sp setpoint
steady steady-state
winter days is analyzed. Average outdoor temperatures for each
of the four winter days are considered as reference temperatures.
Energy and exergy efciencies are on the ranges of 3742% and
4246% respectively. In [8] the authors investigate the inuence of
different supply and return temperatures from the district heating
network in the overall performance of a district heating system.
Outdoor air temperature for a typical winter day in Germany
is regarded as reference temperature (0

C). For a given return


temperature of 30

C, exergy increases from25 to 50% if the district


heating supply temperatures are lowered from 130

C to 40

C.
However, district heatingsystems operateat temperatures close
to the reference environment and, thus, their performance may be
strongly inuenced by the dynamic behaviour of the system and
outdoor air conditions (takenas reference for exergy analysis here).
Inthis paper, the energy andexergy performance of different possi-
ble congurations of a small district heating systemtaken as a case
study are compared. Toallowanaccurate comparisonandshowthe
differences in the performance of the analyzed systems dynamic
energy and exergy analysis are mandatory. Dynamic energy and
exergy assessment of the systems is performed here with TRNSYS
Simulaton Environment [13]. Results show clearly the benets of
the exergy approach for evaluating district heat supply systems in
buildings.
The paper is organized as follows: in Section 2 the district heat-
ing system case study used here is briey introduced. A simplied
Table 1
Main assumptions describing the single family houses in the case study.
Magnitude Value Unit Magnitude Value Unit
U external walls 0.28 W/m
2
K Inltration rate, n 0.6 h
1
U ground oor 0.30 W/m
2
K Internal gains 5 W/m
2
U roof 0.17 W/m
2
K
in,FH
32

C
H

T
0.35 W/m
2
K ret,FH 27

C
The exergy input into the district heat exchanger would need to be evaluated in this
case as a function of the quality factor of the fossil fuel used in the heat plant.
steady-state assessment of the behaviour of district heating sys-
tems is presented in Section 3. Main conclusions derived from
it are compared to ndings in the literature. Based on the rec-
ommendations from steady-state analysis four different hydraulic
congurations are derived and modelled dynamically in TRNSYS.
These congurations are presented in detail in Section 4. Results
fromdynamic energy and exergy assessment are presented in Sec-
tion 5. Main conclusions from the analyses are summarized in
Section 6.
2. Description of the case study
The small neighborhood of Oberzwehren, planned to be erected
in Kassel (Germany), is taken here as case study. In the initial plans
it consisted of 24 single family houses. District heating supply and
return pipes from the local utility company circulate close to the
residential area. It is planned to use the return pipe, i.e. with lower
available temperatures, to supply domestic hot water and space
heating demands. District heating in Kassel is mainly waste heat
from co-generation power plants.
The single family houses are dened as free-standing two storey
buildings with a net useful area of 184.4m
2
. In Table 1 main
assumptions describing the buildings are shown. They represent
well-insulated new buildings complying with requirements from
the German standard [14].
Small DHW storage tanks of 200l are considered in each house.
This allows reducing peak loads for DHWsupply signicantly from
42kWfor each single family house, to 7kWpeak power per house.
According to [15] for DHW supply in single family houses a tem-
perature of 50

C at the outlet of the DHW supply element must


be ensured at all times. An electric heater located at the outlet of
the tank is foreseen as back-up system for this purpose. For DHW
supply a simultaneity factor of 0.39 is considered [16].
A centralized heat exchanger unit is planned to supply heat to
the small neighbourhood as shown in Fig. 1. In this way, the dis-
trict heating network from the local utility company is decoupled
from the building appliances and systems installed, i.e. mass ow
rate and temperature drop in the district heating network are not
directly determined by the mass owrates and temperature drops
in the building systems (e.g. oor heating systems). All houses are
connected in parallel to the local distribution network (secondary
side of the heat supply), as shown in Fig. 1.
Fig. 1. Simplied scheme of district heat supply to the studied neighbourhood of
Oberzwehren (Germany).
H. Toro, D. Schmidt / Energy and Buildings 42 (2010) 16011609 1603
With this conguration the main energy and exergy inputs into
the system are the heat ow from the primary side of the heat
exchanger, i.e. fromthe district heating network, and the pumping
energy both in the primary and secondary sides. Pumping energy
for low temperature district heating systems is estimated to be
around 0.51% of the thermal energy input in the network [8,17].
As it is shown in Section 5, pumping energy for the secondary side
(i.e. for the local heat distribution within the neighbourhood) rep-
resents a marginal part of the energy and exergy supply. Having
much shorter pipes than the secondary side, less pressure losses
can be presumed in the primary side. Thereby, pumping energy in
the primary side is expected to be even smaller and is disregarded
in this study.
3. First preliminary analysis: steady-state behaviour
Preliminarysteady-state analyses have beencarriedout inorder
to understand the exergetic behaviour of the district heating sup-
ply. As a rst simplication, it is assumed that the secondary
side operates under given conditions, i.e. pumping energy which
depends on the mass ow rates and temperature levels chosen
remains unchanged. The main variable is then, the thermal energy
input from the primary side district heating pipe. Steady-state
behaviour of the district heatingsystemis, thus, characterisedbyits
exergy efciency calculated as shown in Eq. (1). Eq. (1) represents
a rational exergy efciency, since only the nal demands of SHand
DHW(i.e. desired output) are regarded as the output fromthe sys-
tem and only the usable output from the district heating system is
regarded as input (i.e. the exergy contained in the return mass ow
rate of the district heating pipes is not regarded as input in the sys-
tem) [4]. Eq. (1) characterises the overall exergy performance of the
district heating supply, i.e. from supply to nal demand. In other
words, all exergy losses happening in the different energy conver-
sion processes in the supply chain (i.e. heat exchangers, thermal
losses inthe pipes, etc.) are included inthe expressionof the exergy
efciency in Eq. (1).

