energy conversion and management - energyville · flow rate on the performance of the power plant...

11
Comparison of shell-and-tube with plate heat exchangers for the use in low-temperature organic Rankine cycles Daniël Walraven a,c , Ben Laenen b,c , William D’haeseleer a,c,a University of Leuven (KU Leuven) Energy Institute, TME Branch (Applied Mechanics and Energy Conversion), Celestijnenlaan 300A Box 2421, B-3001 Leuven, Belgium b Flemish Institute for Technological Research (VITO), Boeretang 200, B-2400 Mol, Belgium c EnergyVille (Joint Venture of VITO and KU Leuven), Dennenstraat 7, B-3600 Genk, Belgium article info Article history: Received 26 March 2014 Accepted 4 July 2014 Keywords: ORC Shell-and-tube heat exchanger Plate heat exchanger System optimization abstract Organic Rankine cycles (ORCs) can be used for electricity production from low-temperature heat sources. These ORCs are often designed based on experience, but this experience will not always lead to the most optimal configuration. The ultimate goal is to design ORCs by performing a system optimization. In such an optimization, the configuration of the components and the cycle parameters (temperatures, pressures, mass flow rate) are optimized together to obtain the optimal configuration of power plant and compo- nents. In this paper, the configuration of plate heat exchangers or shell-and-tube heat exchangers is opti- mized together with the cycle configuration. In this way every heat exchanger has the optimum allocation of heat exchanger surface, pressure drop and pinch-point-temperature difference for the given boundary conditions. ORCs with plate heat exchangers perform mostly better than ORCs with shell-and- tube heat exchangers, but one disadvantage of plate heat exchangers is that the geometry of both sides is the same, which can result in an inefficient heat exchanger. It is also shown that especially the cooling- fluid inlet temperature and mass flow have a strong influence on the performance of the power plant. Ó 2014 Elsevier Ltd. All rights reserved. 1. Introduction Low-temperature geothermal heat sources are widely available [1], but the electricity production efficiency is low due to the low temperature. Many authors have tried to maximize the electricity production efficiency of organic Rankine cycles (ORCs) [2–4] by optimizing the cycle parameters (pressures, temperatures and mass flow rates). To perform such an optimization, simplifying assumptions are made about the components, but these assump- tions can have a strong influence on the performance and the cost of the ORC. As already explained in our previous work [5], this issue is already touched upon in the literature. Madhawa Hettiarachchi et al. [6] minimized the ratio of the total heat-exchanger surface and the net electrical power produced by the cycle, for a fixed heat-exchanger configuration. Quoilin et al. [7] developed an ORC model and plate heat exchanger models based on their experimental set-up, while neglecting the pressure drop in the heat exchangers. They used these models to predict the performance of their set-up in different working conditions. Shengjun et al. [8] maximized the performance of ORCs with shell-and-tube heat exchangers, while keeping the configuration of the heat exchangers fixed. Domingues et al. [9] optimized ORCs with shell-and-tube heat exchangers for vehicle exhaust waste heat recovery and investigated the effect of the number of tubes in the shell-and-tube heat exchangers on the performance of the ORC. Another approach was followed by Franco and Villani [10]. They divided the ORC in a system level and a com- ponent level. In a first step, the system level was optimized, followed by the optimization of the configuration of the components for the obtained optimal system configuration. An iteration between the optimization of both levels was needed to come to the final solution. To obtain the global optimum configuration of the ORC, the sys- tem and the components should be optimized together so that the configuration of the components is optimal for the use in the cycle and so that the components are adjusted to each other. To perform such a system optimization, realistic models, which describe the performance of the components depending on geometric parame- ters, are needed. In this paper models for heat exchangers are implemented and included in the system optimization. Both shell-and-tube heat exchangers and plate heat exchangers are discussed. In previous research [5] we have performed a detailed investigation of shell- http://dx.doi.org/10.1016/j.enconman.2014.07.019 0196-8904/Ó 2014 Elsevier Ltd. All rights reserved. Corresponding author at: University of Leuven (KU Leuven) Energy Institute, TME Branch (Applied Mechanics and Energy Conversion), Celestijnenlaan 300A Box 2421, B-3001 Leuven, Belgium. Tel.: +32 16 32 25 11; fax: +32 16 32 29 85. E-mail addresses: [email protected] (D. Walraven), Ben.Laenen @vito.be (B. Laenen), [email protected] (W. D’haeseleer). Energy Conversion and Management 87 (2014) 227–237 Contents lists available at ScienceDirect Energy Conversion and Management journal homepage: www.elsevier.com/locate/enconman

