+ All documents
Home > Documents > Simulation and multi-objective optimization of a combined heat and power (CHP) system integrated...

Simulation and multi-objective optimization of a combined heat and power (CHP) system integrated...

Date post: 11-Nov-2023
Category:
Upload: concordia
View: 0 times
Download: 0 times
Share this document with a friend
34
Author’s Accepted Manuscript Simulation and multi-objective optimization of a combined heat and power (CHP) system integrated with low-energy buildings Jamasb Pirkandi, Mohammad Ali Jokar, Mohammad Sameti, Alibakhsh Kasaeian, Fazel Kasaeian PII: S2352-7102(15)30038-3 DOI: http://dx.doi.org/10.1016/j.jobe.2015.10.004 Reference: JOBE62 To appear in: Journal of Building Engineering Received date: 6 July 2015 Revised date: 1 October 2015 Accepted date: 25 October 2015 Cite this article as: Jamasb Pirkandi, Mohammad Ali Jokar, Mohammad Sameti Alibakhsh Kasaeian and Fazel Kasaeian, Simulation and multi-objective optimization of a combined heat and power (CHP) system integrated with low energy buildings, Journal of Building Engineering http://dx.doi.org/10.1016/j.jobe.2015.10.004 This is a PDF file of an unedited manuscript that has been accepted fo publication. As a service to our customers we are providing this early version o the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting galley proof before it is published in its final citable form Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain www.elsevier.com/locate/jobe
Transcript

Author’s Accepted Manuscript

Simulation and multi-objective optimization of acombined heat and power (CHP) system integratedwith low-energy buildings

Jamasb Pirkandi, Mohammad Ali Jokar,Mohammad Sameti, Alibakhsh Kasaeian, FazelKasaeian

PII: S2352-7102(15)30038-3DOI: http://dx.doi.org/10.1016/j.jobe.2015.10.004Reference: JOBE62

To appear in: Journal of Building Engineering

Received date: 6 July 2015Revised date: 1 October 2015Accepted date: 25 October 2015

Cite this article as: Jamasb Pirkandi, Mohammad Ali Jokar, Mohammad Sameti,Alibakhsh Kasaeian and Fazel Kasaeian, Simulation and multi-objectiveoptimization of a combined heat and power (CHP) system integrated with low-energy buildings, Journal of Building Engineering,http://dx.doi.org/10.1016/j.jobe.2015.10.004

This is a PDF file of an unedited manuscript that has been accepted forpublication. As a service to our customers we are providing this early version ofthe manuscript. The manuscript will undergo copyediting, typesetting, andreview of the resulting galley proof before it is published in its final citable form.Please note that during the production process errors may be discovered whichcould affect the content, and all legal disclaimers that apply to the journal pertain.

www.elsevier.com/locate/jobe

1

Simulation and multi-objective optimization of a combined heat and

power (CHP) system integrated with low-energy buildings

Jamasb Pirkandi1, Mohammad Ali Jokar2, Mohammad Sameti2*

Alibakhsh Kasaeian2, Fazel Kasaeian3

1Department of Aerospace Engineering, Malek Ashtar University of Technology, Tehran, Iran

2Department of Renewable Energies, University of Tehran, Tehran, Iran

3Department of Materials Science and Engineering, Sharif University of Technology, Tehran, Iran

*Corresponding Author:

Mohammad Sameti

Department of Renewable Energies, Faculty of New Sciences and Technologies

University of Tehran, North Karegar St., Tehran, Iran

E-mails: [email protected]

Tel: +98 9132124120

Fax: +98 21 88617087

2

Abstract One of the novel applications of gas turbine technology is the integration of

combined heat and power (CHP) system with micro gas turbine which is spreading widely in

the field of distributed generation and low-energy buildings. It has a promising great

potential to meet the electrical and heating demands of residential buildings. In this study, A

MATLAB code was developed to simulate and optimize the thermoeconomic performance

of a gas turbine based CHP cycle. Three design parameters of this cycle considered in this

research are compressor pressure ratio, turbine inlet temperature, and air mass flow rate.

Firstly, two objective functions including exergetic efficiency and net power output were

chosen to achieve their maximum level. Genetic algorithm (GA) was used as the

optimization technique to determine the optimum behavior of the system. Variation of

exergy destruction rate and exergetic efficiency with turbine inlet temperature (TIT) and air

mass flow rate were also studied for each component. Based on the exergetic analysis,

suggestions were given for reducing the overall irreversibility of the thermodynamic cycle.

To have a good insight into this study, a sensitivity analysis for important parameters was

also carried out. Finally, based on the exergy analysis and utilization of economic and

environmental functions, a multi-objective approach was performed to optimize the system

performance.

Keywords low-energy buildings, cogeneration, micro gas turbine (MGT), optimization,

genetic algorithm (GA), economic analysis, environmental consideration

Nomenclature

Cost of fuel per energy unit ($/kJ) C Cost flow rate ($/s) Specific heat at constant pressure (kJ/kg.K)

3

CRF Capital recovery factor

Exergy flow rate (kW)

Exergy destruction rate (kW) e Specific exergy (kJ/kmol) h Enthalpy (kJ/kg) k Specific heat ratio (Cp/Cv) LHV Lower heating value (kJ/kg) Mass flow rate (kg/s) N Annual number of the operation hours of the unit P Pressure (KPa) R Gas constant (kJ/kg.K) Compressor pressure ratio

S Entropy (kJ/kg.K) T Temperature (K) U Overall heat transfer coefficient (W/m2.K)

Net power output (kW) Z Capital cost of the component ($)

Capital cost rate of the component ($/s) ΔP Pressure drop (KPa) ΔTLM Log mean temperature difference (K) ƞ efficiency ξ Ratio of chemical exergy to Lower Heating Value Specific heat ratio Maintenance factor of the equipment Molar fraction Subscripts a Air AC Air compressor CC Combustion chamber e Exergetic f Fuel FC Fuel compressor g Combustion gases gen Generator GT Gas turbine HE Heat exchanger Q Heat rate (kW) rec Recuperator W Work (kW) Superscripts ch Chemical ph Physical

4

1. Introduction

A Low-energy building is any type of building that from design, technologies, and building

products uses less energy, from any source, than a traditional or average contemporary

house [1]. They are the practice of sustainable design, sustainable architecture, low energy

building, energy-efficient landscaping [2], and energy system optimization [3]. Meanwhile,

distribute generation (DG) is predicted to play an increasing role into the electric power

system for buildings in the near future [4]. Distributed energy resources are small modular

power generation systems that can be located at or near the site where energy is used. In

conventional energy systems, electrical power is conveyed from large-scale plants located

far away from the consuming region, while energy for heating is supplied separately as fuel.

In this way, more than 50% of the energy content of the fuel is lost at the power plant alone

because of energy conversion inefficiencies and is discharged in the form of waste. Further

losses occur in the electric power transmission and distribution network in the form of

electric current losses and power transformation losses [5, 6]. DG with a cogeneration

system is one of the options because it can efficiently utilize exhaust heat. Following are the

benefits of such a power generation [7]:

This system can be easily and effectively installed and operated both in high-demand

or rural areas.