DH,steady
=

Ex
dem,SH,ave
+

Ex
dem,DHW,ave

Ex
in,DH
=

Q
SH,ave

_
1
T
0,HP,ave
Tr
_
+

Q
dem,DHW,ave
(T
dem,DHW
T
net
)

_
(T
dem,DHW
T
net
) T
0,HP,ave
ln
_
T
dem,DHW
T
net
__
m
prim,DH
c
p
_
(T
in,prim,DH
T
ret,prim,DH
) T
0,HP,ave
ln
_
T
in,prim,DH
T
ret,prim,DH
__ , (1)
where

Ex
dem,SH,ave
is the average exergy loadfor space heating (SH),
calculated as a function of the average energy load for SH
.
Q
SH,ave
whichherehas a valueof 76.43kW, andtheabsolutevalueof design
indoor air temperature T
r
(assumed to be 293K). The average load
for domestic hot water (DHW) supply

Ex
dem,DHW,ave
is calculated
as a function of the average energy load for DHW supply
.
Q
DHW,ave
which here has a value of 9.54kW, the absolute value of the tem-
perature demanded for DHW supply T
dem,DHW
(i.e. 323K) and the
temperature fromthe local cold water network T
net
(assumed to be
283K).

Ex
in,DH
is the thermal exergy input fromthe primary side of
the district heating network. It is important to remark that

Ex
in,DH
can be calculated in this way as long as it is waste heat available,
otherwise, e.g. if it would be heat from a fossil fuel powered heat
plant the expression for estimating the exergy input shown here is
not valid.
1
Primary side mass ow rate m
prim,DH
, inlet and return
temperatures T
in,prim,DH
and T
ret,prim,DH
are varied in order to check
the efciency of district heat supply for different operating con-
ditions. T
0,HP,ave
is the absolute value of the average outdoor air
temperature during the heating period (HP), i.e. October to April,
which here has a value of 4.8

C. For estimating the primary side


1
The exergy input input into the district heat exchanger would need to be evalu-
ated in this case as a function of the quality factor of the fossil fuel used in the heat
plant.
Fig. 2. Exergy efciency for different primary side supply and return temperatures
(
in,DH
and ret,DH respectively) and primary side mass ow rates. Only thermal
energy ows are regarded here.
mass owrate required to supply the given SHand DHWdemands,
Eq. (2) is used.
m
prim,DH
=
(

Q
SH,ave
+

Q
dem,DHW,ave
) (1 + f
ls,SH+DHW
)
c
p
(T
in,prim,DH
T
ret,prim,DH
)
, (2)
where the factor f
ls,SH+DHW
depicts the share of thermal losses in
the supply network and DHW storage tanks, and is considered as
0.17 in this case.
Fig. 2 shows steady-state exergy efciencies for supplying the
average SHand DHWloads at different sets of operating conditions
(inlet, outlet temperatures and mass ow rates). Mass ow rates
and temperature levels in Fig. 2 correspond to the primary side of
the heat exchanger in Fig. 1. Continuous lines indicate the exergy
efciency for different primary side return temperatures. Dashed
lines represent exergy efciencies for different primary side supply
temperatures.
Lower supply temperatures increase the exergy efciency of the
heat supply, i.e. lower the quality at which the heat owis supplied
tothe single family houses allowing a better matching of the energy
supply and demand. Lower return temperatures also increase sig-
nicantly the exergy efciency of the heat supply. Similar results
were found by Dtsch and Bargel [8] and in [10]. It is important to
remark that for a given supply temperature, lower return temper-
atures lead to an increase in the exergy efciency. In other words,
maximizing the degradation of the thermal potential of the pri-
mary mass ow increases the exergy efciency. These trends and
strategies for improved operation of district heating networks are
coherent with those found in [8]. However, Dtsch and Bargel [8]
analyze a district heating network without hydraulic separation by
means of aheat exchanger, i.e. thewholedistributionnetworkis the
primary side of energy supply. In consequence, different mass ow
rates and temperature drops have a strong inuence on the elec-
tricity demand for the pumps in the network and pumping energy
1604 H. Toro, D. Schmidt / Energy and Buildings 42 (2010) 16011609
Fig. 3. Exergy efciency for different reference temperatures. The same conditions of primary side supply and return temperatures (
in,DH
and ret,DH respectively) and
primary side mass ow rates as in Fig. 2 are considered: (a) a reference temperature of 0