Upload: duongtu

Post on 21-Jul-2018

214 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Energy Conversion and Management - EnergyVille · flow rate on the performance of the power plant are also investi- ... c and the baffle spacing at the inlet L b;i, ... chevron,

Energy Conversion and Management 87 (2014) 227–237

Contents lists available at ScienceDirect

Energy Conversion and Management

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

Comparison of shell-and-tube with plate heat exchangersfor the use in low-temperature organic Rankine cycles

http://dx.doi.org/10.1016/j.enconman.2014.07.0190196-8904/� 2014 Elsevier Ltd. All rights reserved.

⇑ Corresponding author at: University of Leuven (KU Leuven) Energy Institute,TME Branch (Applied Mechanics and Energy Conversion), Celestijnenlaan 300ABox 2421, B-3001 Leuven, Belgium. Tel.: +32 16 32 25 11; fax: +32 16 32 29 85.

E-mail addresses: [email protected] (D. Walraven), [email protected] (B. Laenen), [email protected] (W. D’haeseleer).

Daniël Walraven a,c, Ben Laenen b,c, William D’haeseleer a,c,⇑a University of Leuven (KU Leuven) Energy Institute, TME Branch (Applied Mechanics and Energy Conversion), Celestijnenlaan 300A Box 2421, B-3001 Leuven, Belgiumb Flemish Institute for Technological Research (VITO), Boeretang 200, B-2400 Mol, Belgiumc EnergyVille (Joint Venture of VITO and KU Leuven), Dennenstraat 7, B-3600 Genk, Belgium

a r t i c l e i n f o

Article history:Received 26 March 2014Accepted 4 July 2014

Keywords:ORCShell-and-tube heat exchangerPlate heat exchangerSystem optimization

a b s t r a c t

Organic Rankine cycles (ORCs) can be used for electricity production from low-temperature heat sources.These ORCs are often designed based on experience, but this experience will not always lead to the mostoptimal configuration. The ultimate goal is to design ORCs by performing a system optimization. In suchan optimization, the configuration of the components and the cycle parameters (temperatures, pressures,mass flow rate) are optimized together to obtain the optimal configuration of power plant and compo-nents. In this paper, the configuration of plate heat exchangers or shell-and-tube heat exchangers is opti-mized together with the cycle configuration. In this way every heat exchanger has the optimumallocation of heat exchanger surface, pressure drop and pinch-point-temperature difference for the givenboundary conditions. ORCs with plate heat exchangers perform mostly better than ORCs with shell-and-tube heat exchangers, but one disadvantage of plate heat exchangers is that the geometry of both sides isthe same, which can result in an inefficient heat exchanger. It is also shown that especially the cooling-fluid inlet temperature and mass flow have a strong influence on the performance of the power plant.

� 2014 Elsevier Ltd. All rights reserved.

1. Introduction

Low-temperature geothermal heat sources are widely available[1], but the electricity production efficiency is low due to the lowtemperature. Many authors have tried to maximize the electricityproduction efficiency of organic Rankine cycles (ORCs) [2–4] byoptimizing the cycle parameters (pressures, temperatures andmass flow rates). To perform such an optimization, simplifyingassumptions are made about the components, but these assump-tions can have a strong influence on the performance and the costof the ORC.

As already explained in our previous work [5], this issue isalready touched upon in the literature. Madhawa Hettiarachchiet al. [6] minimized the ratio of the total heat-exchanger surfaceand the net electrical power produced by the cycle, for a fixedheat-exchanger configuration. Quoilin et al. [7] developed an ORCmodel and plate heat exchanger models based on their experimentalset-up, while neglecting the pressure drop in the heat exchangers.

