Você está na página 1de 11

Energy Conversion and Management 95 (2015) 149–159

Contents lists available at ScienceDirect

Energy Conversion and Management


journal homepage: www.elsevier.com/locate/enconman

Calculation of first-law and second-law-efficiency of a Norwegian


combined heat and power facility driven by municipal waste
incineration – A case study
Tore Solheimslid, Hanne Katrin Harneshaug, Norbert Lümmen ⇑
Department for Mechanical and Marine Engineering, Bergen University College, Postboks 7030, 5020 Bergen, Norway

a r t i c l e i n f o a b s t r a c t

Article history: Both first-law and second-law-efficiency of a combined heat and power plant were calculated. The plant
Received 13 October 2014 is located in Bergen, Norway. Both household and industrial waste is converted into electricity and dis-
Accepted 7 February 2015 trict heating by incineration. The fictive molecule C6H10O2.41N0.1S0.01 represents the average chemical
Available online 27 February 2015
composition of the waste. The chemical exergy of the waste is calculated in several different ways.
Calculations of chemical exergy based on the heating value and entropy change during combustion are
Keywords: compared to direct calculation by means of correlation functions. The results obtained with the different
Municipal waste incineration
methods are in good agreement. Based on the lower heating value of the waste an energy utilization fac-
Exergy efficiency
Absolute entropy
tor of 40.6% is found for the combined heat and power plant in this case study. Its second-law-efficiency
Chemical exergy is 17.3%. The effect of moisture in the incinerated waste on the different efficiencies is investigated as
well. It appears that analyses of the waste composition seem to underestimate the total moisture in
the waste burnt in the incinerators. This in turn underestimates the actual efficiencies and the second
law efficiency could be at 20% or more in the presented case.
Ó 2015 Elsevier Ltd. All rights reserved.

1. Introduction dumped in landfills. Ng et al. recently evaluated the potential of


energy recovery from municipal solid waste in Malaysia [7] by
In the past decades, municipal waste has undergone a transfor- studying a proposed supply network for municipal waste treat-
mation from being a problem to being a valuable resource. The ment that included different operations like, for example, compost-
management of waste and its sub processes like collection, trans- ing, incineration and gasification.
port, processing, etc. have become more of an important task than Waste incineration has undergone the development from sim-
before [1]. Advances in combustion technology and the need to ply reducing the waste volume to being an economic and energy
find a replacement for dumping waste in landfills have led to a efficient operation. It has the potential to partially replace fossil
situation where more and more waste is turned into useful energy fuels in heat and electricity production on a local scale. This can
like electricity and district and process heat. Energy recovery from be important for societies that want to reduce the use of and their
municipal waste can be operated in a cost-effective way and can dependency on fossil fuels and base their energy supply on, for
provide energy with a low carbon footprint to communities [2]. example, combined heat and power plants based in waste incin-
In 2009, 1.1 million tons of waste were delivered to the 19 waste eration or renewable fuels instead and reduce the number of large
incineration plants in Norway, which were operative at this time fossil fuel based power plants.
[3]. While municipal waste incineration is an established activity Reporting of recovered and delivered energy from waste to
in many European countries [4], its application is still under review households and the corresponding energy efficiency to the public
and development in other parts of the world. In Saudi Arabia [5], is not necessarily standardized, although simple definitions for
energy recovered from waste is considered to be part of the coun- thermal efficiency, utilization factor, or second-law-efficiency are
try’s future energy mix due to an increasing demand for electricity. available from standard textbooks on thermodynamics (for exam-
In India, Chakraborty et al. [6] have studied the possibilities of ple [8]). At the other end of the scale, complicated frameworks for
energy generation from Dehli’s municipal waste that is usually defining and calculating the efficiency of energy recovery opera-
tions exists, like the R1-formula introduced by the European Union
⇑ Corresponding author. [9]. Among the various possibilities to define efficiency it is attrac-
E-mail address: nlu@hib.no (N. Lümmen). tive for operators to employ definitions that express the successful

http://dx.doi.org/10.1016/j.enconman.2015.02.026
0196-8904/Ó 2015 Elsevier Ltd. All rights reserved.
150 T. Solheimslid et al. / Energy Conversion and Management 95 (2015) 149–159

Nomenclature

Symbol Explanation (Unit) Ru universal gas constant (J/mol K)


A weight percent ash on dry basis (wt% dry basis) S weight percent sulphur on dry basis (wt% dry basis)
A1 universal constant in Eq. (14) (g/mol) S entropy (kJ/K)
A2 universal constant in Eq. (14) (g2/mol2) s amount of sulphur in Eq. (16) (kmol/kg (daf))
C weight percent carbon on dry basis (wt% dry basis) s0 specific entropy relative to 0 K (kJ/kg K)
C1 universal constant in Eq. (14) (g J/mol2 K) s molar entropy (kJ/kmol K)
C2 universal constant in Eq. (14) (g2 J/mol3 K) s0 molar entropy relative to 0 K (standard entropy, abso-
D1 universal constant in Eq. (14) (g J/mol2 K) lute entropy) (kJ/kmol K)
D2 universal constant in Eq. (14) (g2 J/mol3 K) T absolute temperature (K)
c amount of carbon in Eq. (16) (kmol/kg (daf)) T0 absolute surroundings temperature (K)
cP specific heat capacity at constant pressure (J/g K) T exhaust exhaust gas (flue gas) temperature (K)
cP molar heat capacity at constant pressure (J/mol K) T in absolute temperature at inlet (K)
Eproduced sum of produced electricity and heat according to the T out absolute temperature at outlet (K)
R1-formula (MW h) X exergy (MJ)
Efuel chemical energy of fuel (MW h) X chem chemical exergy (MJ)
Eimported imported energy (MW h) X expended expended exergy (MW h)
Ewaste chemical energy of waste (MW h) X recovered recovered exergy (MW h)
H weight percent hydrogen on dry basis (wt% dry basis) X waste chemical exergy of waste (MJ)
HHV higher heating value (MJ/kg) x specific exergy (MJ/kg)
h amount of hydrogen in Eq. (16) (kmol/kg (daf)) xchem specific chemical exergy (MJ/kg)

h molar enthalpy (kJ/kmol) xwaste specific chemical exergy of waste (MJ/kg)
h
h 0 sensible enthalpy relative to the standard reference w weight percent moisture (–)
state (kJ/kmol) W el electric energy (MW h)
0
h molar enthalpy of formation (kJ/kmol)
f
LHV lower heating value (MJ/kg) Greek symbols
M molar mass (g/mol) a similarity variable (mol/g)
m mass (kg) b factor (–)
N weight percent nitrogen on dry basis (wt% dry basis) eu utilization factor (–)
N amount of substance (kmol) gII second-law-efficiency, exergy efficiency (–)
Nr amount of substance of a reactant (kmol) h universal constant in Eq. (14) (K)
Np amount of substance of a product (kmol) t number of moles (–)
n amount of nitrogen in Eq. (16) (kmol/kg (daf))
O weight percent oxygen on dry basis (wt% dry basis)
Abbreviations
o amount of oxygen in Eq. (16) (kmol/kg (daf)) ar as received
Q heat (MW h) CHP combined heat and power
Q combustion heat released in steady combustion process (MJ) daf dry and ash free basis
Q district district heat (MW h)
dry dry basis
Q process process heat (MW h)
R1 efficiency as defined by the R1-equation (–)