C is considered; (b) a reference temperature of 10

C is considered.
demands were included in the balances. For each set of operat-
ing conditions , Dtsch and Bargel [8] size a distribution network,
whereby pressure losses and electricity demands for the pumps
can be estimated. This leads to optimum values of the tempera-
ture drop at the primary side of about 10K, with lower values of
the temperature difference leading to a greater importance of the
pumpingenergyand, thereby, decreasingthe exergyefciency. Yet,
for a givensupplytemperature, lower returntemperatures increase
the exergy efciency, being similar to conclusions from Fig. 2.
As stated above, in Fig. 2 the average outdoor air temperature
during the heating period (4.8

C) has been regarded as reference


temperature. InFig. 3 the inuence of the reference temperature on
the results from exergy analysis is graphically shown. In Fig. 3(a)
and (b) a reference temperature of 0

C and 10

C are considered
respectively. The variation of the thermal energy losses in the pipes
has beenestimatedinaccordancewith[8]. It canbeclearlyseenthat
lower reference temperatures (i.e. Fig. 3(a) corresponding to e.g.
colder climates) increase the exergy efciency of the district heat-
ing systemas compared to Fig. 2. Exergy efciencies in Fig. 3(a) are
in very good agreement with those shown in [8], where the same
reference temperature is chosen. In turn, increasing the reference
temperature decreases the exergy efciency of the district heat-
ing system (see Fig. 3(b)). This trend is mainly due to the greater
reduction in the exergy demands (numerator in Eq. (1)) as com-
pared to the exergy supply (denominator in Eq. (1)) for higher
reference temperatures. Yet, regarding the inlet and return tem-
peratures fromthe district heating pipe, the same trends as in Fig. 2
can be observed.
FromFig. 2, two mainstrategies to increase the exergy efciency
of district heating supply can be derived:
1. Minimizereturntemperaturetothedistrict heatingnetwork:
this can be achieved by proper sizing of the heat exchanger
between the district heating pipe and the local heat distribution
network. However, minimum achievable return temperatures
to the district heating pipe strongly depend on the building
systems, e.g. low temperature space heating systems allow
lower return temperatures. Therefore, the use of appropriate
(i.e. low temperature) building systems is of great importance
for promoting this strategy. The use of lowtemperature systems
succeeds only if existing heating loads are low enough. For this
purpose an improved building shell is necessary, i.e. increased
insulation level and careful design are therefore a must.
2. Minimize supply temperature from the district heating net-
work: supply temperatures are determined by the temperature
prole available from the district heating pipe and cannot be
directly inuenced. However, according to German regulations
for DHWpreparation a supply temperature of 50

C must always
be ensured [15]. In turn, maximum required inlet temperature
for space heating is 32

C. Thus, a way to reduce the supply tem-


perature for the lowtemperature use (i.e. space heatingdemand)
is by cascading the energy demands according to their required
temperature level.
4. Models for dynamic energy and exergy analysis
According to the strategies developed frompreliminary steady-
state analysis shown in Fig. 2, four different system congurations
have been derived and are analyzed. A schema of their congura-
tion and main features are shown in Fig. 4:
- SystemI corresponds to a conventional high temperature district
heat supply, with a supply temperature at the primary side of
95

C. DHW and SH supply are combinedly delivered by a single


centralized heat exchanger.
- System II corresponds to a low temperature district heat supply
with primary side supply temperatures between 50 and 65

C.
DHW and SH supply is also done with a single centralized heat
exchanger.
- SystemIII is a lowtemperature district heat supply with primary
side supply temperatures between 50 and 65