They used these models to predict the performance of their set-upin different working conditions. Shengjun et al. [8] maximized theperformance of ORCs with shell-and-tube heat exchangers, whilekeeping the configuration of the heat exchangers fixed. Domingueset al. [9] optimized ORCs with shell-and-tube heat exchangers forvehicle exhaust waste heat recovery and investigated the effect ofthe number of tubes in the shell-and-tube heat exchangers on theperformance of the ORC. Another approach was followed by Francoand Villani [10]. They divided the ORC in a system level and a com-ponent level. In a first step, the system level was optimized, followedby the optimization of the configuration of the components for theobtained optimal system configuration. An iteration between theoptimization of both levels was needed to come to the final solution.

To obtain the global optimum configuration of the ORC, the sys-tem and the components should be optimized together so that theconfiguration of the components is optimal for the use in the cycleand so that the components are adjusted to each other. To performsuch a system optimization, realistic models, which describe theperformance of the components depending on geometric parame-ters, are needed.

In this paper models for heat exchangers are implemented andincluded in the system optimization. Both shell-and-tube heatexchangers and plate heat exchangers are discussed. In previousresearch [5] we have performed a detailed investigation of shell-

Page 2: Energy Conversion and Management - EnergyVille · flow rate on the performance of the power plant are also investi- ... c and the baffle spacing at the inlet L b;i, ... chevron,

Nomenclature

Greekb corrugation angle (�)Dp pressure drop (Pa)DT temperature difference (�C)g efficiency (–)K corrugation width (m)l dynamic viscosity (Pa s)U area enlargement factor (–)q density (kg/m3)h angle (�)

Romana corrugation amplitude (m)A area (m2)Bo boiling number (–)cp specific heat capacity (J/kg K)do tube outside diameter (m)D diameter (m)e specific exergy (kJ/kg)G mass velocity (kg/m2 s)h heat transfer coefficient (W/m2 K)h specific enthalpy (J/kg)lc baffle cut length (m)Lb distance between baffles (m)Lh length of a plate (m)Lp distance between ports (m)_m mass flow (kg/s)

Nu Nusselt number (–)p pressure (bar)Pr Prandtl number (–)pt tube pitch (m)_Q heat flow (kW)

Re Reynolds number (–)T temperature (�C)W plate width (m)_W mechanical power (kW)

X corrugation parameter (–)

Sub- and superscripts0 dead state1� 9 number of the stateac accelerationcycle cycleen energeticex exergeticfr frictionalh hydraulicid idealin inletmax maximummin minimumnet nettl longitudinalotl outermost tubesout outletplant plants shellsource heat sourcet transversetot totalwf working fluid

228 D. Walraven et al. / Energy Conversion and Management 87 (2014) 227–237

and-tube heat exchangers integrated in ORCs. The shell-and-tubeheat exchangers are modeled with the Bell-Delaware method[11,12], which is a mature model and can be used for single-phaseflow, condensing and evaporation.

The purpose of our research reported in this paper is to inte-grate plate heat exchangers in ORCs and to compare the result tothat obtained with shell-and-tube heat exchangers. Martin [13]developed a model for plate heat exchangers with single-phaseflow. This model is based on physical reasoning and many experi-mental data are used. Such generally applicable models do notexist for two-phase plate heat exchangers used as evaporators orcondensers, although much research has been performed on thetopic [14–20]. The authors of those references propose correlationsfor heat-transfer coefficient and pressure drop based on ownexperiments and these correlations are therefore only valid forthe investigated cases. We shall utilize the correlations of Hanet al. [16,17] for evaporation and condensation in plate heatexchangers, respectively. These papers correlate the performanceof plate heat exchangers to many geometrical parameters.

This paper extends the work performed in Walraven et al. [5], inwhich ORCs with shell-and-tube heat exchangers are optimized fora reference case. The conclusion from that work is that it is optimalto use the 30� and 60� tube configurations in single and two-phaseheat exchangers, respectively. In this paper, models for plate heatexchangers are added and ORCs with plate heat exchangers arecompared to ORCs with shell-and-tube heat exchangers (with theoptimal tube configuration). The influence of the heat-source-inlettemperature, heat-source-outlet temperature, total heat exchangersurface, cooling-fluid inlet temperature and the cooling fluid mass

flow rate on the performance of the power plant are also investi-gated. The comparison between ORCs with the two different typesof heat exchangers is performed in a wide range of parameters andfor many fluids.