transformation of waste into useful energy with large numbers the evaluation of the up-and-running system and proposals for its
(above 50% at best) that on the other hand may not have a clearly improvements [12]. Ray et al. [13] give an example of a situation in
defined root in the language of thermodynamics. It may also be which second law analysis is better suited to identify the contribu-
difficult to compare efficiencies between different plant operators tions of single components to the overall efficiency of a power
if the respective definitions are not clearly documented and may plant than first law analysis. Padilla et al. [14] demonstrate the
differ from each other. comparison of the suitability of different working fluids for a
In 2010, Grosso et al. [4] published a study, which related the domestic refrigerator based on second law analysis.
R1-formula to exergy efficiencies of different types of waste incin- Based on the same energy input, a system with higher exergy
eration plants (heat producing, power producing, and combined efficiency is expected to sell a larger amount of useful energy
heat and power (CHP) plants). An average value of 20% exergy effi- and thereby have a potential for larger revenue. It is thus impor-
ciency was reported for Norway and an average energy recovery tant to focus on optimal second law efficiency in addition to cost
efficiency of 0.71 according to the R1-formula. The aim of this effectiveness already in the planning stage of new facilities [15].
investigation was to verify these values based on a case study of The method of second law analysis is used in this study in order
a CHP-plant located in Bergen, Norway. Such a verification of first to evaluate the efficiency of the conversion of chemical exergy of
and second law efficiencies is an important part in the process of waste into district and process heat as well as electricity. Earlier
operating and optimizing this type combined heat and power studies have identified the combustion process in electrical power
plants. generation as the main source of irreversibilities, where about one
Second law analysis (or exergy analysis) has been pointed out as third of the exergy of a fuel is lost [16] already before heat transfer
a valuable and important tool in evaluating efficient use of energy to the working fluid.
[10]. According to Lior [11], the analysis can be applied in various In order to calculate the heating value and chemical exergy of
phases of an energy conversion project: the feasibility analysis, the the combustible waste in this study its chemical composition
determination of the system’s components and working fluids and needs to be known. For the year 2012, the combustible waste from
T. Solheimslid et al. / Energy Conversion and Management 95 (2015) 149–159 151

the Bergen region was first characterized by Husetuft Ommedal 2. Methods


et al. [17]. They compared its heating value with and without the
wet organic fraction. The composition used in this study was A short overview of the methods used in this study is given in
obtained by Harneshaug and Solheimslid [18] and is based on a the following sections, starting with some details of the waste
more detailed analysis of the incinerated waste. Different incineration facility in this case study, the amount and composition
approaches for calculating the chemical exergy of biomass exist. of the incinerated waste and finally details of the calculations of
One can choose between different correlations for directly obtain- the different efficiencies.
ing the chemical exergy based on the weight percentages of the
main elements contributing to the energy release in the combus-
tion process (typically carbon, hydrogen, oxygen, and often nitro- 2.1. About the CHP waste incineration plant in Bergen, Norway
gen, sulphur, and the ash content in addition on dry basis).
Examples for different correlations are given in Section 2.7. The waste incineration plant under investigation is located in
Another approach to calculating chemical exergy is to combine Bergen, on the Norwegian West coast, at 60° northern latitude.
heating value and changes in entropy of waste, combustion air The plant has two combustion chambers and steam generators
and combustion products multiplied by the surroundings which produce steam for driving two turbines in order to produce
temperature to a value representing the available energy from electricity. The temperature in the combustion chambers usually
waste by combustion [8]. For the latter method, the absolute lies around 1000 °C. Part of the generated steam is used internally
entropy of the waste has to be known which may be difficult to as process heat. Two heat exchangers transfer heat from the
obtain as waste is composed by all sorts of materials. Standard expanded steam to the district heating network, which has an out-
entropy values are not as well documented in the literature as going temperature of 120 °C in winter and 110 °C in summer
heating values, for example. Once again, correlations for either or respectively. The return temperature of the water in the district
both heating value and absolute entropy can be used if the average heating network is about 60 °C. The waste incineration plant has
chemical composition of the waste is known [19]. In most cases, an annual capacity of 210,000 tons and both household waste
correlations are limited to a certain type of fuel and one has to and waste from the local industries and recycling stations is
be careful when applying a correlation originally developed for, received. Fig. 1 shows a simplified sketch of the facility and the dif-
for example, coal or fuel oil, to biomass, which has a different ferent mass flows between the main components. Table 1 gives an
chemical composition [20]. overview over the delivered electricity and district/process heat in
A different approach for calculating the absolute entropy of the 2012. Around 22,000 households can be supplied with district
waste was tested in this study. The waste was assumed to behave
as a solid, incompressible material. Its specific heat capacity was
integrated as function of temperature from 0 K to 298.15 K. The Table 1
local surroundings temperature of 279.95 K (6.8 °C; [21]) for the Overview over produced electricity (Wel) and delivered district/process heat (Qdistrict,
Qprocess) as well as imported energy contributions (Efuel, Eimported) relevant for the R1-
CHP plant in this case study was also used as upper temperature
formula. The produced electricity has to be multiplied by a factor of 2.6 in the R1-
limit in the integral. This was done in order to investigate the effect formula, and heat by 1.1.
of its value on the second-law-efficiency compared to the standard
Quantity Amount per year (MW h) Type
value of 25 °C. The employed function for the specific heat capacity
developed by Laštovka et al. [22] is based solely on the chemical W el 85,000 Delivered
Q district 214,000 Delivered
composition and can be used for so-called ill-defined hydrocarbons
Q process 50,000 Delivered
containing up to a certain amount of heteroatoms. The results from Efuel 2210 Imported
this approach are in good agreement with other methods as will be Eimported 1006 Imported
shown in the following sections.

Fig. 1. Simplified sketch of the CHP-facility in this study. The different line types represent the different mass flows: working fluid (water, solid lines), air and flue gas (short
dashed lines), district heat working fluid (water, long dashed lines), process heat working fluid (water, dash-dot-dotted lines), waste and bottom ash (dash-dotted lines). For
simplicity only one incinerator/boiler and steam turbine is shown instead of two.
152 T. Solheimslid et al. / Energy Conversion and Management 95 (2015) 149–159

heating and the produced electricity supplies up to 5000 Table 3


households. Amount of the most important elements characterizing the composition of the waste
and the amount of ash in weight percentage on dry basis.

2.2. Waste composition Carbon Hydrogen Oxygen Nitrogen Sulphur Ash


52.2 7.32 27.9 1.02 0.25 11.4
A detailed analysis has given the following properties of the
waste input to the plant [18]. In order to estimate the chemical
composition, dry mass, dry and ash free (daf) mass and the Wel stands for the produced electricity. The two factors 1.1 and 2.6
corresponding heating values, the Phyllis2 database [23] provided in Eproduced are called equivalence factors. They relate the process
by the Energy Research Centre of the Netherlands (ECN) was heat and electricity produced in a waste incineration facility to
employed because there is no such database available for proper- the average efficiency of European coal-fired power and process
ties of Norwegian waste. 195,000 tons were received by the instal- heat plants. Due to this, the R1-formula is not exactly based on
lation in 2012 of which 169,876 tons were combustible (see the first law like the utilization factor. It may yield values larger
Table 2). On dry basis, the combustible mass was 139,806 tons in than one. R1-values are therefore usually not given in percent.
2012. The dry and ash free mass fed into the incineration process Ewaste is the energy available from waste that goes to incineration
was 123,870 tons. The weight fractions of the most important after pre-treatment and is basically the product of net calorific val-
elements (carbon, hydrogen, oxygen, nitrogen, and sulphur) and ue (lower heating value) and the respective mass on an annual
the ash content are given in Table 3 on dry basis. The waste com- basis. The other two contributions have to be carefully chosen
position can be represented by means of a fictive molecule with according to the R1-guidelines [25] which specify the boundary of
the chemical composition C6H10O2.41N0.10S0.01 like it is done in the system that is to be characterized by the R1-formula. Roughly
other publications [24]. In the calculations, the waste is assumed explained Efuel stands for fuels that are not waste but contribute
to be solid at the surroundings temperature T0 which is important to the generation of steam. Eimported represents amongst others elec-
to note with respect to the calculation of its absolute entropy and tricity and other kinds of non-fuel energy that are used in start-up
chemical exergy as well as the recovered exergy in heat exchang- and shutdown processes, when the facility is not delivering energy
ing processes. across the system boundary, or imported energy used in other parts
of the installation such as the flue gas cleaning system [25].
2.3. Efficiencies The second law efficiency is simply defined as [8]