C. DHW and SH
supply is performed via separated heat exchangers. In this way,
each of the heat exchangers can be sized separately for mini-
mizing primary side return temperatures at its specic operation
conditions. Following, lower outlet temperatures at the primary
side of the heat exchanger are expected.
- In System IV a three way valve connects the primary side return
from DHW supply with the primary side for SH supply. If return
temperatures and mass ow rates from DHW supply are high
enough to supply SH demands mass ow from DHW supply will
circulate also through the SH heat exchanger. Otherwise, mass
ow rate is withdrawn directly from the main district heating
pipe. In this way, cascading of DHW and space heating demand
is achieved. The setpoint for the secondary side of DHW supply
has been kept as a function of the primary side inlet temper-
ature similarly as in the previous cases (see Eq. (3)). However,
primary side mass ow rate from the DHW heat exchanger can
only be used for supplying SH demands (i.e. cascaded) if its tem-
perature is higher than the required SHsetpoint. Thus, secondary
side setpoint for SH supply has been minimized: it is dened as
H. Toro, D. Schmidt / Energy and Buildings 42 (2010) 16011609 1605
Fig. 4. Four system congurations for district heat supply studied here.
a function of the required inlet temperature for the oor heating
(FH) systems
sp,FH
(see Eq. (4)).
In case I, with a conventional high temperature supply, a
constant supply temperature from the district heating network
(primary side in Figs. 1 and 2) of 95

C is assumed. For options II,


III and IV with low temperature supply, a temperature prole as a
function of outdoor air temperature is assumed. Supply tempera-
ture
in,prim,DH
is assumed to vary between 50 and 65

C, according
to the prole shown in Fig. 5(a) [18]. Fig. 5(b) shows the setpoint
inlet temperatures for the FHsystems usedfor space heatingsupply

sp,FH
as a function of outdoor air temperature.
In order to minimize the pumping power for the secondary side,
minimummass owrates needtobe achieved. For this purpose, the
setpoint of the secondary side of the network providing DHW and
SH to the single family houses
sp,sec
is maximized in cases II and
III. As shown in Fig. 4 and Eq. (3), the setpoint is chosen as a func-
tion of the inlet temperature in the primary side of the supply heat
exchangers
in,DH
, i.e. the supply temperature available from the
district heating network. Comparing cases I and II the inuence of
lower supply temperatures from the network can be investigated.
To show clearly the inuence of this parameter on the exergy per-
formance, all other variables need to be the same in both cases.
Thus, thermal energy losses and energy performance of the sec-
ondary side needs to be the same in both cases. For this aim, in case
I a constant setpoint of 55

C has been chosen for the secondary


side. Higher setpoint temperatures lead to higher temperatures in
the upper layers of the DHWtanks, and thus, higher thermal losses
in the storage tanks occur. In turn, with 55

C as setpoint the energy


supplied to both systems and their energy behaviour are similar.
For supplying the required temperature to the oor heating sys-
tems
sp,FH
a mixing valve for recirculating part of the cold return
water from FH systems is foreseen in all cases.

sp,sec
(t
k
) =
in,prim,DH
(t
k
) 2 (3)

sp,sec
(t
k
) =
sp,FH
(t
k
) +2 (4)
The four systems under investigation(Fig. 4) have beendynami-
cally simulatedinTRNSYS witha timestepof 3min. Detailedenergy
ows for every energy conversion step in the energy supply sys-
tems (i.e. heat exchangers, pipes in the networks, storage, etc.)
can be obtained. An input output approach has been applied for
exergy analysis. This means that exergy ows are not derived and
analyzed in detail for every energy conversion step on the district
heating system. Instead, only the exergy associated to the energy
supply (input) and demand (output) are regarded. The heat input
in the primary side of the heat exchangers and pumping energy in
the secondary side are the main inputs in the system. Results from
Fig. 5. (a) Primary side supply temperature for the district heating systemas a func-
tion of the outdoor air temperature [18]. (b) Setpoint for the supply temperature of
the oor heating (FH) systems for space heating supply as a function of the outdoor
air temperature.
1606 H. Toro, D. Schmidt / Energy and Buildings 42 (2010) 16011609
dynamic energy assessment required to calculate the exergy ows
(energy, mass ow rates or temperatures) are used as input in the
equations used to perform exergy analysis for every timestep (t
k
).
Eqs. (5)(9) show the mathematical expressions used to calculate
the exergy of heat input, exergy of SH and DHW demands and the
nal exergyandenergyefciencies over theyear (withNtimesteps)
respectively. Exergy ows are thereby obtained for every timestep
(i.e. 3min):

Ex
in,DH
(t
k
) = m
prim,DH
(t
k
) c
p
_
(T
in,prim,DH
(t
k
) T
ret,prim,DH
(t
k
)) T
0
(t
k
) ln
_
T
in,prim,DH
(t
k
)
T
ret,prim,DH
(t
k
)
__
(5)

Ex
dem,DHW
(t
k
) = m
dem,DHW
(t
k
) c
p
_
_
T
dem,DHW
(t
k
) T
net
(t
k
)
_
T
0
(t
k
) ln
_
T
dem,DHW
(t
k
)
T
net
(t
k
)
__
(6)