2. Organic Rankine cycle

Different types of ORCs exist and are simulated in this paper.The investigated cycles can be of the simple or recuperated type,be subcritical or transcritical and can have one or two pressurelevels. Two examples are given in Fig. 1, in which the scheme ofa single-pressure, recuperated ORC and a double-pressure, simpleORC are shown. All the possible heat exchangers (economizer,evaporator, superheater, desuperheater, condenser and recupera-tor) are shown in the figure, but are not always necessary.

In all configurations it is assumed that state 1 is saturated liquidand that the isentropic efficiencies of the pump and turbine are80% and 85%, respectively. More information of the modeling canbe found in previous work [4,5]. Instead of assuming a fixed pinchpoint temperature difference and ideal heat exchangers, modelsare used to calculate the heat transfer coefficients and pressuredrops in each heat exchanger.

3. Shell-and-tube heat exchanger

The shell-and-tube heat exchanger type has already been stud-ied in Walraven et al. [5]. Here we now recall some elements tosupport the later analysis with plate heat exchangers. TEMA E type

Page 3: Energy Conversion and Management - EnergyVille · flow rate on the performance of the power plant are also investi- ... c and the baffle spacing at the inlet L b;i, ... chevron,

Cooling outCooling in

Heat source inHeat source out

1

2

3

4 56

7

8

9

(a) Single-pressure

Cooling outCooling in

Heat source in

Hea

tso

urce

out

1 9 7

2 4b 5b 6b

1a2a 4a 5a 6a

(b) Double-pressure

Fig. 1. Scheme of a single-pressure, recuperated (a) and double-pressure, simple (b) ORC.

Lb,cLb,i

Lb,oD

s

lc

pt

Dotl

θb

θctl

do

Dctl

Fig. 2. Shell-and-tube geometrical characteristics. Figure adapted from Shah and Sekulic [12]. See also Walraven et al. [5].

1 These parts generally do not have the same physical size.2 This is a starting value. The effect of the pressure drop on the heat load is taken

into account in the last part of the heat exchanger.

D. Walraven et al. / Energy Conversion and Management 87 (2014) 227–237 229

heat exchangers with a single shell pass and with the inlet and theoutlet at the opposite ends of the shell are modeled. The workingfluid always flows on the shell side, so models for the pressure dropand heat-transfer coefficient in single-phase flow, evaporation andcondensation in a TEMA E shell are needed. The tube-side fluid (theheat source and the cooling fluid) will always be single phase.

Fig. 2 recalls the basic geometrical characteristics of a shell-and-tube heat exchanger. These are the shell outside diameter Ds, theoutside diameter of a tube do, the pitch between the tubes pt , thebaffle cut length lc and the baffle spacing at the inlet Lb;i, outletLb;o and the center Lb;c. More detailed information about theshell-and-tube model used in this paper can be found in Walravenet al. [5].

4. Plate heat exchanger

Plate heat exchangers can have many different types of corru-gations [12], but in this paper only heat exchangers withchevron, also known as herringbone, corrugations are used. Thistype of corrugation is commonly used and models whichdescribe the pressure drop and heat transfer depending on theheat exchanger geometry are available [13,16,17]. The numberof passes on both sides of the heat exchangers are assumed tobe equal in this paper.

4.1. Geometry

Fig. 3 shows the geometrical parameters of a chevron plate. Thecorrugations are determined by the corrugation amplitude a, thecorrugation width K and the angle of the corrugations b. The widthof a plate W, the length of a plate between ports Lp and the lengthof a plate for heat transfer Lh are also indicated.

The hydraulic diameter is defined as:

Dh ¼4aU; ð1Þ

where U ¼ 16 1þ

ffiffiffiffiffiffiffiffiffiffiffiffiffiffi1þ X2

pþ 4

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1þ X2=2

q� �is the area-enlargement

factor and X ¼ 2paK the dimensionless corrugation parameter.

4.2. Heat-transfer and pressure-drop correlations

The correlations of Martin [13] are used to predict the heat-transfer coefficient and the pressure drop in the single-phase heatexchangers. For the evaporator and the condenser, the correlationsof Han et al. [16,17] are used, respectively. An overview of thesecorrelations is given in Appendix A.