Efficiencies related to both first and second law of thermody- X recovered


gII ¼ ð4Þ
namics were calculated. X expended
The first one was the utilization factor. Based on the first law of
X expended is equal to the chemical exergy of the waste that is deliv-
thermodynamics it is defined for a combined heat and power plant
ered to the oven and incinerated over the course of one year. The
as [8]
recovered exergy is in case of the studied CHP-plant a sum of three
net work output þ process=district heat delivered contributions:
eu ¼ ð1Þ
total heat input  
T0
The heat input was set to the product of the waste’s dry and ash free X recovered ¼ W el þ Q district  1  þ Q process
T district
(daf) mass per year and the waste’s respective higher heating value  
T0
(daf). The net work output was set to the electricity delivered  1 ð5Þ
T process
(including the internal use in the plant). The plant delivers both dis-
trict heating to the communal network and some process heat is The first two terms on the right hand side are the exergy of the
used internally in the plant for other processes. delivered electricity and district heat to customers outside the facil-
The second energy efficiency was calculated by means of the ity. The third term represents the exergy of the process heat con-
R1-formula published by the European Union [9]. The aim of this sumed internally in the facility. The surroundings temperature T0
formula is to classify waste treatment operations by incineration is set to the annual temperature average in Bergen in 2012, which
as either disposal or energy recovery operations based on thresh- was 279.95 K (6.8 °C) [21]. Following the work of Grosso et al. [4]
old values of 0.6 and 0.65 for facilities that started operation before the other two temperatures (Tdistrict and Tprocess) were set to the
or after the year 2009 respectively. logarithmic mean of the temperatures at the inflow and outflow
Eproduced  ðEfuel þ Eimported Þ of the heat exchangers on the facility side of the system:
R1 ¼ ð2Þ
0:97  ðEwaste þ Efuel Þ T out  T inn
T¼   ð6Þ
The factor 0.97 in the denominator is meant to take care of heat ln TT out
inn
losses due to radiation and bottom ash. The term Eproduced is
calculated by With Tin and Tout equal to 402 K and 368 K respectively for the heat
exchanger to the district heating network one gets Tdistrict = 385 K
Eproduced ¼ 1:1  ðQ district þ Q process Þ þ 2:6  W el ð3Þ
(112 °C). Tprocess = 402 K (128 °C) as Tin and Tout are equal to respec-
The terms Qdistrict and Qprocess represent heat delivered to the tively 443 K and 363 K for the process heat exchanger. All these
district heat and facility internal process heat networks. The symbol temperature values were rounded to three significant digits.

Table 2
Overview over the amount of combustible waste received by the waste incineration plant and the combustible mass on dry and dry and ash free (daf) basis.

Combustible mass (ar) [tons] Moisture content Combustible mass (dry) [tons] Ash content Combustible mass (daf) [tons]
wt% (ar) [tons] wt% (dry) [tons]
169,876 17.7 30,070 139,806 11.4 15,936 123,870
T. Solheimslid et al. / Energy Conversion and Management 95 (2015) 149–159 153

Only the R1-formula takes energy needed for start-up and shut- and ash free basis (daf) by means of Eq. (8) in order to compare
down operations into account of these three different efficiencies. them to the values obtained from the Phyllis2 database [23].
The original textbook definitions of second-law-efficiency and uti-
lization factor were kept which contain only contributions from 2.5. Calculation of the chemical exergy from combustion equation and
steady ‘production’ operations. However, it would not be difficult absolute entropy
to include necessary energy for start-up and shutdown into these
as well if they represent a significant part of the involved energy The expended exergy X expended to be used in the calculation of
input into the system. the second law efficiency is the chemical exergy of the waste avail-
able per year plus eventual other contributions from other energy
2.4. Heating values sources. In the present case, there are no other energy sources con-
tributing to the output of the facility other than the waste. The
As mentioned before the Dutch Phyllis2 database [23] was available work potential from a chemical reaction, Xchem, can be
employed, mainly to obtain the average chemical composition of calculated from the following formula if kinetic and potential ener-
the waste by means of carbon, hydrogen, oxygen, sulphur, nitro- gies are neglected and T0 is the surroundings temperature [8]:
gen, moisture, and ash content. Heating values (both higher and X  
X chem ¼ 0 þ ðh
Nr h h0 Þ  T 0s0
lower) are given as an experimental value and also as result from f
r
a correlation by Milne ([26], originally published in [27]), based X  
  0  
Np hf þ ðh  h Þ  T 0 s0
0  ð10Þ
on the chemical composition of the waste on dry and ash free basis. p
In order to verify the validity of both these values the composition
The indices r and p represent reactants and products respectively. N
data on a dry basis was used to calculate the heating value by
is the amount of substance, h 0 the molar enthalpy of formation at
means of the recently published correlations by Komilis et al. f

[28] and Song et al. [19]. When using the database and different h
the reference state (25 °C, 1 atm), h 0 the sensible enthalpy rela-
correlations it is important to use the correct masses and weight tive to the standard reference state and s0 the absolute molar
percentages, for example. While the Phyllis2 database [23] shows entropy. This equation may be rearranged to
waste properties for dry and ash-free (daf) material, many correla-  
tions expect input values and on a dry basis (db) only and heating X chem ¼ Q combustion  T 0  Swaste þ SO2  SCO2  SH2 O  SSO2 ð11Þ
values are also returned on dry basis. It is possible to convert
Stoichiometric combustion was assumed with CO2, H2O, and SO2
between those values using the following equations [26] where
being the products along with N2 from the dry air entering on the
the subscript ar stands for ‘as received’:
reactant side. The heat Qcombustion set free in this process is then
  simply the higher heating value (HHV, H2O is assumed to be in
wt% moisture
HHVar ¼ HHVdry  1  ð7Þ the liquid phase in the products) of the waste multiplied with the
100
  waste mass for one year. The absolute entropies S ¼ Ns0 for oxygen,
wt% ash
HHVdry ¼ HHVdaf  1  ð8Þ carbon dioxide, water and sulphur dioxide were interpolated from
100 the literature [8] for the surroundings temperature of T0 = 6.8 °C
(see Table 5) except for SO2, where only the value for 25 °C was
Table 4 shows a comparison of the different heating values obtained
available to the authors. This had, however, little to no effect on
from both the database, Milne-, Song- and Komilis-correlations for
the results as the amount of sulphur is very low in the waste com-
the waste composition in the studied case. It was decided to use
pared to the main elements carbon, hydrogen, and oxygen. Only a
the experimental values provided by the database (given as
negligible amount of SO2 turns up in the products. The total entropy
HHVPhyllis in the tables) and mainly employ the correlation used
value for one substance is calculated from the product of absolute
by Song et al. [19] for comparisons of the exergy and exergy effi-
molar entropy and the respective number of moles based on the
ciency calculations. This choice was made because the heating value
balanced combustion equation for the waste. Based on the assump-
calculated with their formula was closest to the experimentally
tion of stoichiometric combustion with dry air one obtains the fol-
obtained higher heating value in the database (see Table 4). The cor-
lowing balanced combustion equation
relation employed by Song et al. [19] was originally published by
Channiwala and Parikh’ [29] and reads C6 H10 O2:41 N0:1 S0:01 þ 7:305ðO2 þ 3:76N2 Þ
! 6CO2 þ 5H2 O þ 0:01SO2 þ 27:52N2 ð12Þ
HHV ¼ 0:3491  C þ 1:1783  H  0:1034  O  0:0151  N
þ 0:1005  S  0:0211  A ð9Þ This combustion reaction is reversible. However, the equilibrium at
the combustion temperature is assumed to be on the product side
It yields the HHV in MJ/kg where C, H, O, N, S, and A are the weight as is typical for fuels with a high amount of carbon and hydrogen
fractions of carbon, hydrogen, oxygen, nitrogen, sulphur and ash on and carbon dioxide and water in the products. The absolute molar
dry basis (db). The heating values in Table 4 were converted to dry entropy of the waste is dependent on the composition of the waste
and has to be calculated separately. This can be achieved, for