Ex
dem,SH
(t
k
) =

Q
dem,SH
(t
k
)
_
1
T
0
(t
k
)
T
r
(t
k
)
_
(7)

DH,n
=
k=N

k=1
Ex
dem,SH
(t
k
) +
k=N

k=1
Ex
dem,DHW
(t
k
)
k=N

k=1
Ex
in,DH
(t
k
) +
k=N

k=1
P
pumps
(t
k
) +
k=N

k=1
P
el,heater
(t
k
)
(8)

DH,n
=
k=N

k=1
Q
dem,SH
(t
k
) +
k=N

k=1
Q
dem,DHW
(t
k
)
k=N

k=1
Q
in,DH
(t
k
) +
k=N

k=1
P
pumps
(t
k
) +
k=N

k=1
P
el,heater
(t
k
)
(9)
The aim of this study is to compare the energy and exergy
performance of different congurations for district heat supply.
Therefore, the systems are compared in terms of nal energy and
exergy supply and efciencies.
5. Results and discussion
Fig. 6 shows the specic nal energy and exergy supply for the
four cases studied. Even in the cases with low temperature dis-
trict heating supply, with a temperature prole between 50 and
65

C, the temperature level is high enough to ensure that most of


the time DHW tanks are heated up to the required level of 50

C
without additional energy input. Additional electricity required to
power the instantaneous heater and guarantee the 50

C required
amounts 435kWh/a in the cases with low temperature supply, i.e.
0.098kWh/m
2
a, being a marginal energy input as it can be seen in
Fig. 6.
In terms of nal energy and for the low temperature district
heating supply, pumping energy represents around 0.1% the ther-
mal energy input into the local network. In the literature these
values are estimated to be around 0.5% and 1% [8,17]. The reason
for the signicantly smaller values in the models developed is that
here a criteria of maximum uid velocity of 11.5m/s was cho-
sen for sizing the network pipes. Since the safety of energy supply
was a key issue in the systems investigated, this was done to make
sure that thermal losses in the network are not underestimated.
On the other hand, as stated in Section 2, pumping energy in the
secondary side is one of the main inputs in the system thereby
inuencing strongly its performance. Criteria chosen here to size
the network aim also at minimizing the pumping energy, thereby
allowing an optimized performance of the systems. In turn, veloci-
ties of up to 2.5m/s in the main pipes would be acceptable [8]. This
leads to bigger pipes and lower pressure losses, i.e. lower pumping
power required. Pipes with bigger diameters required for the net-
works sized with the criteria used here (i.e. with target velocities
of 11.5m/s) are estimated to cost around 17% more than simi-
lar networks sized with conventional criteria. However, the pumps
would be smaller in the networks used here (since maximumpres-
sure losses are also smaller) and thereby cheaper. Estimations done
for case II show that the lower prices for the pumps might even
compensate greater costs for the pipes in the hydraulic network.
Final energy input required to supply the different cases
amounts 92.0, 91.7, 92.3 and 93.0kWh/m
2
a for cases I, II, III, and IV
respectively. The greatest difference can be found between cases II
and IVand amounts only 1.4%of the total energy supply. Interms of
nal energysupply, as it canbe seeninFig. 6(a) all systems analyzed
are equivalent. Thereby the energy performance (efciency) of all
cases is very similar, as it can be seen in Table 2.
In turn, the exergy supplied is signicantly different for the dif-
ferent systems studied: 15.8kWh/m
2
a for case I, 12.8kWh/m
2
a for
case II, 11.8kWh/m
2
a for case III and 11.9kWh/m
2
a for case IV.
The greatest difference can be found between cases I and III and
amounts 25.6% of the exergy supply in case I, being thus a relevant
reduction. This leads to an increased exergy performance for case
III as compared to the rest of cases (see Table 2).
Fig. 6. (a) Final energy supply for the four cases studied and (b) Final exergy supply
for the four cases studied.
H. Toro, D. Schmidt / Energy and Buildings 42 (2010) 16011609 1607
Table 2
Final energy and exergy efciencies for the four cases studied. Annual average values of the primary side supply and return temperatures are also shown for each case.
(Final) energy
efciency,
DH,n
[%]
(Final) exergy
efciency
DH,n
[%]
Primary side supply
temperature
in,primDH
[

C]
Primary side return
temperature
ret,primDH
[

C]
I 1HX, 95

C 81.7 32.0 95.0 32.4


II 1HX, 6550

C 81.9 39.3 57.7 40.8


III 2HX, 6550

C 81.4 43.0 57.7 37.7


IV 2HX cascading, 6550

C 80.7 42.4 57.5 37.8


Table 2 shows the nal energy and exergy efciencies for the
four cases analyzed. In addition, annual average values for the pri-
mary side inlet
in,primDH,ave
and return temperatures
ret,primDH,ave
for the cases studied are also shown for being representative for
the operating conditions in each case.
Case I has the lowest exergy efciency, being between 10.4
and 7.3% lower than the rest of systems studied. Case II repre-
sents a system with the same hydraulic conguration but with
lower supply temperatures (5065