Power can be distributed and transmitted with low losses.

Exhaust heat can be used efficiently by this system.

This system can either be used independently or as a supportive system.

Many studies have been conducted many aspects of cogeneration systems. Today the main

potential for CHP dissemination seems to be in the residential sector. Until the present,

several technical, environmental, economic and legislative problems have curbed the spread

of CHP technology in this sector, especially for electric power sizes of a few kW [8-10].

Additionally, authors proposed numerous researches in the operation of CHP systems in

large power plants e.g. [11, 12]. Tehrani et al. [11] presented a method of design a

trigeneration plant. Their idea is to recover the exhaust hot gases of a GT power plant in

order to supply dynamic HVAC (heating, ventilating, and air conditioning) load of the new

town of Parand, and also use the rest heat potential to feed a steam turbine cycle. Karaali et

al. [12] introduced a novel thermoeconomic optimization method for real complex cycles.

The objective of this paper is to apply this method to four cogeneration cycles that are

simple cycle, inlet air cooling cycle, air preheated and air-fuel preheated cycles for analyzing

and optimizing. The four cycles are thermoeconomically optimized for constant power and

steam mass (30 MW and 14 kg/s saturated steamflow rate at 2000 kPa), for constant power

(30 MW) and for variable steam mass, and for variable power and steam mass by using the

cost equation method and the effect of size on equipment method.

Cogeneration systems utilizing internal combustion engines and gas turbines in open cycle

are the most utilized technologies in this field worldwide. The interest of MGTs as

distributed energy systems lies in their low environmental impact in terms of pollutants.

5

MGTs present some unique characteristics compared with the larger gas turbine engines

such as the high rotational speeds, the ability to burn various fuels and the radial

turbomachines [13, 14].

MGTs are usually designed for natural gas, but there is the ability to utilize other fuels such

as those based on biomass. Mozafari et al. [15] performed the optimization of MGT by

exergetic-economic-environmental analysis considering various fuels for the system. The

results showed that and the trends of variations of second law efficiency and cost rate of

owning and operating the whole system are independent of the fuels. In this article the fuel

which is used in the combustion chamber is methane (CH4).

Thermodynamic analysis can be a perfect tool for identifying the ways for improving the

efficiency of fuel use, and determining the best configuration and equipment size for a

cogeneration plant [16]. As mentioned above, extending this system to combined heat and

power generation is a way of increasing productivity with recovery of the heat discarded

from the inefficient energy conversion of producing electrical power. These systems are

expected to find applications in the cogeneration market for various heating and power

requirements [17].

A thermodynamic analysis is proposed in this paper to study MGT to minimize fuel

consumption and maximize net exergetic efficiency. Energetic, economic and environmental

performance of the system is investigated. The energy and entropy balance equations of the

entire system will be additionally obtained. To find the optimum design parameters of the

system genetic algorithm is used and a simulation program is developed in MATLAB

software. This simulation investigates the effects of various performance parameters, such

as the compression ratio (rp), mass flow rate of air and turbine inlet temperature (TIT), on

the exergetic efficiency and irreversibility of the plant. Sensitivity studies for two important

parameters including net power generation and exergetic efficiency is done to show the

abilities and authenticity of simulated system. Next, a common economic analysis for the

system is carried out. The quantity of NOx and CO emissions is considered for environmental

purposes. In the present study, the cost of pollution damage is considered to be added

directly to total cost rate of the system production. Therefore, the third objective function is

sum of the thermodynamic and environomic objectives.

In the heat production sector, there is several choices to select the kind of heat production

including hot water, steam water, hot air and even producing cold water after an absorption

chiller cycle. In this article, we really did not discuss this issue and to demystify the exergetic

efficiency of the system and calculate the total cost rate, we assume that a water pump is

driven using electricity produced by micro gas turbine to pump the water in ambient

temperature into the heat exchanger and hot water temperature of 80 °C will be delivered

to the residential consumer.

In the open literature, various configurations of integrating micro gas turbine and

regenerator with a process plant have been analyzed through different techniques such as

exergy analysis, thermoeconomics, physical insight based pinch analysis, R-curve analysis

and mathematical optimization (mixed integer linear programming and nonlinear

6

programming) [18, 19]. Ameri et al. [20] presented a thermodynamic analysis of a tri-

generation system based on micro gas turbine with a steam ejector refrigeration system and

heat recovery steam generator. Caresana et al. focused on the effect of ambient

temperature on the performance of a microturbine in cogeneration arrangement and by

providing a simulation code, entered in detail into the machines’ behavior. A performance

chart has been drawn showing how the MGT working range changes with temperature in

specific intervals [21]. A conceptual trigeneration system is proposed in ref. [22] based on

the conventional gas turbine cycle for the high temperature heat addition while adopting

the heat recovery steam generator for process heat and vapor absorption refrigeration for

the cold production.

Several researchers carried out the exergoeconomics analysis and optimization for thermal

systems. Barzegar et al. [23] has carried out an excellent review in this issue. The reviewed

results show that although exergoeconomic analyses are so useful in power generation,

they cannot find the optimal design parameters in such systems. So, using the optimization

procedure as thermoeconomics is essential with respect to thermodynamic laws.

Optimization in engineering design has always been of great importance and interest

particularly in solving complex real-world design problems. There are many calculus-based

methods including gradient approaches to search for mostly local optimum solutions and

well documented in ref. [24]. However, some basic difficulties in the gradient methods such

as their strong dependence on the initial guess can cause them to find a local optimum

rather than a global one. This has led to other heuristic optimization methods, particularly

genetic algorithms (GAs) being used extensively during the last decade. The main difference

between GA and traditional calculus based techniques is that GAs work with a population of

candidate solutions, not a single point in search space. This helps significantly to avoid being

trapped in local optima as long as the diversity of the population is well preserved [25].

2. System description

The first generation of micro-turbines has been introduced to the energy market recently.

Their size range from 10 to 250 kW with electrical efficiency of about 30%, and coupled with

thermally activated devices (to produce either heating or cooling by using the exhaust heat),

may achieve high total efficiencies in the range of 65% as shown in recent realistic studies

[26] although the initial analysis envisaged higher efficiencies (in the range of 75%) [27].

Some advantages of microturbines are as follows [28]:

They can be installed on-site especially if there are space limitations. Also they are

compact in size and light in weight with respect to traditional combustion engines.

They are very efficient (more than 80%) and have lower emissions (less than 10 ppm

NOx) with respect to large scale ones.

They have well-known technology and they can start-up easily, have good load

tracking characteristics and require less maintenance due their simple design.

7

They have lower electricity costs and lower capital costs than any other DG

technology costs.

They have a small number of moving parts with small inertia not like a large gas

turbine with large inertia.