Table 4
Heating values for dry and ash free (daf) material. The first three heating values were Table 5
obtained from the Phyllis2 database [23]. The heating value calculated by the Absolute molar entropies for O2 (g), CO2 (g), H2O (l) and the waste C6H10.0O2.41N0.1-
correlation of Komilis et al. [28] is based only on carbon, hydrogen and oxygen S0.01(s) at 6.8 °C and 25 °C. The value for SO2 is for 25 °C in both cases.
content while the HHV calculated with the correlation by Channiwala and Parikh’
(Ch.P.; [29]) employed by Song et al. [19] also takes the sulphur and nitrogen content Molecule Absolute molar entropy Absolute molar entropy
of the combustible mass into account. The latter two heating values were converted [J/(mol K)] at 6.8 °C [J/(mol K)] at 25 °C
from dry basis to dry and ash free basis by means of Eq. (8). O2 (g) 203.19 205.04
CO2 (g) 211.49 213.80
LHVPhyllis HHVPhyllis HHVMilne HHVCh.P. HHVKomilis
H2O (l) 65.15 69.92
24.99 MJ/kg 26.71 MJ/kg 26.91 MJ/kg 26.77 MJ/kg 25.95 MJ/kg SO2 (g) 248.22 248.22
Experimental Experimental Correlation Correlation Correlation C6H10.0O2.41N0.1S0.01(s) 160.82 170.56
154 T. Solheimslid et al. / Energy Conversion and Management 95 (2015) 149–159

Pn
example, by means of semi-empirical correlations or the method m
i¼1 i
presented in the next subsection.
a ¼ Pn ð15Þ
m  Mi
i¼1 i

In this equation, the sum goes over the n elements the molecule
2.6. Calculation of absolute entropy
consists of and mi and Mi are number of moles and molar mass of
element i respectively. It can easily be shown that the value of a
Access to reliable values for heating values and absolute
is independent of choice of the base of carbon atoms in the virtual
entropy of biomass of arbitrary composition is important in calcu-
molecule, representing the composition of the waste. Laštovka et al.
lating the chemical exergy of (bio) fuels. While heating values of
report their model [22] to be of similar and, for certain types of sub-
many types of fuels are well documented and many different cor-
stances, of higher precision than the Goodman et al. [30] power-law
relation functions exist, reference values for absolute entropy are
equation. The extra material provided with this article contains a
more difficult to obtain. An approach for calculating the solid
more detailed comparison of these two approaches to calculating
waste’s absolute entropy based on its dry base composition was
the molar heat capacity.
chosen that, to the author’s knowledge, has not been tried before
It was necessary to numerically integrate the integral in the
in this context. Laštovka et al. [22] have developed a similarity
equation for the absolute entropy (Eq. (13)). The reason for this
variable for estimating the heat capacity of ill-defined organic
is several terms with division by absolute temperature T in Eq.
solids. As the composition of municipal waste is usually character-
(14) and thereby a division by zero at the lower boundary of the
ized by a fictive molecule, no information about its possible chemi-
integral. The freely available computer algebra system Maxima
cal structure is available. It would be hard to choose one that
[31] was used to carry out the integration.
would be representative for the average chemical and physical
Two correlations for the absolute entropy of solid fuels were
properties of the waste, when incinerated.
chosen in order to compare the results for the entropy of the solid
The model by Laštovka et al. [22] circumvents this problem by
waste obtained by the approach described above to results obtain-
correlating the specific heat capacity at constant pressure of many
able with other methods. The first correlation is given in the book
molecules with like composition but different structures to each
of Bejan et al. [20]. Even though it was developed for coal, char and
other. Over 2000 data points of 72 organic solids were used in
fuel oils authors of other publications used it to calculate the molar
the derivation of the parameters of this model. The authors could
entropy of biomass at normal temperature (25 °C) (for example
show that for solid organic molecules with a limit on the amount
[32]). The equation is
of heteroatoms, the specific heat capacity is insensitive to the
  
structure over a wide range of temperatures. The model is accurate h
s0Bejan ¼ c  37:1653  31:4767 exp 0:564682 
between 50 K and the fusion temperature. In this study, it was suc- cþn
cessfully employed to integrate the following equation from 0 K to 
o n s
the surroundings temperature T0: þ 20:1145  þ 54:3111  þ 44:6712 
cþn cþn cþn
Z T0 ð16Þ
cP ðTÞ
s0 ¼ dT ð13Þ
0K T and c, h, o, n, and s have to be given in kmol/kg dry and ash free
The waste is assumed to be solid and incompressible and therefore material. The molar entropy has the unit kJ/kmol K.
the pressure or volume related terms in calculation of the entropy The other correlation used for comparison was published by
can be neglected [8]. Song et al. [19] as part of their work on a correlation for the chemi-
According to Laštovka et al. [22] the variation of the molar heat cal exergy of solid and liquid biomass. It reads
capacity of solid organic compounds (in J mol1 K1) with tem-
s0Song ¼ 0:0086  C þ 0:0780  H þ 0:0106  O þ 0:0103  N
perature is
 2  þ 0:0118  S ð17Þ
  h exp Th
cP ðTÞ ¼ 3 A1 a þ A2 a2 Ru   2
T exp Th  1 which returns the entropy in units of kJ/kg K at 25 °C and
   2 where C, H, O, N, and S are the contents of the elements carbon,
T T
þ ðC 1 a þ C 2 a2 Þ þ ðD1 a þ D2 a2 Þ ð14Þ hydrogen, oxygen, nitrogen, and sulphur in wt% dry material. It is
K K based on 162 standard entropy values of organic solids and liquids
In this equation T is the absolute thermodynamic temperature in at 25 °C.
Kelvin, Ru the universal gas constant, and A1, A2, C1, C2, D1, D2, and
h are universal constants of this correlation (see Table 6 for values 2.7. Calculation of chemical exergy of solid biomass by means of
and units). The letter ‘K’ in some of the factors in parentheses stands correlations
for the SI-unit Kelvin and leads to the correct unit for the molar heat
capacity, cP ðTÞ. The factor a is a similarity variable that is calculated The alternative to calculation of the chemical exergy by means
from the composition of the molecule of interest by: of heating values and entropies are correlation functions. They give
the chemical exergy of a fuel based on its chemical composition on
either dry or dry and ash free basis. Three different correlations
Table 6 were used to calculate the chemical exergy of the waste in this
Factors in the equation of molar heat capacity of organic solids by Laštovka et al. [39]. study.
Parameter Value Unit Song et al. [19] have published a unified correlation for calcula-
A1 0.013183 g/mol
tion of the specific chemical exergy of solid biomass on dry basis.
A2 0.249381 (g/mol)2 The chemical exergy of the inorganic part is ignored as it could
C1 0.026526 g J/(mol2 K) be shown to have negligible impact on the results. Chemical exergy
C2 0.024942 g2 J/(mol3 K) in units of kJ/kg is then calculated from
D1 0.000025 g J/(mol2 K)
D2 0.000123 g2 J/(mol3 K) xchem;dry ¼ 362:008  C þ 1101:841  H  86:218  O þ 2:418
h 151.8675 K
Ru 8.314472 J/(mol K)  N þ 196:701  S  21:1  A ð18Þ
T. Solheimslid et al. / Energy Conversion and Management 95 (2015) 149–159 155

where C, H, O, N, S, and A represent weight percent dry basis [22] the value of the similarity variable a is 0.151251 at a molar
(wt% db) of carbon, hydrogen, oxygen, nitrogen, sulphur and ash mass of 122.47 g/mol. At T0 = 6.8 °C the absolute molar entropy is
contents of the fuel. This correlation is applicable to various 160.82 kJ/kmol K and it is 170.56 kJ/kmol K at 25 °C. The correla-
types of fuels (biomass, coal, petroleum, for example) with compo- tion of Song et al. [19] (Eq. (17)) gives a value of 162.74 kJ/kmol K
sitions in the range 27.33% 6 C 6 89.10%, 2.46% 6 H 6 14.00%, 1.1% 6 at 25 °C and the one by Bejan et al. [20] (Eq. (16)) 199.41 kJ/kmol K.
O 6 46.92%, 0.00% 6 N 6 9.27%, 0.01% 6 S 6 5.54%, and 0.00% 6 A 6 The three contributions in recovered exergy (see Eq. (5)) sum up
51.96%. to X recovered = 158 GW h at 6.8 °C surroundings temperature and
A correlation for the chemical exergy of biomass originally pub- 146 GW h at 25 °C surroundings temperature. The difference in
lished by Szargut and Styrylska [33] and used by many authors, for surroundings temperature thus leads to a difference of about 8%
example Prins et al. [34], reads in the amount of recovered exergy.
Different values for the second-law-efficiency gII of the energy
xbiomass ¼ b  LHVbiomass ð19Þ
conversion process were obtained depending on the method for
where the factor b is calculated by calculating the exergy of the incinerated waste and the choice of
 surroundings temperature. See Table 7 for detailed results.
1:0412 þ 0:2160 HC2  0:2499 OC2 1 þ 0:7884 HC2 þ 0:0450 NC2 The first set of results in Table 7 is based on using heating value