C) and shows an exergy input


of 12.8kWh/m
2
a, representingareductionof 19%inthenal exergy
input as compared to case I. This shows the importance of reduc-
ing supply temperatures in district heating networks for increasing
their exergy performance. Cases II, III and IV have a more sim-
ilar exergy supply (Fig. 6b) and, in consequence, a more similar
performance (Table 2).
Values for the exergy efciencies in Table 2 are between 32
and 43%. Similar values can be found in the literature: in [9,12]
exergy efciencies for two geothermal district heating systems are
found to be 50 and 45.66% respectively. Efciencies in [9,12] how-
ever, do not consider the nal demand to be supplied, i.e. space
heating or domestic hot water supply, as the nal output of the sys-
tem. Instead, in [9,12] the thermal exergy output from the district
heating system to the building systems is regarded as the output.
This leads to slightly higher values in the exergy efciencies. On
the contrary, the exergy efciencies in the present paper consider
the exergy demands for SH and DHW supply as the nal output,
regarding therefore a further conversion step. Following, exergy
efciencies presentedhere (see Table 2andFig. 2) are lower. Dtsch
and Bargel [8] also follow this last approach, i.e. they regard the
exergy of SH demands to be supplied as the nal desired output
of the district heating system. Values of the exergy efciencies in
[8] are in very good agreement with those shown in the present
paper.
In [12] the authors conclude that higher supply temperatures
increase the exergy efciency of the district heating system. This
results, contradictorytothoseshowninthepresent paper, areagain
due to the boundary chosen for exergy analysis: as stated above, in
[9] the authors regard the exergy supplied by the district heating
systemas the desired output. In consequence, higher temperatures
of the energy supplied increase the exergy efciency of the system.
Inturn, inthepresent paper theSHandDHWdemands areregarded
as the nal desired output of the district heating system. Supply-
ing these demands with higher temperatures (e.g. by using high
temperature radiators) leads to increased exergy losses inside the
building systems. Following this approach, for a given demand at a
given temperature (e.g. SH at 20

C) lowering the supply tempera-


ture from the district heating network increases the performance
of the system. Similar conclusions can be found in [8], where the
same approach as in the present paper is followed. This shows the
importance andgreat inuence of the chosenboundaries for exergy
analysis on the results and conclusions that can be derived from it.
Comparing cases II and III the inuence of supplying DHW and
SHdemands separately, i.e. with separate heat exchangers and sec-
ondary networks, can be seen. Separate supply (i.e. case III) allows
minimizing primary side return temperatures to the district heat-
ing network (see Table 2). Fig. 7(a) and (b) shows the quality factor
of theenergysuppliedF
Q
, primarysidereturntemperatures
ret,prim
and mass owrates for both cases II and III. For completeness, heat
rate supplied to the heat exchanger(s)
.
Q
in,DH
and the exergy rate
associated to it
.
Ex
in,DH
are also shown. Supply temperatures are the
same at all times, and therefore not shown in Fig. 7. The dynamic
behaviour of the heat transfer is quite similar in both cases. How-
ever, it happens at signicantly different mass ow rates, and in
consequence, different temperature drops occur in the primary
side, i.e. different return temperatures are found. The grey shad-
owedareas inFig. 7showthe correlationbetweenthe quality factor
of the energy suppliedF
Q
andthe primary side returntemperatures