Modern power electronic interface between the MT and the load or grid increases

its flexibility to be controlled efficiently.

In this paper a combined cycle micro gas turbine is investigated and for this purpose, a

commercially available C30 engine manufactured from Capstone, which is also sold as part

of the C30 CHP cogeneration unit is considered [29-31]. The unit is usually fired with natural

gas; biofuel (gas or oil) and diesel fuel versions are also available. This micro gas turbine

produces 30 kW of electrical power with 26 ± 2% efficiency at ISO (international standards

organization) conditions. Design parameters are presented in Table 1. In the present paper,

the system is only analyzed at the design point. Thus, all components required for start-up

and partial-load operation are not considered. So, this model consists of following

components: a centrifugal air compressor; a centrifugal fuel compressor; a plate-fin

recuperator, air-gas turbine exhaust heat exchanger; a plate-fin recuperator, fuel-gas

turbine exhaust heat exchanger; a combustion chamber, a radial turbine and a heat

exchanger. Future work will also take into account partial-load operation by using

characteristic maps of the turbine and the compressors. All the rotating components are

mounted on a single shaft. The layout proposed for the plant is shown in Figure 1.

Table 1 Capstone C30 design parameters [29, 30]

Figure 1 Schematic diagram of the cycle power plant

2.1. Thermodynamic modeling

To determine the exergy at different points of the power plant, the thermodynamic

properties of the cycle should be specified. This is done usually through physical modeling.

This section includes the rules that form the physical model. The major portion of this

modeling consists of the application of the mass and energy balance equations. The

principle of operation can be summarized as follows:

Air is compressed by the air compressor up near to the combustion chamber

operating pressure. The air is then preheated in the plate-fin heat exchanger and

brought to the combustion chamber (state point 3).

Similarly, the fuel (natural gas) is compressed by the fuel compressor, preheated in

the fuel-exhaust gas, plate-fin heat exchanger and then brought to the combustion

chamber (state point 3f).

8

The chemical reactions, occurring in the combustion chamber, produce thermal

energy. The high energy stream at state point 4 is first enter to the turbine and then

used to preheat air and fuel in the counter-flow heat exchangers.

The expansion in the gas turbine supplies mechanical energy and then electric

power. As mentioned, the gas turbine outlet stream (state point 5) can be used to

preheat both fuel and air flows. Any residual thermal energy left over is available for

cogeneration purposes: to this scope, a water–gas, plate-fin heat exchanger is

included.

For the purpose of analysis the following assumptions are made:

The whole system operates at a steady state condition.

Ideal-gas mixture principles apply for the air, fuel and the combustion products.

All components operate without heat loss.

The exit temperature is above the dew point temperature of the combustion

product.

The reference environmental state for the system is T0 =250C and P0= 101.3 KPa.

The turbine and the compressors have been assumed as adiabatic.

The energy variation and the kinetic and potential exergies have been assumed as

negligible.

The result of friction and other irreversibilities for flow through the compressor is

explained with an isentropic efficiency.

Pressure drop in combustion chamber, two recuperators and heat exchanger are as

follow: P P P P .

Developing equation for each component leads to the following systems of equations:

Air compressor, (1) (2) (T T )

(3) T T

{

[

( ) ⁄

]

}

(4) P P

First recuperator,

(5) (T T ) (T T )

(6) P P ( ΔP )

(7) P P ( ΔP )

Fuel compressor, (8)

9

(9) (T T )

(10) T T

{

[

( ) ⁄

]

}

Second recuperator,

(11) P P

(12) (T T ) (T T )

(13) P P ( ΔP )

(14) P P ( ΔP )

Combustion chamber,

(15) L ( ) L

(16) (T T )

(17) (T T )

(18) P P ( ΔP ) (19)

Since we should have the molar values of output gases in order to obtain the chemical

exergy at the exit point of the combustion chamber, we need to solve the combustion

equation of the combustion chamber. In this case, the molar values of air are considered as

fixed and by using the equation below, the molar values of gases leaving the combustion

chamber are obtained from:

(20) ( ) ( )

In the equation (20), the value of λ and α obtained from the energy law in the combustion

chamber, and knowing this value, the molar values of gases exiting the combustion chamber

are calculated and subsequently used in the step associated with the calculation of chemical

exergy:

(21)

(22)

(23)

(24)

Gas turbine,

(25)

(26) T T

{

[

P P

( ) ⁄

]

}

(27) (T T )

Heat exchanger,

(28) (T T ) (T T )

(29) (T T )

(30) ( T )

(31) P

10

(32) P P ( ΔP )

The net electrical power output of the gas turbine, , can be expressed as:

(33) ( )

(34) ( )

The other important criterion is exergetic efficiency of the system, and it can be given as the

ratio of total exergy output to exergy rate of fuel:

(35) L ξ

(36)

L ξ

Here, LHV is lower heating value (for methane it is equal to 50000 KJ/Kg). Equation 35 is

considered for the power generation system regardless of the heat exchanger and equation

36 for the CHP system.

It should be noted that the calculation of the specific heat of the materials is often done

using relationships depicted in [32, 33]. In this article we use the relationships in [32].

Comparison between these references is carried out in Figure 2.

Figure 2 Comparison between thermodynamic properties of material used in this article

based on references [31, 32]

2.2. Exergy analysis

In summary, exergy is a measure of the maximum capacity of a system to perform useful

work as it proceeds to a specified final state in equilibrium with its surrounding. And it can

be divided into four distinct components. The two important ones are the physical exergy

and chemical exergy. In this study, the two other components which are kinetic exergy and

potential exergy are assumed to be negligible as the elevation and speed have negligible

changes. The physical exergy is defined as the maximum theoretical useful work obtained as

a system interacts with an equilibrium state [34, 35]. The chemical exergy is associated with

the departure of the chemical composition of a system from its chemical equilibrium. The

chemical exergy is an important part of exergy in combustion process. Applying the first and

the second law of thermodynamics, the following exergy balance is obtained:

(37) ∑

In this equation is the exergy destruction. Other terms in this equation is as follow:

(38) ( T T)

(39) (40)

(41) ( ) T ( )

11

Where and are the corresponding exergy of heat transfer and work which cross

the boundaries of the control volume, T is the absolute temperature and (K) and (0) refers

to the ambient conditions respectively. The mixture chemical exergy is defined as follows

[35]:

(42) [∑

T ∑ L

]

For the evaluation of the fuel exergy, the above equation cannot be used. Thus, the

corresponding ratio of simplified exergy is defined as the following [35]:

(43) ξ L ⁄ Due to the fact that for the most of usual gaseous fuels, the ratio of chemical exergy to the

LHV is usually close to unity, one may write:

(44) ξ ξ

For gaseous fuel with CxHy, the following experimental equation is used to calculate ξ:

(45) ξ

Recently exergy analyses have been employed for analysis, design, performance

improvement and optimization of thermal systems, including microCHP plants. It is well-

known that exergy can be used as a potential to determine the location, type and true

magnitude of exergy loss (or destruction) too [36]. Therefore, it can play an important issue

in developing strategies and in providing guidelines for more effective use of energy in the

existing power plants [37]. In the present work, for the exergy analysis of the plant, the

exergy of each line is calculated at all states and the changes in the exergy are determined

for each major component. The source of exergy destruction (or irreversibility) in

combustion chamber is mainly combustion (chemical reaction) and thermal losses in the

flow path. However, the exergy destruction in the heat exchangers of the system is due to

the large temperature difference between the hot and cold fluids. The exergy destruction

rate and the exergetic efficiency for each component for the whole system in the power

plant are shown in Table 2.