1  0:3035 OC2 and absolute entropy for calculating the exergy of the waste where
ð20Þ three different sources for the higher heating value have been
used: values from the Phyllis2 database [23], and the correlations
The element/molecule symbols represent the weight fractions of by Channiwala and Parikh’ [29] and Komilis et al. [28] presented
carbon, hydrogen, oxygen, and nitrogen and the ratio of O2/C must in Section 2.4. The absolute entropy of the waste was calculated
be less than or equal to 2.67. based on either the integration of specific heat capacity [22] or
A third correlation used for comparison is the one by Kotas [35] the correlations by Song et al. [19] or Bejan et al. [20].
and reads The second law efficiency at 6.8 °C surroundings temperature is
17.3% for the given chemical composition of the waste and mois-
xbiomass ¼ ðLHVbiomass þ 2442  wÞ ture content. The combination of HHV from the Phyllis2 database
 
H O N [23] and entropy from the Laštovka et al. [22] method was used.
 1:0437 þ 0:1882  0:0610 þ 0:0404 ð21Þ
C C C At 25 °C surroundings temperature the second law efficiency is
15.9%. This is mainly due to the lower recovered exergy in process
which is valid for O/C < 0.667. The element symbols represent and district heat at this temperature while the difference in speci-
weight percent on dry basis, w is weight percent moisture and fic chemical exergy of the waste is negligible. The same results are
the unit of xbiomass is kJ/kg. obtained, when the heating value is replaced with values calculat-
ed with the correlation by Channiwala and Parikh’ [29] which was
3. Results used by Song et al. [19] to derive their correlation for the chemical

Based on the waste composition analysis, the thermal proper-


ties of the waste and the energy conversion processes in the waste
Table 7
incineration installation the following results for utilization factor,
Overview over the chemical exergy of the waste (Xwaste) in absolute (GW h) and in
R1-efficiency, absolute entropy, and second-law-efficiency were specific form (MJ/kg, on dry basis) as well as the resulting second law efficiency from
obtained. different ways of calculation. The first three columns give the references from which
When using the product of dry and ash free mass and the cor- values for either higher heating value (HHV) and absolute entropy (S) or the chemical
exergy (Xchem) were taken for calculating the exergy of the waste, Xwaste.
responding higher heating value from the Phyllis2 database [23]
(HHVPhyllis from Table 4) as heat input in Eq. (1) for eu one obtains References T0 Xwaste xwaste gII
a utilization factor of 38.0% and 40.6% with the corresponding low- HHV S Xchem °C GWh MJ/kg (%)
er heating value (LHVPhyllis) from the same table and source. A (db)
much higher value of energy utilization is obtained if one only Phyllis2 [23] Laštovka 6.8 915 23.57 17.3
takes the thermal energy into account in the denominator of eu that et al. [22]
is actually transferred into the steam cycle in the heat exchanging Laštovka 25 916 23.60 15.9
process from combustion chamber to working fluid (locally called et al. [22]
Song et al. 25 917 23.61 15.9
the ‘Norwegian method’). This amount of energy is given by
[19]
486,000 MW h by the plant operator and yields 71.8% energy Bejan et al. 25 914 23.53 16.0
utilization. [20]
Regarding the R1-efficiency, certain energy amounts with Channiwala & Laštovka 25 918 23.65 15.9
respect to start-up and shutdown-processes of the waste-incin- Parikh‘ [29] et al. [22]
eration plant have to be taken into account as well. These were Song et al. 25 919 23.66 15.9
[19]
Efuel = 2210 MW h and Eimported = 1006 MW h in 2012 according to
Bejan et al. 25 915 23.56 16.0
the plant operator. The energy input from the waste has to be given [20]
based on the waste’s lower heating value, which is 24.99 MJ/
Komilis et al. [28] Laštovka 25 890 22.91 16.4
kg (daf) (LHVPhyllis from Table 4). This gives 860 GW h for Ewaste . et al. [22]
The ‘produced’ energy term has a value of 475 GW h. The result Song et al. 25 890 22.92 16.4
is a R1-efficiency of 0.568, which is short of being a recovery opera- [19]
tion according to the EU directive [9] as the facility in its current Bejan et al. 25 887 22.83 16.5
[20]
setup has been in operation since the beginning of 2010. Song et al. 25 946 24.35 15.4
The different approaches to the calculation of the absolute [19]
entropy of the waste with composition C6H10O2.41N0.1S0.01 gave Prins et al. 25 897 23.09 16.3
results that differed up to 19% from each other. In case of the [34]
Kotas [35] 25 1025 26.39 14.2
method based on the heat capacity equation by Laštovka et al.
156 T. Solheimslid et al. / Energy Conversion and Management 95 (2015) 149–159