4.3. Implementation of the models

To account for non-uniform fluid properties, each heat exchan-ger is divided into five parts with an equal heat load.1 This numberis chosen, because it leads to a reasonable accuracy and calculationtime. For each heat exchanger, the configuration (a; K; b andWNp), the inlet states at one side of the heat exchanger and a neces-sary outlet condition (e.g. the working fluid has to be saturated vaporat the end of the evaporator) are needed. The variables W and Np arecombined to one variable, because only their product appears in theequations. With these data, the total heat that has to be transferredin the case of no pressure drop2 can be calculated. In each of the fiveparts one fifth of this total heat will be exchanged. With the

Page 4: Energy Conversion and Management - EnergyVille · flow rate on the performance of the power plant are also investi- ... c and the baffle spacing at the inlet L b;i, ... chevron,

Fig. 3. Geometrical parameters of a chevron plate. Figure from Shah and Sekulic [12].

Table 1Optimization variables and constraints used for plate-type heat exchangers and theirlower and upper boundaries.

Optimization variable Lowerboundary

Upperboundary

Corrugation amplitude a 1 mm 200 mmCorrugation width K 1 mm 200 mmCorrugation angle b 0� 90�Product of plate width and number of

channels WNp

10 mm 10 m

Ratio of corrugation width to corrugationamplitude K=a

2.5 3.5

230 D. Walraven et al. / Energy Conversion and Management 87 (2014) 227–237

equations above, the heat-transfer coefficient and the pressure dropin the first part can be calculated. In this way, the state after the firstpart, the necessary heat-transfer surface and the fictive plate lengthof the first part can be calculated. This procedure is repeated for theother parts, except in the last part for which the heat to be trans-ferred is corrected for the pressure drop in the previous parts.

3 Atot is a non-linear function of the optimization variables.

5. Optimization

5.1. Objective function

The goal of the optimization is to find a system configurationwhich maximizes the mechanical work output for a given heatsource. This is the same as maximizing the exergetic plant effi-ciency [4], defined as:

gplantex ¼

_Wnet

_msourceesourcein

; ð2Þ

with _Wnet the net mechanical-power production of the power plant,_msource the mass flow of the source and esource

in the flow exergy of thesource.

5.2. Optimization variables and constraints

The optimization variables which determine the cycle configu-ration in a single-pressure, simple cycle are the temperature andsaturation temperature at the pressure before the turbine T6 andTsatðp6Þ, the mass flow of working fluid _mwf and the temperatureafter the condenser T1. For a recuperated cycle, the temperaturedifference between states 2 and 8 is added as an optimization var-iable. For a double-pressure cycle, Tb

6; Tsatðpb

6Þ and _mbwf are added as

optimization variables. More information can be found in Walrav-en et al. [5].

The optimization variables and constraints used for the plate-type heat exchangers are given in Table 1.

If no constraint on the heat-exchanger surface of each heatexchanger is imposed, the pinch-point-temperature differenceswould become very small and the total heat exchanger surfacewould become huge. Therefore, a non-linear constraint3 on thetotal heat exchanger surface of all heat exchangers together Atot isimposed: Atot

6 Atotmax. In this way, the optimizer can choose itself

how to distribute the available surface optimally amongst the differ-ent heat exchangers.

A last constraint is a limit on the heat source outlet tempera-ture. In some circumstances, the heat source outlet temperaturecannot be too low, e.g. to use the heat source for heating or to avoidscaling with geothermal brines. This is again a non-linear con-straint: Tsource

out P Tsourcemin .

Table 2 recalls the optimization variables and constraints forthe shell-and-tube case, analyzed in Walraven et al. [5].

5.3. Optimization approach

The cycle configuration is optimized together with the configu-ration of the heat exchangers, as done in Walraven et al. [5]. Thismeans that all the above mentioned optimization variables canbe adapted at every iteration step by the optimizer in order tomaximize the exergetic plant efficiency. Fluid properties areobtained from REFPROP [21] and the complex-step derivativemethod [22] is used to obtain the derivative of the fluid properties.These properties are available in Python, in which our model iswritten, by using F2PY [23]. The CasADi software [24] calculatesthe gradient of the objective function and constraints with auto-matic differentiation in reverse mode and these gradients are usedby the optimizer WORHP [25] together with the value of the objec-tive function to find the optimum.