Table 2 The exergy destruction rate and exergy efficiency equations for plant components

The code developed for thermodynamic optimization purposes is written in MATLAB

software and is based on a number of built-in functions, tools and externally developed

subroutines. The model includes 6 important fixed parameters (see Table 3) and the

variables selected for the optimization are: The compressor pressure ratio, Pr, the turbine

inlet temperature (TIT) and the air mass flow rate ( ). Fixed parameters remain constant

during the optimization. The decision variables can vary in a given range and represent the

independent variables in design optimization of the system (Table 4). Obviously, only some

of the possible sets of values for the decision variables correspond to actual feasible

designs. The others are automatically rejected by the code. For any acceptable design, the

12

model calculates all energy, entropy, and exergy flow rates entering and leaving each

component.

Table 3 Important fixed parameters and their values

Table 4 Decision variables and their values

2.3. Thermoeconomical Analysis

Thermoeconomics combines the exergy analysis with the economic aspects and consider

the related costs of the thermodynamic inefficiencies in the total product cost of the

system. The total cost rate of operation for the installation is obtained from:

(46) ∑

Where is the total cost rate of fuel ($/s) and is the capital cost rate ($/s) of the kth

equipment item. is introduced in the next section.

(47) ⁄

Where Zk, CRF, N (8000 hours) and (1.06) are the purchase cost of kth component in

dollar, the capital recovery factor, the annual number of the operation hours of the unit and

the maintenance factor. The expression for each component of the MGT plant and

economic model is presented in Table 5. Constants used in the equation of Table 5 are

showed in Table 6. The CRF depends on the interest rate as well as estimated equipment life

time. (Value of CRF this paper is: 0.182) [33, 38, 39].

Cost of fuel rate is defined as follows:

(48) L

Where Cf = 0.004 $/MJ is the regional cost of fuel per unit of energy (on LHV basis) [33, 38],

is the fuel mass flow rate.

Table 5 Equation for calculation the purchase costs for the system components

Table 6 Constants used in the equation of Table 5 for the purchase cost of the components

2.4. Environmental Analysis

In order to minimize the environmental impacts, the objective is to increase the efficiency of

energy conversion processes and, thus, decrease the amount of fuel and the related overall

environmental impacts, especially the release of carbon dioxide as a major greenhouse gas.

13

Therefore, optimization of thermal systems based on this fact has been an important

subject in recent years. Although there are many papers in the literature, dealing with

optimization of CHP plants, they consider no environmental impacts. For this reason, one of

the major goals of the present work is to consider the environmental impacts as producing

the CO and NOx. As discussed in ref. [40], the amount of CO and NOx produced in the

combustion chamber and combustion reaction also change mainly by the adiabatic flame

temperature. The adiabatic flame temperature in the primary zone of the CC is derived as

follows:

(49) T ( ( ) )

Where π is a dimensionless pressure p/pref (p being the combustion pressure p3, and pref =

101.3 KPa); θ is a dimensionless temperature T/Tref (T being the inlet temperature T3, and

Tref = 300 K); ψ is the H/C atomic ratio (ψ = 4, the fuel being pure methane); σ = φ for being

the fuel to air equivalence ratio (φ is assumed constant); It is considered 0.64 for the

combustion equation with the fuel of methane. x, y and z are quadratic functions of σ; A, α,

β and λ are constants (different sets of constants are used for different ranges of θ). The

constants for equations (49 – 52) are obtained from references [33, 38].

(50)

(51)

(52)

The adiabatic flame temperature is used in the semi analytical correlations proposed by Rizk

et al. [41] to determine the pollutant emissions in grams per kilogram of fuel:

(53) ( T ⁄ )

P (ΔP P ⁄ )

(54) ( T ⁄ )

P (ΔP P ⁄ )

Where τ is the residence time in the combustion zone (τ is assumed constant and is equal to

0.002 s); Tpz is the primary zone combustion temperature; p3 is the combustor inlet

pressure; Δp3/p3 is the non-dimensional pressure drop in the combustor. The cost of

environmental impacts is derived as follow:

(55)

Where:

(56) ⁄ (57) ⁄

3. The Genetic Algorithm

The genetic algorithm (GA) is a stochastic global search method that mimics the metaphor

of natural biological evolution. GAs operates on a population of potential solutions applying

the principle of survival of the fittest to produce (hopefully) better and better

approximations to a solution. At each generation, a new set of approximations is created by

the process of selecting individuals according to their level of fitness in the problem domain

14

and breeding them together using operators borrowed from natural genetics. This process

leads to the evolution of populations of individuals that are better suited to their

environment than the individuals that they were created from, just as in natural adaptation

[25]. Individuals, or current approximations, are encoded as strings, chromosomes,

composed over some alphabet(s), so that the genotypes (chromosome values) are uniquely

mapped onto the decision variable (phenotypic) domain. Having decoded the chromosome

representation into the decision variable domain, it is possible to assess the performance, or

fitness, of individual members of a population. This is done through an objective function

that characterizes an individual’s performance in the problem domain. Thus, the objective

function establishes the basis for selection of pairs of individuals that will be mated together

during reproduction. During the reproduction phase, each individual is assigned a fitness

value derived from its raw performance measure given by the objective function. This value

is used in the selection to bias towards more fit individuals. Highly fit individuals, relative to

the whole population, have a high probability of being selected for mating whereas less fit

individuals have a correspondingly low probability of being selected [35].

A GA usually some operators that act on the chromosomes of each generation include

recombination and mutation. The recombination operator is used to exchange genetic

information between pairs, or larger groups of individuals. A further genetic operator, called

mutation, causes the individual genetic representation to be changed according to some

probabilistic rule. Mutation is generally considered to be a background operator that

ensures that the probability of searching a particular subspace of the problem space is never

zero. After recombination and mutation, the individual strings are then, if necessary,

decoded, the objective function evaluated, a fitness value assigned to each individual and

individuals selected for mating according to their fitness, and so the process continues

through subsequent generations. In this way, the average performance of individuals in a

population is expected to increase, as good individuals are preserved and multiplied with

one another and the less fit individuals die out. The GA is terminated when some criteria are

satisfied, e.g. a certain number of generations, a mean deviation in the population, or when

a particular point in the search space is encountered [42]. The basic steps for the application

of a GA for an optimization problem are summarized in Figure 3.