exergy of solid biomass. The value of the specific chemical exergy is hydrocarbons and a hydrocarbon with heteroatoms based on the
ca. 23.6 MJ/kg on dry basis, which corresponds to 916 GW h rever- different approaches presented in Section 2.6 were compared to
sible work potential for incinerated waste in 2012 in the studied literature values. These substances were Octadecane (C18H38),
facility. When the correlation by Komilis et al. [28] is employed Eicosane (C20H42), and Dotriocontane (C32H66) each in their linear
to find the higher heating value, smaller values for the waste’s conformation and 1,6-Hexanediol (C6H14O2). The latter is similar
specific chemical exergy are the result. This is due to the smaller in molecular composition and molar mass compared to the compo-
heating values returned by this correlation compared to both Phyl- sition of the waste in this study. The literature values of the select-
lis2 database [23] and Channiwala & Parikh’ correlation [29]. As a ed substances were taken from the NIST databases [36] except for
consequence, larger values for the second law efficiency of 16.4% 1,6-Hexanediol, which was taken from Smirnova et al. [37]. The
and 16.5% are obtained with the different approaches to the calcu- choice of the three pure hydrocarbons was made with respect to
lation of absolute entropy. Both absolute entropy and the recov- the limits for use of the Laštovka et al. [22] approach. It works best
ered exergy are unaffected by the value for the heating value. for hydrocarbons with M > 200 g/mol and with less than 15 wt%
The last comparison in Table 7 is between the results from dif- heteroatoms. As mentioned above, C6H10O2.41N0.1S0.01 is the chemi-
ferent correlations for specific chemical exergy. The results range cal composition of the waste in this study. The weight percentage
from 23.09 MJ/kg to 26.39 MJ/kg for the chemical exergy and of heteroatoms is 32.9 wt%. With the exception of Eicosane, the
16.0–14.2% for the second law efficiency respectively. The results Song et al. correlation [19] had the smallest error, less than 1%.
from the Song et al. correlation [19] are 24.35 MJ/kg and 16.0% The results closest to the literature value of Eicosane came from
and closest to the results obtained with the before mentioned the Laštovka et al. [22] approach with 0.4% deviation. The literature
methods. The Prins et al. correlation [34] gives similar results value of 1,6-hexanediol at 25 °C (232.1 kJ/kmol K, [37]) was best
(23.09 MJ/kg and 16.3%) while the values from the correlation approximated by the Song et al. correlation [19] while the other
employed by Kotas [35] (26.39 MJ/kg and 14.2%) deviate by about two approaches gave large errors of 13.5% (Laštovka et al. [22])
11% from the typical 23.6 MJ/kg and 15.9%. and 8.4% (Bejan et al. [20]). The large error of the Laštovka et al.
[22] method may be due to the low molar mass of just
118.17 g/mol. This is much lower than the recommended lower
4. Discussion limit of 200 g/mol. See the extra material provided with this article
for a more detailed comparison between the different methods for
The utilization factor of just around 40% is put into perspective the case of 1,6-Hexanediol.
by the second-law-efficiency of 17.3%. The largest part of the uti- As the Song et al. [19] approach had the best approximation of
lized energy from the combustion of waste is turned into dis- the available correlations to the absolute entropy of 1,6-Hexanedi-
trict/process heat, which has a much lower work potential than ol it was assumed that it has a similar low deviation from the abso-
the delivered electricity. This is due to the Carnot-efficiencies that lute entropy of the fictive waste molecule as well. It therefore
the heat delivered to process heat network and district heat net- served as reference for the calculation of the deviation of the values
work has to be multiplied with in Eq. (5) in order to obtain the given by the other two methods. The value obtained from the
respective exergy. The Carnot-efficiencies based on the tem- Laštovka et al. [22] method is 8.8% larger and the Bejan et al.
peratures in the heat exchangers (see Section 2.3 and Eq. (6)) are [20] correlation gives a 23% larger value for the absolute molar
30.3% and 27.2% at 6.8 °C surroundings temperature and 25.7% entropy of the waste.
and 22.5% at 25 °C respectively. Heat transfer through finite tem- The choice of number of carbon atoms in the molecule formula
perature differences in the heat exchangers at relatively low tem- of the waste is arbitrary. A number of 8 or 10 carbon atoms could
peratures are therefore one of the main causes of exergy have been chosen as well. The number of hydrogen and oxygen
destruction and relatively low exergy recovery. However, exergy atoms would have been adjusted accordingly, resulting in a larger
recovery and second law efficiency would be even lower without molar mass. In this way, it would have been easy to satisfy the con-
any utilization of waste heat from the steam turbines as district straint for the molar mass. The most important thing to note is
or process heat. however, that independent of the choice of the number of carbon
Grosso et al. [4] related the R1-values to second-law-efficiencies atoms in the representative waste molecule, all molecules would
for European waste incinerations plants. For the Norwegian plants, have the same value for the similarity value a, which is the only
an average second-law-efficiency of 20% was predicted. Another property of the molecule that finally enters the equation for the
result in this study relates the average R1-efficiency of combined specific heat capacity as function of absolute temperature (Eq.
heat and power plants of 0.71 to a second-law-efficiency of 20.9% (14)). Consequently, it does not matter, that the fictive waste mole-
while a combination of R1 = 0.568 and 17.3% second-law-efficiency cule of choice has a much lower molar mass than recommended.
at the local surroundings temperature of 6.8 °C was found in this Laštovka et al. have used solid substances with molar masses down
study. This is characteristic for a mainly heat producing plant, to 124.22 g/mol and weight fractions heteroatom of up to 32 wt%
according to Grosso et al. [4] for which these authors have found in their training sets in their two studies published in 2008
a combination of 0.64 for the R1-value and 18.8% exergy efficiency. [38,39]. This gives the authors reason to believe, that the results
A comparison of the different heating values (Table 4) shows, for the absolute entropy of the waste using the method of Laštovka
that the correlation of Channiwala and Parikh’ [29] gives the value et al. [22] are reasonable despite the observed large deviations in
that comes closest to the experimental result from the Phyllis2 case of 1,6-Hexanediol.
database [23] for the given chemical composition of the waste with The different values for the higher heating value have the stron-
a deviation of only 0.2%. The values from the correlations by Milne gest influence on the result for the second-law-efficiency. This can
[26] and Komilis et al. [28] differ by 0.7% and 2.8% from the easily be seen in Table 7. When the entropy value based on the
experimental value. Laštovka et al. method [22] was replaced with the result from
One of the different methods to obtain the absolute entropy of the entropy correlation from Song et al. [19], the second-law-effi-
the waste was the integration of the specific heat capacity divided ciency was still 15.9% with 25 °C as surroundings temperature.
by absolute temperature with respect to temperature from abso- Almost the same result (16.0%) was obtained when employing
lute zero to the annual average of the surroundings temperature the correlation given by Bejan et al. [20]. This is a bit surprising
of the facility or the standard reference temperature of 25 °C (Eq. because the Bejan et al. correlation [20] was developed for coal,
(13)). Results obtained for the absolute entropy of three selected char and fuel oil. On the other hand, the contribution from the
T. Solheimslid et al. / Energy Conversion and Management 95 (2015) 149–159 157

factor T0DS in the approach where it is subtracted from the heat- T exhaust = 180 °C and the local surroundings temperature T0. Water
ing value is typically two orders of magnitude smaller than the is treated as superheated vapour at Texhaust and compressed liquid
exergy value itself. Therefore, the much larger entropy value com- at T0 with absolute pressure set to 100 kPa for simplicity in both
pared to the selected reference value does not have a large effect states. For all substances tabulated values for enthalpy and entropy
on the exergy value and second law efficiency. from the book of Çengel & Boles [8] were used in
The correlation for chemical exergy by Song et al. [19] is a com- X 
bination of two correlations, one for the higher heating value, X exhaust ¼ Ni ðh 0 Þ  T 0 ðs  s0 Þ
h ð23Þ
i
originally developed by Channiwala and Parikh’ [29], and one for i
entropy of the fuel, both based on the carbon, hydrogen, oxygen,
nitrogen, sulphur, and ash content (dry basis) of the fuel. In con- The sum goes over the substances CO2, H2O, and N2. The index 0
trast to many other correlations, which are only valid for a certain represents the state of the surroundings (T0, P0) and h  and s are
type of fuel, Song et al. [19] have developed their correlation(s) for molar enthalpy and entropy. The amount of substance N can be
solid and liquid fuels in general. They even provide a combined found with the combustion equation and the mass of waste inciner-
equation independent of the phase (solid or liquid) of the fuel. ated per year. The result is an exergy loss of 29.1 GW h per year,
The range of applicability of the correlations is limited by the about 3.2% of the expended exergy and ca. 15 times the amount
weight percent (dry basis) intervals for the different elemental of exergy loss due to bottom ash removal. The value of 3.2% is in
constituents and ash content for both the underlying heating value good agreement with the order of magnitude of exergy loss with
and entropy correlations. flue gases given by Lior (5%) [11]. Still, these two contributions from
When comparing the results for specific chemical exergy and bottom ash and exhaust gas can be neglected because they do not
second-law-efficiency to those calculated directly from different lead to a significantly lower upper limit for the achievable sec-
correlations for chemical exergy, a certain spread in the values ond-law-efficiency. One thing to keep in mind is that the result
can be seen. At 25 °C surroundings temperature, the values range for X exhaust is for stoichiometric combustion, while excess air is used
from 26.39 MJ/kg and 14.2% (correlation by Kotas [35]) to in the actual combustion process.
24.35 MJ/kg and 15.4% (Song et al. [19]) and 23.09 MJ/kg and The obtained values for utilization factor, R1-value and second
16.3% (Prins et al. [34]). The results from the correlations by Song law efficiency are surprisingly low with respect to the age of the
et al. [19] and Prins et al. [34] are closest to those obtained by facility. One factor that has not been accounted for yet is the varia-
the other method. The much larger specific chemical exergy from tion of heating values, absolute entropy and the chemical exergy of
the Kotas correlation [35] leads to a much lower second law effi- the waste with the moisture content in the as received and com-
ciency compared to the other results. bustible waste. The climate at the Norwegian west coast around
Many factors that have an impact on the second law efficiency Bergen is quite humid with an average relative humidity of 81%
of a real installation keep it from reaching the theoretical upper in 2012 [21] and large amounts of precipitation. Due to the high
limit of 100%. These are mainly the combustion process and heat relative humidity, the operator sometimes has trouble in getting
transfer through finite temperature differences in the various heat the waste dry enough for incineration. Some of the industrial waste
exchangers, which have already been mentioned. Exergy destruc- is collected in open containers from which rainwater can only eva-
tion during combustion can be up to one third of the fuel’s exergy porate due to lack of drainage. This may lead to a larger amount of
while the loss in heat transfer to the working fluid can be assumed moisture content than the values obtained from the Dutch Phyllis2
to be on the order of 15% [11]. Other causes are friction and database [23] suggest. The analysis, which led to the chemical
unavoidable losses due to thermal radiation, exhaust gas and bot- composition of the waste used in this study, is based on a small
tom ash removal. The latter two mass flows both have tem- number of regular sample inspections of a few hundred kilograms
peratures above the surroundings temperature when they leave waste carried out per year. The type and amount of waste is
the combustion chamber. The exhaust gas needs to be kept at a categorized, for example in paper/pulp, food waste, plastics, etc.
certain temperature to avoid condensation of corrosive species in However, the amount of moisture or liquids in the waste itself or
the exhaust system of the plant and to insure a certain buoyancy in the containers is not measured. This is a root of error in the
of the gases in the surroundings that they are released into. Both amount total moisture in the waste that goes directly from the
bottom and flue gas/fly ash are removed at temperatures above containers and garbage collection trucks into the waste reservoir
the surroundings and therefore have a non-zero exergy. from where the combustion chamber is fed with a crane. Only
The amount of bottom ash removed was 42,000 tons in the ref- small amounts of moisture can evaporate and evaporation can
erence year. With a typical heat capacity of 0.8 kJ/kg K [40] and therefore not be considered to reduce the amount of moisture sub-
assuming the material is incompressible, its exergy can be calculat- stantially prior to incineration.
ed from [8] Another fact that points towards the need to correct the weight
  percentage moisture in the waste is the maximum heat input rate
T the CHP plant in this case study is designed for. The design heat
X bottomash ¼ mcP T  T 0  T 0 ln ð22Þ
T0 input rate is 88 MW. A typical time of steady operation is 8000 h
per year. With the amount of combustible waste from Table 2
Here, T is the temperature of the bottom ash when it leaves the and the heating value of the waste obtained from the Phyllis2 data-
combustion chamber, which is 748.15 K (=475 °C) according to base [23] the average heat input rate would be about 108 MW.
the operator of the plant. The result is 1.89 GW h lost exergy due Table 8 shows a comparison of some key properties of the waste
to bottom ash removal. However, this is just 0.2% of the expended and results for the different efficiencies as function of weight per-
exergy even though the removal temperature appears quite high cent moisture in the combustible fraction of the received waste. As
relative to the surroundings temperature. the presented calculations are based on both dry mass and dry and
The exergy loss due to the exhaust gas can be calculated based ash free mass, an increase in weight percentage moisture leads to a
on the amount of CO2, N2 and H2O available from the combustion smaller combustible mass on both dry and dry and ash free basis.
equation and the mass of waste incinerated per year. The contribu- This on the other hand leads to increases in chemical exergy and
tion from SO2 was neglected as it has shown throughout the study increasing efficiency values. A doubling in the weight percent
that it does not contribute significantly to the results. Both CO2 and moisture leads only to slight changes in the chemical composition
N2 can be treated as ideal gases at the exhaust gas temperature of of the waste as can be seen in the last column of the table. The
158 T. Solheimslid et al. / Energy Conversion and Management 95 (2015) 149–159