6. Results

6.1. Reference parameters

100 kg/s of clean water is used as the heat source. The parame-ters for the reference case are given in Table 3. The influence ofthese parameters is investigated in the following sections. Thetemperature of the heat source after the ORC is unconstrained

Page 5: Energy Conversion and Management - EnergyVille · flow rate on the performance of the power plant are also investi- ... c and the baffle spacing at the inlet L b;i, ... chevron,

Table 2Optimization variables and constraints used for shell-and-tube heat exchangers andtheir lower and upper boundaries.

Optimization variable Lowerboundary

Upperboundary

Shell diameter Ds 0.3 m 2 mTube outside diameter do 5 mm 50 mmRelative tube pitch pt=do 1.15 2.5Relative baffle cut lc=Ds 0.25 0.45Baffle spacing Lb;c 0.3 m 5 mRatio of tube diameter to shell diameter

do=Ds

/ 0.1

Table 3Reference parameters.

Parameter Symbol Value

Heat source inlet temperature Tsourcein 125 �C

Heat source mass flow _msource 100 kg/sMaximum allowed heat exchanger surface Atot

max4000 m2

Cooling fluid inlet temperature Tcoolingin

20 �C

Cooling fluid mass flow _mcooling 800 kg/s

D. Walraven et al. / Energy Conversion and Management 87 (2014) 227–237 231

and the ORC is of the simple, single-pressure type. In a previouswork [5] it was shown that it is best to use a 30�- and 60�-tube con-figuration in single-phase and two-phase shell-and-tube heatexchangers, respectively. This combined configuration is used forthe comparison in this paper.

2,000 4,000 6,000 8,000 10,000

20

30

40

50

Maximum allowed surface [m2]

pla

teη

pla

nt

ex[%

]

Isobutane PropaneR134a R218R227ea R245faR1234yf RC318

(a) Plate heat exchangers

Fig. 4. Exergetic plant efficiency for single-pressure, simple ORCs with all plate hea

2,000 4,000 6,000 8,000 10,0006

8

10

12

Maximum allowed surface [m2]

pla

teη

cycl

een

[%]

Isobutane PropaneR134a R218R227ea R245faR1234yf RC318

(a) Plate heat exchangers

Fig. 5. Energetic cycle efficiency for single-pressure, simple ORCs with all plate hea

6.2. Influence maximum allowed heat-exchanger surface

In this section the influence of the maximum allowed heatexchanger surface is investigated. The other parameters are theones given in Table 3. Fig. 4 shows the exergetic plant efficiencyfor simple ORCs with all plate heat exchangers or all shell-and-tubeheat exchangers for different working fluids as a function of themaximum allowed heat exchanger surface. The efficiency increaseswith increasing Atot

max, as expected.This increase of the exergetic plant efficiency with increasing

maximum allowed surface is explained by the increase of the ener-getic cycle efficiency (Fig. 5) and the decrease of the heat-sourceoutlet temperature (Fig. 6). So, more heat is added to the cycleand this heat is converted more efficiently to mechanical powerwhen the total heat-exchanger surface increases. This cycle effi-ciency is defined as:

gcycleen ¼

_Wnet

_Q; ð3Þ

with _Q the heat added to the cycle.For low values of Atot

max ORCs with all plate heat exchangers pro-duce more electrical power than ORCs with all shell-and-tube heatexchangers. It is generally known that plate heat exchangers canachieve higher heat-transfer coefficients and therefore lowerpinch-point-temperature differences as seen in Fig. 7. For highvalues of Atot

max ORCs with all shell-and-tube heat exchangersperform equally well or even better. The pinch-point temperaturedifferences in Fig. 7 become very small and the efficiency of theORCs almost reaches the upper limit.

2,000 4,000 6,000 8,000 10,000

20

30

40

50

Maximum allowed surface [m2]

shel

l ηpla

nt

ex[%

]

Isobutane PropaneR134a R218R227ea R245faR1234yf RC318

(b) Shell-and-tube heat exchangers

t exchangers (a) and all shell-and-tube heat exchangers (b) for different fluids.