Figure 3 GA flow chart

4. Results and discussion

4.1. Model verification

After applying the thermodynamic relations listed in section 2, and using the developed

simulation code, the results of thermodynamic optimization are obtained. The assumptions

of the design variables are presented in Table 7. In order to validate the modeling output

results, the operating parameters were compared with the corresponding data from the

15

available literature [29] for the same input parameters. The paper deals with the

examination of a hybrid system consisting of a pre-commercially available high temperature

solid oxide fuel cell and an existing recuperated microturbine. The irreversibilities and

thermodynamic inefficiencies of the system are evaluated after examining the full and

partial load exergetic performance and estimating the amount of exergy destruction and the

efficiency of each hybrid system component. Table 8 compare two groups of data and their

corresponding differences in percentage. Results show that the model was capable of

predicting the thermal performance of the system quite precisely.

Table 7 Comparison between Capstone C30 parameters and proposed design variables in

this study

Table 8 Comparison between ref. [29] and optimized data in this study

There are always things to be improved, but as seen from Table 7 and Table 8, the model

shows a great similarity with the measured values from the real microturbine and the

proposed reference.

4.2. Optimization results

For optimization using genetic algorithm package written in MATLAB software, the

parameters of Table 9 was used and stopping criteria are shown in Table 10 were

considered. The optimal values obtained are presented in Table 11, Table 12. The results

presented in Table 11 are to maximize exergetic efficiency and the results presented in

Table 12 are to maximize net power output. As seen, due to the desired accuracy for

optimization, a range of optimum pressure ratio is presented in both tables.

Table 9 GA parameters

Table 10 Stopping criteria for optimization with GA

Table 11 Optimized values for pressure ratio of the system based on exergetic efficiency

optimization in different TITs

Table 12 Optimized values for design parameters of the system based on maximum net

power generation

As one can see from results:

To optimize the exergetic efficiency, air mass flow differences don’t create large

differences. In other words, after calculating the optimum pressure ratio to achieve

16

maximum exergetic efficiency, due to the required output power the rate of air mass

flow should be calculated.

By increasing the maximum temperature of the cycle, which means turbine inlet

temperature, exergetic efficiency and power output increase. It should be noted that

the maximum temperature of the cycle has its own limitations determined according

to user needs and the investment rationale for the design of the gas turbine.

By increasing the turbine inlet temperature, the optimal amount of pressure ratio for

maximum exergetic efficiency and power output increases.

Since in this model assumed that fuel before entering the combustion chamber,

enters a heat recovery heat exchanger with hot exhaust gases from the turbine,

lower fuel consumption than similar systems are obtained. The reason for the higher

system efficiency of this system than similar systems is this assumption too.

As mentioned above in section 2, energy analysis which is based on the first law of

thermodynamics, does not provide a clear picture of thermodynamic efficiency and losses.

Exergy analysis overcomes these deficiencies and can help identify pathways to sustainable

development. Exergy is a useful tool for determining the location, type and true magnitude

of exergy losses, which appear in the form of either exergy destructions or waste exergy

emissions. Therefore, exergy can assist in developing strategies and guidelines for more

effective use of energy resources and technologies. Figures 4 to 9 shows the variation of

exergy destruction rate with TIT and air mass flow rate of each components of the system

respectively. It is seen that the highest exergy loss takes place at the combustion chamber.

The sources of exergy destruction (or irreversibility) in combustion chamber are mainly the

combustion or chemical reaction and thermal losses in the flow path. Another important

source of exergy loss is the heat exchanger of the system i.e. two recuperators and heat

exchanger, which is related to the big temperature difference between the hot and cold

fluids.

Figure 4 Exergy destruction rate with TIT and air mass flow rate for air compressor

Figure 5 Exergy destruction rate with TIT and air mass flow rate for first recuperator

Figure 6 Exergy destruction rate with TIT and air mass flow rate for fuel compressor

Figure 7 Exergy destruction rate with TIT and air mass flow rate for second recuperator

Figure 8 Exergy destruction rate with TIT and air mass flow rate for combustion chamber

Figure 9 Exergy destruction rate with TIT and air mass flow rate for gas turbine

Figure 10 shows the variation of exergetic efficiency with TIT of each components of the

system respectively. As seen from it, the value of exergetic efficiency of combustion

chamber is lower than that of other components, and can be increased by increasing the

17

combustion inlet temperature (T3) and turbine inlet temperature (T4). However, it should be

noted that due to physical constraints, the turbine material resistance to creep and capital

cost limitations, these temperatures can be changed only within allowable extents. This

means that the improvement of the exergetic efficiency by increasing T3 and T4 may move

the design point from the optimum situation to a new situation at which, the objective

function is not minimum.

Figure 10 Variation of exergetic efficiency with TIT for each component of system

(Compressor pressure ratio is optimized in each TIT (4, 4.75 and 5.4))

4.3. Sensitivity analysis

In this section, to have an understanding of variation of each design parameters on the

objective function a sensitivity analysis has been performed. This analysis is carried out

based on the change in a related parameter as well as some other modeling parameters and

helps us to predict the results while some modifications are necessary in modeling and

optimization. Figure 11 shows the effect of compressor pressure ration and mass flow rate

on system net output power in fixed turbine inlet temperatures. As seen from it variation of

net power output is low sensitive to the compressor pressure ratio. Figures 12 and 13 show

the variation of compressor pressure ratio on the cycle exergetic efficiency. These figures

are proving the ability of desired model to identify the optimum point of the cycle to

achieve maximum exergetic efficiency. For simplicity, in each figure one of abovementioned

TITs is considered.

Figure 11 Variation of compressor pressure ratio and mass flow rate of air on net power

generation (TIT = 1000 ˚K)

Figure 12 Variation of compressor pressure ratio on exergetic efficiency I (without heat

generation) (TIT = 1100 ˚K)

Figure 13 Variation of compressor pressure ratio on exergetic efficiency II (with heat

generation) (TIT = 1200 ˚K)

4.4. Economical results

Figure 14 shows the Pareto frontier solution for CHP system with the objective functions

described in previous sections including exergetic efficiency and total cost rate regarding

environmental aspects. In this figure three TITs are assumed 1000, 1100 and 1200 K. it is

worth to mention that compressor pressure ratio varies around its suggested range resulted

from physical constraints (Table 4).

18

Figure 14 Distribution of Pareto optimal points solutions for exergetic efficiency and total

cost rate of the CHP system

As seen from this figure (for TIT = 1000 K as an example) while the total exergetic efficiency

increases from 67.69% to about 71.52%, the total cost rate increases only slightly from 2.20

to 2.57 ($/hour). In addition, increase in the exergetic efficiency from 71.52% to a little

higher value (71.53%) leads to a drastic increase of the total cost rate from 2.57 to 3.32. This

is corresponding to the moderate increase in the fuel cost rate as a result of increasing air

mass flow rate.