Table 8
Comparison of absolute molar entropy, specific chemical exergy, utilization factor based on the lower heating value (LHV), R1-value, second law efficiency at the local
surroundings and standard reference temperature, and the total mass of moisture in the waste as function of weight percent moisture in the combustible fraction of the received
waste. The last column shows the changes in the chemical composition of the waste with increasing moisture content.

wt% moisture s0 at 25 °C xwaste at 25 °C eu LHV R1 gII at 6.8 °C gII at 25 °C Mass moisture Chemical composition of the
(db) (J/mol K) (MJ/kg) (%) (%) (%) (tons) combustible waste
17.7 170.56 23.60 40.6 0.568 17.3 15.9 30,070 C6H10O2.41N0.1S0.01
20 170.50 23.83 41.4 0.579 17.6 16.2 33,979 C6H10.02O2.38N0.1S0.01
25 170.35 24.38 43.1 0.603 18.4 16.9 42,398 C6H10.08O2.33N0.1S0.01
30 170.18 25.03 45.0 0.630 19.2 17.7 50,968 C6H10.14O2.26N0.09S0.01
35 170.00 25.79 47.1 0.659 20.1 18.5 59,538 C6H10.2O2.19N0.09S0.01
40 169.79 26.68 49.3 0.689 21.0 19.3 67,958 C6H10.27O2.11N0.08S0.01

changes in absolute entropy are about 0.3% only. The specific che- predicted by Grosso et al. [4] for Scandinavian waste incineration
mical exergy of the waste increases by 13% from 23.6 MJ/kg (db) at plants of the recovery-operation type. This value is surprisingly
17.7 wt% moisture to 26.66 MJ/kg (db) at 40 wt% moisture. The low and suggests that there is room for improvement. On the other
combustible mass on dry and dry and ash free basis decreases hand, it is more likely that the local climate is not adequately
accordingly. The combined effect is that the same amount of recov- accounted for with respect to humidity and precipitation by means
ered exergy is based on a smaller amount of expended exergy. This of the moisture content in the waste. The second law efficiency can
leads to an increase in both the second-law-efficiency and the therefore be expected to be ca. 20% or larger at the local surround-
other efficiencies. By increasing the weight percent moisture, the ings temperature. In this case, energy utilization would be above
utilization factor, R1-value and exergy efficiency increase towards 47% (instead of 40.6%) and the R1-efficiency larger than 0.65 com-
values that appear more realistic, according to the plant operator. A pared to the calculated 0.568.
study conducted by the SP Technical Research Institute of Sweden Focusing on the production of electricity from waste can give
on Swedish waste and waste imported from the UK gave moisture larger increases in exergy recovery and exergy efficiency than
contents of 38 wt% and 32 wt% respectively [41]. A moisture con- increasing the delivery of district or process heat although the lat-
tent of 35 wt% in this study corresponds to a mass of almost ter is necessary to reach exergy efficiencies of 20% or more, as
60,000 tons H2O. The corresponding second law efficiencies are Grosso et al. [4] have shown. Waste heat recovery from slag and
then 18.5% at 25 °C and 20.1% at the local surroundings tem- especially exhaust gases also has the potential to contribute to
perature (see Table 8). Together with an R1-value of 0.659, the an increase in exergy recovery. This will be considered for future
facility in this case study would fulfil the criteria for a recovery upgrades of the plant.
operation as defined by the European Union [9]. The efficiencies
would in this situation also be closer to the average properties of
Norwegian waste incineration facilities as estimated by Grosso Acknowledgements
et al. [4]. The actual efficiencies of the CHP-plant in this study
are therefore likely to be higher than the presented calculations The authors are grateful to Ragnar Gjengedal, Jarle S. Diesen,
suggest which are based on the waste analysis as it is conducted Karl Weydahl (Bergen University College), Camilla Tangenes and
nowadays. Efficiencies corresponding to moisture content between Øyvind Underdahl Holm (both BIR AS) for useful discussions and
35 wt% and 40 wt% are more likely as the theoretical heat input detailed information about the waste incineration plant.
rate drops below the design value of 88 MW just above 30 wt%.
Future measurements of the actual water and moisture content
Appendix A. Supplementary material
in regular sample inspections of the received waste are planned.
They are expected to lead to a clarification and better understand-
Supplementary data associated with this article can be found, in
ing of the influence of the local climate and waste handling on the
the online version, at http://dx.doi.org/10.1016/j.enconman.2015.
moisture content in the combustible fraction of the waste.
02.026.