ret,prim,DH
: whenever returntemperatures arelower for anyof both
cases, quality factors are also lower. This means that the quality of
the energy supplieddecreases withlower returntemperatures. The
instantaneous exergysuppliedat this timesteps, however, might be
higher if the energy supplied is also greater. Thereby, nal exergy
Fig. 7. (a) Quality factors FQ and mass owrates m for cases II and III; (b) Exergy
and heat transfer rates supplied Ex and Q and return temperatures ret for
cases II and III.
1608 H. Toro, D. Schmidt / Energy and Buildings 42 (2010) 16011609
Fig. 8. (a) Quality factors FQ andmass owrates m for cases III andIV; (b) Exergy
and heat transfer rates supplied Ex and Q as well as inlet and return in and
ret temperatures for cases III and IV.
input requiredis reducedto11.7kWh/m
2
a, representingadecrease
of 8.4% as compared to case II. This shows the inuence of minimiz-
ing return temperatures for increasing the exergy performance of
district heating systems.
Fig. 8(a) and (b) show the quality factor of the energy supplied
F
Q
, primary side inlet
in,prim
and return temperatures
ret,prim
and
mass owrates for both cases III and IVas well as the heat rate sup-
plied to the heat exchangers
.
Q
in,DH
and the exergy rate associated
to it
.
Ex
in,DH
. Inlet temperatures are the same for bothsystems when
cascading is not possible and lower in case cascading of the mass
owrate for DHWsupply might be used for SHsupply in case IV, as
shown in Fig. 8(b). The grey shadowed areas in Fig. 8 highlight the
correlationbetweenthequalityfactor of theenergysuppliedF
Q
and
the primary side supply temperatures
in,prim,DH
: whenever supply
temperatures are lower for case IV, i.e. when cascading is possible,
quality factors are also lower despite return temperatures at those
timesteps are higher for case IV. Whenlower inlet temperatures for
case IV succeed mass ow rates for this case increase and become
greater than for case III. The instantaneous heat transfer is nearly
the same in both cases at all times. In turn, the quality factor asso-
ciated to the heat transfer for those situations is lower for case IV.
This means that the quality of the energy supplied decreases with
lower supply temperatures. This shows the importance of reducing
supply temperatures when possible.
Unfortunately, cascading between DHWand space heating heat
exchangers succeeds only 485h/a of the 3594h/a when space heat-
ing energy demand exist and it represents only 2% of the total
energy and exergy supplied to space heating. Therefore, the benet
of this improved exergy performance due to cascading represents
a very small amount of the total energy and exergy supplied.
Whencascading is not possible, returntemperatures are slightly
higher for case IV as for case III, leading to also slightly higher qual-
ity factors. In addition, due to the control strategy of the secondary
side network for SH supply in case IV higher mass ow rates are
required, leadingtohigher pumpingpower demands: electricityfor
pumping amounts 0.047kWh/m
2
a in case III and 0.122kWh/m
2
a
in case IV. These adverse factors are responsible for the worse per-
formance of case IV as compared to case III. Yet, pumping power
for operating the secondary network would not vary if cascading of
DHWand SHdemands was possible more often. In turn, if frequent
cascading wouldreduce the quality of the heat suppliedina greater
share than the increase in pumping energy system IV would show
the best performance of all cases. This requires, however, develop-
ing suitable system concepts promoting cascaded use of the mass
ow rate.
6. Conclusions
Four different systems for district heat supplyhavebeendynam-
ically analyzed. As a rst step, simplied steady-state exergy
analysis was performed to show the exergetic behaviour of a dis-
trict heat supply system. Systemconcepts studied were developed
based on conclusions from this analysis. Main conclusions and
trends derived from this preliminary steady-state analysis have
been conrmed by results from dynamic analysis.
The energy performance of the systems studied is very similar.
Maximum differences in the nal energy efciency of the systems
studiedamount only1.2%. Their exergyperformanceshows, inturn,
signicant differences with values for the nal exergy efciencies
from 32% to 43%. These results show that exergy analysis is a more
powerful tool for depicting the performance of the systems ana-
lyzed than mere energy analysis. By depicting the quality level of
the energy supplied, exergy analysis allows using available thermal
energy ows ina more efcient way, i.e. degrading lower amount of
their potential todeliver work(exergy). This might beasuitabletool
for organizing available energy ows in district heating networks
in a more efcient way, where energy supplied and demanded can
be matched in terms of their quality level.
The main conclusions related to the performance of district
heating systems obtained from this study can be summarized as
follows:
- Results from dynamic analysis show that reducing the inlet tem-
peratures from 95

C to an average annual value of 57.7

C allow
increasing the exergy efciency of the district heating system in
7.3%. Similar behaviour is conrmed by simplied steady-state
analysis (Fig. 2).
- Reducing the average annual return temperature in 3.1