2,000 4,000 6,000 8,000 10,0006

8

10

12

Maximum allowed surface [m2]

shel

l ηcy

cle

en[%

]

Isobutane PropaneR134a R218R227ea R245faR1234yf RC318

(b) Shell-and-tube heat exchangers

t exchangers (a) and all shell-and-tube heat exchangers (b) for different fluids.

Page 6: Energy Conversion and Management - EnergyVille · flow rate on the performance of the power plant are also investi- ... c and the baffle spacing at the inlet L b;i, ... chevron,

232 D. Walraven et al. / Energy Conversion and Management 87 (2014) 227–237

For the working fluids R245fa, isobutane and RC318 the exer-getic plant efficiency of ORCs with all shell-and-tube heat exchang-ers is better than the one for ORCs with all plate heat exchangersfor Atot

max > 2000; 3000 and 6000 m2, respectively. Comparison ofFig. 7a and b shows that the pinch-point temperature differencebetween the working fluid and cooling fluid DTLow

min decreases muchfaster in the case of shell-and-tube heat exchangers with increas-ing Amax. This is due to fact that the cold and the hot side of plateheat exchangers with the same number of passes at both sideshave exactly the same geometry, which is of course not the casefor shell-and-tube heat exchangers. The mass flow of the cooling

2,000 4,000 6,000 8,000 10,00040

50

60

70

80

90

Maximum allowed surface [m2]

pla

teT

out[◦ C

]

Isobutane PropaneR134a R218R227ea R245faR1234yf RC318

(a) Plate heat exchangers

Fig. 6. Outlet temperature of the heating fluid for single-pressure, simple ORCs with alfluids.

2,000 4,000 6,000 8,000 10,0000

5

10

15

Maximum allowed surface [m2]

pla

teΔ

TL

ow

min

[◦ C]

Isobutane PropaneR134a R218R227ea R245faR1234yf RC318

(a) Plate heat exchangers, ΔTmin between work-ing fluid and cooling fluid

2,000 4,000 6,000 8,000 10,0000

5

10

15

Maximum allowed surface [m2]

pla

teΔ

TH

igh

min

[◦ C]

Isobutane PropaneR134a R218R227ea R245faR1234yf RC318

(c) Plate heat exchangers, ΔTmin between heat-ing fluid and working fluid

Fig. 7. Minimum temperature difference between working fluid and cooling fluid (a and(c and d) for single-pressure, simple ORCs with all plate heat exchangers (a and c) and

water is much higher than the one of the working fluid, whichcan result in a large difference between the optimal geometry ofthe cold side and the one of the hot side in a plate heat exchanger,leading to relatively high pinch-point-temperature differences.

It can also be noticed from Fig. 7 that DTHighmin is lower than DTLow

min

for subcritical cycles (e.g. Isobutane, R245fa), while they are aboutequal for transcritical cycles (e.g. R227ea, R1234yf). Because of thehigh cooling fluid mass flow, DTLow

min is relatively close to the averagetemperature difference, which applies also to DTHigh

min in transcriticalcycles. In subcritical cycles, DTHigh

min is much lower than the averagetemperature difference.

2,000 4,000 6,000 8,000 10,00040

50

60

70

80

90

Maximum allowed surface [m2]

shel

l Tout[◦ C

]

Isobutane PropaneR134a R218R227ea R245faR1234yf RC318

(b) Shell-and-tube heat exchangers

l plate heat exchangers (a) and all shell-and-tube heat exchangers (b) for different

2,000 4,000 6,000 8,000 10,0000

5

10

15

Maximum allowed surface [m2]

shel

l ΔT

Low

min

[◦ C]

Isobutane PropaneR134a R218R227ea R245faR1234yf RC318

(b) Shell-and-tube heat exchangers, ΔTmin be-tween working fluid and cooling fluid

2,000 4,000 6,000 8,000 10,0000

5

10

15

Maximum allowed surface [m2]

shel

l ΔT

Hig

hm

in[◦ C

]

Isobutane PropaneR134a R218R227ea R245faR1234yf RC318

(d) Shell-and-tube heat exchangers, ΔTmin be-tween heating fluid and working fluid

b) and minimum temperature difference between heating fluid and working fluidall shell-and-tube heat exchangers (b and d) for different fluids.