5. Conclusion References

The different methods employed for calculating the exergy effi- [1] Demirbas A. Waste management, waste resource facilities and waste
ciency of the energy-recovery-from-waste-operation in this study conversion processes. Energy Convers Manage 2011;52(2):1280–7.
[2] Finney KN, Chen Q, Sharifi VN, Swithenbank J, Nolan A, White S, et al.
yield comparable results. A comparison of the influence of heating Developments to an existing city-wide district energy network: Part II –
value and entropy contributions to the exergy of the waste shows, Analysis of environmental and economic impacts. Energy Convers Manage
that the entropy part has a minor impact on the results that 2012;62:176–84.
[3] Marthinsen J, Sannberg K, Johansen M. Fornybarandel i avfall til norske
appears almost negligible in case of the waste in this study. This forbrenningsanlegg. Oslo, Norway: NVE; 2011.
also means that the method to calculate the entropy of the waste [4] Grosso M, Motta A, Rigamonti L. Efficiency of energy recovery from waste
based on its chemical composition by means of the equation for incineration, in the light of the new Waste Framework Directive. Waste
Manage 2010;30(7):1238–43.
the specific heat by Laštovka et al. [22] can be a useful and reliable [5] Ouda OKM, Cekirge HM, Raza SAR. An assessment of the potential contribution
alternative to other correlations. The observed difference between from waste-to-energy facilities to electricity demand in Saudi Arabia. Energy
calculated and literature values for the absolute entropy for select- Convers Manage 2013;75:402–6.
[6] Chakraborty M, Sharma C, Pandey J, Gupta PK. Assessment of energy
ed hydrocarbons does not fall into account in a significant manner
generation potentials of MSW in Delhi under different technological options.
but clearly shows the need for more reliable correlations and Energy Convers Manage 2013;75:249–55.
experimental investigations of heat capacity and entropy of [7] Ng WPQ, Lam HL, Varbanov PS, Klemeš JJ. Waste-to-Energy (WTE) network
biomass as function of temperature. synthesis for Municipal Solid Waste (MSW). Energy Convers Manage
2014;85:866–74.
The obtained second-law efficiency of 17.3% for the local [8] Çengel YA, Boles MA. Thermodynamics. An engineering approach. 7th (SI
surroundings temperature lies below the range of efficiencies Units) ed. New York (USA): McGraw-Hill; 2011.
T. Solheimslid et al. / Energy Conversion and Management 95 (2015) 149–159 159

[9] European Union. Directive 2008/98/EC of the European Parliament and of the [26] Energy Research Centre of the Netherlands. Phyllis2 – Help Page [cited 10th
Council og 19th November 2008 on waste repealing with certain directives. January 2015]. <https://www.ecn.nl/phyllis2/Home/Help>.
Official Journal of the European Union. 2008. [27] United States Institute of Gas Technology – Division of Coal Conversion. Coal
[10] Verkhivker GP, Kosoy BV. On the exergy analysis of power plants. Energy Conversion Systems: Technical Data Book: U.S. Government Printing Office;
Convers Manage 2001;42(18):2053–9. 1978.
[11] Lior N. Thoughts about future power generation systems and the role of exergy [28] Komilis D, Evangelou A, Giannakis G, Lymperis C. Revisiting the elemental
analysis in their development. Energy Convers Manage 2002;43(9– composition and the calorific value of the organic fraction of municipal solid
12):1187–98. wastes. Waste Manage 2012;32(3):372–81.
[12] Ishida M. The role and limitations of endoreversible thermodynamics. Energy [29] Channiwala SA, Parikh PP. A unified correlation for estimating HHV of solid,
1999;24(12):1009–14. liquid and gaseous fuels. Fuel 2002;81(8):1051–63.
[13] Ray TK, Datta A, Gupta A, Ganguly R. Exergy-based performance analysis for [30] Goodman BT, Wilding WV, Oscarson JL, Rowley RL. Use of the DIPPR database
proper O&M decisions in a steam power plant. Energy Convers Manage for development of quantitative structureproperty relationship correlations:
2010;51(6):1333–44. heat capacity of solid organic compounds. J Chem Eng Data 2003;49(1):24–31.
[14] Padilla M, Revellin R, Bonjour J. Exergy analysis of R413A as replacement of [31] Maxima, a Computer Algebra System. (Version 5.34.1) 2014 [cited 15th
R12 in a domestic refrigeration system. Energy Convers Manage December 2014]. <http://maxima.sourceforge.net/>.
2010;51(11):2195–201. [32] Mountouris A, Voutsas E, Tassios D. Solid waste plasma gasification:
[15] Kaviri AG, Jaafar MNM, Lazim TM. Modeling and multi-objective exergy based equilibrium model development and exergy analysis. Energy Convers
optimization of a combined cycle power plant using a genetic algorithm. Manage 2006;47(13–14):1723–37.
Energy Convers Manage 2012;58:94–103. [33] Szargut J, Styrylska T. Approximate evaluation of the exergy of fuels.
[16] Dunbar WR, Lior N. Sources of combustion irreversibility. Combust Sci Technol Brennstoff Wärme Kraft 1964;16:589–96.
1994;103(1–6):41–61. [34] Prins MJ, Ptasinski KJ, Janssen FJJG. Thermodynamics of gas-char reactions:
[17] Husetuft Ommedal HK, Haugvaldstad L, Rikstad L. Separation of wet organic first and second law analysis. Chem Eng Sci 2003;58(3–6):1003–11.
material in waste treatment. Bergen (Norway): Bergen University College; [35] Kotas TJ. The exergy method of thermal plant analysis. Butterworth-
2013. Heinemann; 1985.
[18] Harneshaug HK, Solheimslid T. Calculation of second law efficiency of energy [36] National Institute of Standards and Technology. NIST Chemistry Webbook
recovery from waste. Bergen (Norway): Bergen University College; 2014. [cited 15th July 2014]. <http://webbook.nist.gov/chemistry/>.
[19] Song G, Xiao J, Zhao H, Shen L. A unified correlation for estimating specific [37] Smirnova NN, Kandeev KV, Bykova TA. The thermodynamic properties of 1,6-
chemical exergy of solid and liquid fuels. Energy 2012;40(1):164–73. Hexanediol in the temperature range from T near 0 K to 370 K. Russ J Phys
[20] Bejan A, Tsatsaronis G, Moran M. Thermal design & optimization. New York Chem A 2005;79(6):857–61.
(USA): John Wiley & Sons, Inc.; 1996. [38] Laštovka V, Sallamie N, Shaw JM. A similarity variable for estimating the heat
[21] Iden K, Kristiansen S, Mamen J, Szewczyk-Bartnicka H, TilleyTajet HT. Været i capacity of solid organic compounds: Part I. Fundamentals. Fluid Phase
Norge, Klimatologisk oversikt. Året 2012. Oslo: Meteorologisk institutt; 2013. Equilibria 2008;268(1–2):51–60.
[22] Laštovka V, Shaw JM. Predictive correlation for Cp of organic solids based on [39] Laštovka V, Fulem M, Becerra M, Shaw JM. A similarity variable for estimating
elemental composition. J Chem Eng Data 2007;52(4):1160–4. the heat capacity of solid organic compounds: Part II. Application: heat
[23] Energy Research Centre of the Netherlands. Phyllis2 [cited 2nd December capacity calculation for ill-defined organic solids. Fluid Phase Equilib
2014]. <https://www.ecn.nl/phyllis2/>. 2008;268(1–2):134–41.
[24] Themelis NJ, Kim YH, Brady MH. Energy recovery from New York City [40] Klein R, Nestle N, Niessner R, Baumann T. Numerical modelling of the
municipal solid wastes. Waste Manage Res 2002;20(3):223–33. generation and transport of heat in a bottom ash monofill. J Hazard Mater
[25] European Commission. Guidelines on the interpretation of the R1 energy 2003;100(1–3):147–62.
efficiency formula for incineration facilities dedicated to the processing of [41] Bisaillon M, Johansson I, Jones F, Sahlin J. Fuel Quality – Composition and
municipal solid waste according to annex II of directive 2008/98/EC on waste properties of waste fuel to waste-to-energy plants. Borås (Sweden): Waste
2011 [cited 16th June 2014]. <http://ec.europa.eu/environment/waste/ Refinery, SP Sveriges Tekniska Forskningsinstitut; 2013.
framework/pdf/guidance.pdf>.

Você também pode gostar