Another conclusion from Figure 14 is that increase in TIT leads to increase in the exergetic

efficiency, however, it results in decrease of the total cost rate first and then this increase

causes a drastic increment in total cost rate of the system. It should be mentioned that in

multiobjective optimization, a process of decision-making for the selection of optimal

solution is necessary. In the Pareto solution, each point can be considered as an optimized

set. Therefore, choosing of the optimum solution depends on preferences and criteria of

each decision-maker. Therefore, they may select a different point as the final optimum

solution which better suits with they requirements.

5. Conclusions

Micro gas turbine engine offers solution to reduce largely the cost and reliability of micro

CHP. A MGT system for combined heat and power generation has been evaluated by means

of system modeling and simulation and optimized for various power outputs (sizes) using

genetic algorithm optimization. The objective functions were selected as the total exergetic

efficiency and the system net power. Some energetic results of the developed model have

been compared with those of literature to indicate its capability at steady-state conditions.

The important parameters of MGT such as fuel consumption, exergetic efficiency and net

power have been analyzed in the wide range of pressure ratio and rate of air mass flow to

have an insight into their influences on exergetic performances of the CHP system. The

results from sensitivity analysis proved the validity of proposed model and also showed that

increasing gas turbine inlet temperature decreases the exergy destruction rate in

combustion chamber (and recuperator) and saves fuel consumption as well. For TIT 1000 K,

the total exergetic efficiency increases from 67.69% to about 71.52% while the total cost

rate increases only slightly from 2.20 to 2.57 ($/hour).

References

1. Sameti, M., Electrical energy efficient building through distributed generation. International Journal of Renewable Energy Research (IJRER), 2014. 4(3): p. 777-783.

2. Sameti, M. and A. Kasaeian. Numerical simulation of combined solar passive heating and radiative cooling for a building. in Building Simulation. 2015. Springer.

19

3. Bahari, S.S., et al., Optimisation of a combined Stirling cycle–organic Rankine cycle using a genetic algorithm. International Journal of Ambient Energy, 2014(ahead-of-print): p. 1-5.

4. Viral R., K.D.K., Optimal planning of distributed generation systems in distribution systems: A review. Renewable and Sustainable Energy Reviews, 2012. 16: p. 5146-5165.

5. Moran A., M.P.J., Chamra L.M., Thermoeconomic modeling of micro-CHP (micro-cooling, heating, and power) for small commercial applications. Energy Res, 2008. 32(808-823).

6. Ferreira E.M., B.J.A.P., Zanardi M.A., Optimization analysis of dual-purpose systems. Desalination, 2010. 250: p. 936-944.

7. Basrawi, F., T. Yamada, and S.y. Obara, Theoretical analysis of performance of a micro gas turbine co/trigeneration system for residential buildings in a tropical region. Energy and Buildings, 2013. 67: p. 108-117.

8. Barbieri, E.S., P.R. Spina, and M. Venturini, Analysis of innovative micro-CHP systems to meet household energy demands. Applied Energy, 2012. 97: p. 723-733.

9. Suárez, I., M. Prieto, and F. Fernández, Analysis of potential energy, economic and environmental savings in residential buildings: Solar collectors combined with microturbines. Applied Energy, 2013. 104: p. 128-136.

10. Ehyaei, M., et al., Feasibility study of applying internal combustion engines in residential buildings by exergy, economic and environmental analysis. Energy and Buildings, 2012. 55: p. 405-413.

11. Tehrani, S.S.M., et al., Development of a CHP/DH system for the new town of Parand: An opportunity to mitigate global warming in Middle East. Applied Thermal Engineering, 2013. 59(1): p. 298-308.

12. Karaali, R. and İ.T. Öztürk, Thermoeconomic optimization of gas turbine cogeneration plants. Energy, 2015.

13. Pepermans G, D.J., Haeseldonckx D, Belmans R, D’haeseleer W. , Distributed generation: definition, benefits and issues. Energy Policy, 2005. 33: p. 787-798.

14. Yazdi, B.A., et al., Optimization of micro CHP gas turbine by genetic algorithm. Thermal Science, 2013(00): p. 141-141.

15. Ehyaei, M. and A. Mozafari, Energy, economic and environmental (3E) analysis of a micro gas turbine employed for on-site combined heat and power production. Energy and Buildings, 2010. 42(2): p. 259-264.

16. Ahmadi, P., I. Dincer, and M.A. Rosen, Thermodynamic modeling and multi-objective evolutionary-based optimization of a new multigeneration energy system. Energy Conversion and Management, 2013. 76: p. 282-300.

17. Angrisani, G., C. Roselli, and M. Sasso, Distributed microtrigeneration systems. Progress in Energy and Combustion Science, 2012. 38(4): p. 502-521.

18. Bade, M.H. and S. Bandyopadhyay, Analysis of gas turbine integrated cogeneration plant: Process integration approach. Applied Thermal Engineering, 2015. 78: p. 118-128.

19. Nikpey, H., M. Assadi, and P. Breuhaus, Development of an optimized artificial neural network model for combined heat and power micro gas turbines. Applied Energy, 2013. 108: p. 137-148.

20

20. Ameri, M., A. Behbahaninia, and A.A. Tanha, Thermodynamic analysis of a tri-generation system based on micro-gas turbine with a steam ejector refrigeration system. energy, 2010. 35(5): p. 2203-2209.

21. Caresana, F., et al., Microturbogas cogeneration systems for distributed generation: Effects of ambient temperature on global performance and components’ behavior. Applied Energy, 2014. 124: p. 17-27.

22. Khaliq, A., Exergy analysis of gas turbine trigeneration system for combined production of power heat and refrigeration. International Journal of Refrigeration, 2009. 32(3): p. 534-545.

23. Barzegar Avval, H., et al., Thermo‐economic‐environmental multiobjective optimization of a gas turbine power plant with preheater using evolutionary algorithm. International Journal of Energy Research, 2011. 35(5): p. 389-403.

24. Arora, J.S., Introduction to optimum design. 1989, New York: McGraw-Hill. 25. Besarati S.M, A.K., Jamali A, Hajiloo A, Nariman-zadeh N., Multi-objective

thermodynamic optimization of combined Brayton and inverse Brayton cycles using genetic algorithms. Energy Conversion and Management, 2010. 51: p. 212-217.

26. EPA, U., Catalog of CHP Technologies. 2008, US EPA, http://www. epa. gov/chp/technologies. html.

27. Scott, W.G., Micro-turbine generators for distribution systems. IEEE Ind Appl Magazine, 1998. 4: p. 57-62.

28. El-Khattam, W. and M. Salama, Distributed generation technologies, definitions and benefits. Electric power systems research, 2004. 71(2): p. 119-128.

29. Bakalis D.P., S.A.G., Full and part load exergetic analysis of a hybrid micro gas turbine fuel cell system based on existing components. Energy Conversion and Management, 2012. 64: p. 213-221.