C leads
to an increase in the exergy efciency of 3.7%. Thus, it can be
stated that lower average annual return temperatures to the dis-
trict heating network lead to an increase in the exergy efciency
of the system. One strategy to achieve lower return temperatures
consists on supplying SH and DHW demands with separate heat
exchangers. This allows sizing andoperating eachheat exchanger
so as to minimize return temperatures to the district heating
network at all times.
- Supply temperatures have a greater inuence on the exergy
performance of the district heating system studied than return
temperatures. This has been shown by the comparison of cases
III and IV (cascaded supply of DHW and SH demands). When
H. Toro, D. Schmidt / Energy and Buildings 42 (2010) 16011609 1609
cascading between DHW and SH demands is possible lower sup-
ply temperatures and higher return temperatures are found for
case IV. At lower supply temperatures (and despite simultane-
ous higher return temperatures) similar heat transfer happens
at lower quality levels (F
Q
), i.e. less exergy input is required to
provide the same thermal energy demands. This indicates that
the rst step for a more efcient district heat supply should be
to reduce supply temperatures. For this aim, suitable network
concepts need to be developed which allow energy demands
at different temperature levels to be supplied at a suitable
temperature.
- Due to the low simultaneity between DHW and SH demands
cascading between DHW and SH heat exchangers succeeds only
485h/aof the3594h/awhenspaceheatingenergydemandexists,
representing only 2% of the total energy and exergy supplied to
space heating. Therefore, the benet of the improved exergy per-
formance due to cascading represents a very small amount of the
total energy and exergy supplied. Yet, if cascading would succeed
more often the performance of case IV might greatly increase.
Thus, hydraulic concepts promoting a cascaded supply of DHW
and SH demands (e.g. by means of storage concepts) need to be
developed and investigated.
- For cascaded supply between DHW and SH demands higher
mass ows are required on the secondary side for SH supply. An
increase inthe pumping power requiredto operate the secondary
network follows. In the case studied, the increase in the pumping
energy was greater than the benets from cascaded supply from
an exergy perspective. Thus, the performance of cascaded supply
(case IV) as compared to separated supply (case III) is reduced in
0.6%.
Acknowledgement
The authors warmly thank the German Federal Foundation for
Environment and the German Federal Ministry of Economy and
Technology for their nancial support.
References
[1] G. Hauser, Energieefzientes BauenUmsetzungsstrategien und Perspektiven,
FVEE Themen (2008) 819.
[2] J. Ahern, The Exergy Method of Energy System Analysis, Wiley Interscience
Publication, John Wiley and Sons, New York, 1980.
[3] J. Szargut, Exergy Method - Technical and Ecological Applications, WIT Press,
Southampton, United Kingdom, 2005.
[4] H. Toro, A. Angelotti, D. Schmidt, Exergy analysis of renewable energy-
based climatisation systems for buildings, Energy and Building 41 (2009)
248271.
[5] D. Schmidt, New ways for energy systems in sustainable buildingsincreased
energy efciency and indoor comfort through the utilisation of low exergy
systems for heating and cooling of buildings, in: Proceedings of the Plea2004
Conference on Passive and Low Energy Architecture, Eindhoven, the Nether-
lands, 2004, pp. 1922.
[6] H. Torio, D. Schmidt, More sustainable buildings through exergy analysissolar
thermal and/or ventilationsystems? in: Proceedings of the International CLIMA
2007 Wellbeing Indoors Conference, Helsinki, Finland, 2007.
[7] M.J. Moran, H.N. Shapiro, Fundamentals of Engineering Thermodynamics, 3rd
edition, John Wiley & Sons, New York, USA, 1998.
[8] C. Dtsch, S. Bargel, Theoretische Betrachtung zur idealen Temperatur in Fer-
nwrmenetzen auf Grundlage einer exergetischen Bewertung, in: Proceedings
of the LowEx Symposium, Kassel, Germany, 2009.
[9] Y. Kalinci, A. Hepbasli, I. Tavman, Determination of optimum pipe diam-
eter along with energetic and exergetic evaluation of geothermal district
heating systems: modeling and application Energy and Buildings 40 (2008)
742755.
[10] K. Comakli, B. Zksel, . Comakli, Evaluation of energy and exergy
losses in district heating network, Applied Thermal Engineering 24 (2004)
10091017.
[11] L. Ozgener, A. Hepbasli, I. Dincer, Effect of reference state onthe performance of
energy and exergy evaluation of geothermal district heating systems: Balcova
example, Building and Environment 41 (2006) 699709.
[12] L. Ozgener, A. Hepbasli, I. Dincer, Energy and exergy analysis of geothermal
district heating systems: an application, Building and Environment 40 (2005)
13091322.
[13] TRNSYS 16. A TRnsient SYstem Simulation program, version 16. Solar Energy
Laboratory, University of Winsconsin-Madison, 2007.
[14] EnEv:09. Verordnung zur nderung der Energieeinsparverordnung. Bunde-
sanzeiger Verlag, April 2009.
[15] AGFW-Merkblatt FW526. Thermal reduction of legionella
growthimplementation of DVGW working sheet W 551 in district heat
supply. Arbeitsgemeinschaft fr Fernwrme, AGFW 2009.
[16] H. Recknagel, E. Sprenger, E.R. Schramek, Taschenbuch fr Heizung und Kli-
matechnik. Oldenbourg Industrieverlag 07/08, 2007.
[17] Olsen, et al., A new low-temperature district heating system for low-energy
buildings, in: The 11th International Symposiumon District Heating and Cool-
ing, Reykjavik, Iceland, 2008.
[18] J. Kaiser. KFW-09-05-13 Calculations FW. Internal working document. Fraun-
hofer Institute for Building Physics, Kassel (Germany), 2009.

You might also like