30. do Nascimento, M.A.R., et al., Micro Gas Turbine Engine: A Review. 2014. 31. Calise F., D.A.M., Vanoli L., Spakovsky M. R., Full load synthesis/design optimization

of a hybrid SOFC_GT power plant. Energy, 2007. 32: p. 446-458. 32. Cengel, Y.A., M.A. Boles, and M. Kanoglu, Thermodynamics: an engineering

approach. Vol. 1056. 1998: McGraw-Hill New York. 33. Bejan, A. and M.J. Moran, Thermal design and optimization. 1996: John Wiley &

Sons. 34. Kotas, T.J., The exergy method of thermal plant analysis. 2013: Elsevier. 35. Kaviri, A.G., M.N.M. Jaafar, and T.M. Lazim, Modeling and multi-objective exergy

based optimization of a combined cycle power plant using a genetic algorithm. Energy Conversion and Management, 2012. 58: p. 94-103.

36. Rosen, M.A. and I. Dincer, Exergoeconomic analysis of power plants operating on various fuels. Applied Thermal Engineering, 2003. 23(6): p. 643-658.

37. Balli, O. and H. Aras, Energetic and exergetic performance evaluation of a combined heat and power system with the micro gas turbine (MGTCHP). International Journal of Energy Research, 2007. 31(14): p. 1425-1440.

38. Dincer, I. and M.A. Rosen, Exergy: energy, environment and sustainable development. 2012: Newnes.

39. Valero, A., et al., CGAM problem: definition and conventional solution. Energy, 1994(19): p. 279-286.

40. Gülder, . ., Flame temperature estimation of conventional and future jet fuels. Journal of engineering for gas turbines and power, 1986. 108(2): p. 376-380.

21

41. Rizk, N.K. and H. Mongia. Semianalytical correlations for NOx, CO, and UHC emissions. in ASME 1992 International Gas Turbine and Aeroengine Congress and Exposition. 1992. American Society of Mechanical Engineers.

42. Chipperfield A.J., F.P.J., Pohlheim H. and Fonseca C.M., Genetic Algorithm Toolbox User's Guide. ACSE Research Report No. 512. 1994: University of Sheffield.

Figure 1 Schematic diagram of the cycle power plant

22

Figure 2 Comparison between thermodynamic properties of material used in this article

based on references [31, 32]

28

38

48

58

68

78

300 400 500 600 700 800 900 1000 1100 1200

Spec

ific

hea

t at

co

nst

ant

pre

ssu

re (

kJ/k

g.K

)

Temperature (˚K)

H2 Cengel H2 Bejan CH4 Cengel CH4 Bejan CO2 Cengel CO2 Bejan

H2O Cengel H2O Bejan N2 Cengel N2 Bejan O2 Cengel O2 Bejan

23

Figure 3 GA flow chart

Coding of parameter

space

Random creation of

initial population

Application of

operators

Population evaluation

Finesses New population

(Replacement of the old)

Is any of

stop criteria

satisfied?

End

No

Yes

Begin

24

Figure 4 Exergy destruction rate with TIT and air mass flow rate for air compressor

Figure 5 Exergy destruction rate with TIT and air mass flow rate for first recuperator

5.8406

7.0087

8.1768

6.3677

7.6412

8.9147

6.7405

8.0886

9.4367

0.25 0.3 0.35

0

1

2

3

4

5

6

7

8

9

10

Air mass flow rate (kg/s)

Eser

gy d

estr

uct

ion

rat

e (k

w/s

)

TIT = 1000 ˚K, rp=4 TIT = 1100 ˚K, rp=4.75 TIT = 1200 ˚K, rp=5.4

8.3355

10.0026

11.6697

9.2051

11.0461

12.8871

10.4597

12.5516

14.6436

0.25 0.3 0.35

0

2

4

6

8

10

12

14

16

Air mass flow rate (kg/s)

Exer

gy d

estr

uct

ion

rat

e (k

W)

TIT = 1000 ˚K, rp=4 TIT = 1100 ˚K, rp=4.75 TIT = 1200 ˚K, rp=5.4

25

Figure 6 Exergy destruction rate with TIT and air mass flow rate for fuel compressor

Figure 7 Exergy destruction rate with TIT and air mass flow rate for second recuperator

0.0819

0.0983

0.1146 0.109

0.1308

0.1525

0.1342

0.161

0.1879

0.25 0.3 0.35

0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

0.18

0.2

Air mass flow rate (kg/s)

Exer

gy d

estr

uct

ion

rat

e (k

W)

TIT = 1000 ˚K, rp=4 TIT = 1100 ˚K, rp=4.75 TIT = 1200 ˚K, rp=5.4

1.0693

1.2831

1.4969

1.2972

1.5567

1.816

1.5486

1.8584

2.1681

0.25 0.3 0.35

0

0.5

1

1.5

2

2.5

Air mass flow rate (kg/s)

Exer

gy d

estr

uct

ion

rat

e (k

W)

TIT = 1000 ˚K, rp=4 TIT = 1100 ˚K, rp=4.75 TIT = 1200 ˚K, rp=5.4

26

Figure 8 Exergy destruction rate with TIT and air mass flow rate for combustion chamber

Figure 9 Exergy destruction rate with TIT and air mass flow rate for gas turbine

42.8323

51.3988

59.9389

48.2726

57.9324

67.5659

53.126

63.7511

74.3763

0.25 0.3 0.35

0

10

20

30

40

50

60

70

80

Air mass flow rate (kg/s)

Exer

gy d

estr

uct

ion

rat

e (k

W)

TIT = 1000 ˚K, rp=4 TIT = 1100 ˚K, rp=4.75 TIT = 1200 ˚K, rp=5.4

4.2766

5.1319

5.9872

4.8722

5.8466

6.821

5.2983

6.358

7.4176

0.25 0.3 0.35

0

1

2

3

4

5

6

7

8

Air mass flow rate (kg/s)

Exer

gy d

estr

uct

ion

rat

e (k

W)

TIT = 1000 ˚K, rp=4 TIT = 1100 ˚K, rp=4.75 TIT = 1200 ˚K, rp=5.4

27

Figure 10 Variation of exergetic efficiency with TIT for each component of system

(Compressor pressure ratio is optimized in each TIT (4, 4.75 and 5.4))

Figure 11 Variation of compressor pressure ratio and mass flow rate of air on net power

generation (TIT = 1000 ˚K)

87.86 88.54 89.01

92.38 92.79 92.84 91.2 91.53 91.86

99.16 99.16 99.14

77.2 78.37

79.45

94.59 95.01 95.39

1000 1100 1200

70

75

80

85

90

95

100

TIT (˚K)

Exer

geti

c ef

fici

ency

(%

)

Air Compressor First Recuperator Fuel Compressor

Second Recuperator Combustion Chamebr Gas Turbine

20

22

24

26

28

30

3.75 3.85 3.95 4.05 4.15 4.25

Net

po

wer

(kW

)

Compressor pressure ratio

Air mass flow rate = 0.25 (Kg/s) Air mass flow rate = 0.3 (Kg/s) Air mass flow rate = 0.35 (Kg/s)

28

Figure 12 Variation of compressor pressure ratio on exergetic efficiency I (without heat

generation) (TIT = 1100 ˚K)

Figure 13 Variation of compressor pressure ratio on exergetic efficiency II (with heat

generation) (TIT = 1200 ˚K)

23.15

23.2

23.25

23.3

23.35

23.4

23.45

23.5

23.55

4.5 4.55 4.6 4.65 4.7 4.75 4.8 4.85 4.9 4.95 5

Exer

geti

c ef

fici

ency

I (%

)

Compressor pressure ratio

71

71.1

71.2

71.3

71.4

71.5

71.6

71.7

71.8

5.15 5.2 5.25 5.3 5.35 5.4 5.45 5.5 5.55 5.6 5.65

Exer

geti

c ef

fici

ency

II (

%)

Compressor pressure ratio

29

Figure 14 Distribution of Pareto optimal points solutions for exergetic efficiency and total

cost rate of the CHP system

Table 1 Capstone C30 design parameters [29, 30]

Electrical power (kW) 28-30

Air mass flow rate (kg/s) 0.31

Net Heat Rate (MJ/kWh) 13.8

Exhaust Gas Flow (Kg/s) 0.31-0.32

Exhaust gas temperature (0C) 275

System efficiency (%) 26 ± 2

Compressor isentropic efficiency (%) 79.6

Turbine isentropic efficiency (%) 84.6

Generator efficiency (%) 95

Table 2 The exergy destruction rate and exergy efficiency equations for plant components

Components Exergy destruction rate Exergetic efficiency

Air Compressor

Recuperator (I) ( ) ( )

Fuel Compressor

Recuperator (II) ( ) ( )

2

2.5

3

3.5

4

67.2 67.7 68.2 68.7 69.2 69.7 70.2 70.7 71.2 71.7

Tota

l co

st r

ate

($/h

ou

r)

Exergetic efficiency (%)

TIT = 1000 ˚K TIT = 1100 ˚K TIT = 1200 ˚K

30

Combustion Chamber

Gas Turbine

Table 3 Important fixed parameters and their values

Description Value

Compressor isentropic efficiency 79.6%

Turbine isentropic efficiency 84%

Recuperator efficiency 90%

Recuperator effectiveness 80%

Invertor efficiency 95%

Generator efficiency 90%

Table 4 Decision variables and their values

Description Unit Min. value Max. value

compressor pressure ratio --- 3 5

turbine inlet temperature C 1000 1200

air mass flow rate Kg/s 0.25 0.35

Table 5 Equation for calculation the purchase costs for the system components

Components Capital or investment cost functions

Air compressor

P P

P P

First recuperator ( ( )

(ΔTLM) )

Fuel compressor

P P

P P

Second recuperator ( ( )

(ΔTLM) )

Combustion chamber

P P ⁄

( ( T ))

Gas turbine

P P ( ( T ))

Heat exchanger (

(ΔTLM) )

Table 6 Constants used in the equation of Table 5 for the purchase cost of the components

31

Components Capital or investment cost functions

Air compressor & fuel compressor ( ⁄ )⁄

Recuperators ⁄ ( )⁄

Combustion chamber ( ⁄ )⁄

Gas turbine ( ⁄ )⁄

Heat exchanger ( ⁄ ) ⁄

( ⁄ )⁄ ( ⁄ ) ⁄

Table 7 Comparison between Capstone C30 parameters and proposed design variables in

this study

Capstone C30 Proposed values in this study

Air mass flow (Kg/s) 0.31 0.31

Compressor pressure ratio 3.6 3.6

System efficiency 26 ± 2 24.87

Net power (kW) 30 33.57

Table 8 Comparison between ref. [29] and optimized data in this study

Optimized data

in ref. [29] Optimized data

in this study

Difference between

values (%)

Air mass flow (Kg/s) 0.307 0.31 ---

Compressor pressure ratio 3.6 3.6 ---

System efficiency 25.1 26.4 ---

Net power (kW) 31.1 33.57 ---

Compressor inlet temperature (K) 298 298.15 0.05

Compressor outlet temperature (K) 462 459.68 0.50

Turbine inlet temperature (K) 1113 1117 0.35

Turbine outlet temperature (K) 885 886 0.11

Turbine exergy destruction (kW) 5.2 4.541 12.67

Turbine exergy efficiency 94 95.55 1.65

Compressor exergy destruction (kW) 7.1 6.815 4.01

Compressor exergy efficiency 86 87.73 2.01

Table 9 GA parameters

Population type Double vector

Population size 20

Creation function Constraint dependent

32

Scaling function Rank

Selection function Stochastic uniform

Elite count 2

Crossover fraction 0.8

Mutation function Constraint dependent

Crossover function Scattered

Migration Direction Forward

Migration Fraction 0.2

Migration Interval 20

Constraint parameters Double vector

Population type 20

Table 10 Stopping criteria for optimization with GA

Generation 100

Time limit 106

Fitness limit 10-6

Stall generations 50

Stall time limit 20

Table 11 Optimized values for pressure ratio of the system based on exergetic efficiency

optimization in different TITs

Tmax = 1000 K Tmax = 1100 K Tmax = 1200 K

Pressure

ratio

Exergetic

efficiency

Pressure

ratio

Exergetic

efficiency

Pressure

ratio

Exergetic

efficiency

3.95 – 4.05 20.1 4.6 – 4.9 23.7 5.2 – 5.6 26.7

Table 12 Optimized values for design parameters of the system based on maximum net

power generation

Tmax = 1000 K Tmax = 1100 K Tmax = 1200 K

Air mass

flow

(Kg/s)

Pressure

ratio

Net

power

(kW)

Air

mass

flow

(Kg/s)

Pressure

ratio

Net

power

(kW)

Air

mass

flow

(Kg/s)

Pressure

ratio

Net

power

(kW)

0.1 4.6 – 5.2 8.6 0.1 5.7 – 6.1 12.3 0.1 7 – 7.5 16.2

0.2 4.6 – 5.2 17.7 0.2 5.7 – 6 25.3 0.2 7.75 32.3

0.3 4.7 – 5.1 26.7 0.3 5.9 - 6 37.9 0.3 7.1 – 7.4 48.2

0.4 4.7 – 5.1 34.8 0.4 5.9 - 6 49.5 0.4 7 – 7.4 61.5

0.5 4.7 – 5.1 42.9 0.5 5.9 - 6 61.7 0.5 7 – 7.4 80.3

33

Graphical abstract

Highlights

A methodology was proposed for design of MGT based CHP systems to be used by

decision makers.

A computer code was developed to simulate the performance of the building

integrated CHP system.

Multiobjective genetic optimization is used for Pareto approach for system

performance.

Suggestions were offered to reduce the overall system irreversibilities.

The thermoenviroeconomic objective and the exergetic efficiency reached their

optimum levels.


Recommended