Page 7: Energy Conversion and Management - EnergyVille · flow rate on the performance of the power plant are also investi- ... c and the baffle spacing at the inlet L b;i, ... chevron,

100 110 120 130 140 150

30

35

40

45

Heat source inlet temperature [◦C]

pla

teη

pla

nt

ex[%

]

Isobutane Propane R134aR218 R227ea R245faR1234yf RC318

(a) Plate heat exchangers

100 110 120 130 140 150

30

35

40

45

Heat source inlet temperature [◦C]

shel

l ηpla

nt

ex[%

]

Isobutane Propane R134aR218 R227ea R245faR1234yf RC318

(b) Shell-and-tube heat exchangers

Fig. 8. Exergetic plant efficiency for single-pressure, simple ORCs with all plate heat exchangers (a) and all shell-and-tube heat exchangers (b) for different fluids.The parameters are those of Table 3.

40 50 60 70 80 90

100

110

120

130

Heat source outlet temperature [◦C]

pla

terec

upη

pla

nt

ex/p

late

standardη

pla

nt

ex[%

]

Isobutane PropaneR134a R218R227ea R245faR1234yf RC318

(a) Simple cycle

40 50 60 70 80 90

100

110

120

130

Heat source outlet temperature [◦C]

shel

lrec

upη

pla

nt

ex/sh

ell

standardη

pla

nt

ex[%

]

Isobutane PropaneR134a R218R227ea R245faR1234yf RC318

(b) Recuperated cycle

Fig. 9. Exergetic plant efficiency for single-pressure, simple and recuperated ORCs with all plate heat exchangers (a) and all shell-and-tube heat exchangers (b) for differentfluids. Other parameters are those of Table 3.

10 15 20 25 30 35 4020

30

40

50

Cooling fluid inlet temperature [◦C]

pla

teη

pla

nt

ex[%

]

Isobutane PropaneR134a R218R227ea R245faR1234yf RC318

(a) Exergetic plant efficiency

10 15 20 25 30 35 400

2

4

6

8

Cooling fluid inlet temperature [◦C]

pla

teΔ

TL

ow

min

[◦ C]

Isobutane Propane R134aR218 R227ea R245faR1234yf RC318

(b) ΔTmin between working fluid and cooling

fluid

Fig. 10. Exergetic plant efficiency (a) and DTLowmin (b) for single-pressure, simple ORCs with all plate heat exchangers for different fluids. Other parameters are those of Table 3.

D. Walraven et al. / Energy Conversion and Management 87 (2014) 227–237 233

6.3. Influence heat-source inlet temperature

In this section the heat-source inlet temperature is variedbetween 100 and 150 �C, while keeping the other parameters con-stant as given in Table 3. It is seen from Fig. 8 that every fluid per-forms optimally for a certain heat source-inlet temperature.4 ORCs

4 For some fluids this optimal temperature is outside the range of the figure.

with all shell-and-tube heat exchangers have their maximum plantefficiency at a lower heat-source inlet temperature than plate heatexchangers. A heat source with an higher inlet temperature containsmore energy and more energy will be transferred in the heatexchangers. However, for a fixed maximum allowed heat-exchangersurface, the pinch-point-temperature differences will increase withincreasing heat-source inlet temperature. This increase will in mostcases be faster if shell-and-tube heat exchangers are used instead ofplate heat exchangers. Higher pinch-point-temperature differences

Page 8: Energy Conversion and Management - EnergyVille · flow rate on the performance of the power plant are also investi- ... c and the baffle spacing at the inlet L b;i, ... chevron,

234 D. Walraven et al. / Energy Conversion and Management 87 (2014) 227–237

have of course a negative influence on the exergetic plant efficiencyand this efficiency will therefore start to decrease earlier when shell-and-tube heat exchangers are used instead of plate heat exchangers.

6.4. Influence heat-source outlet temperature

Often the heat source cannot be cooled down too much; e.g. toavoid scaling in geothermal brines or when the heat source afterthe ORC is used for heating. In this section, the heat-source outlettemperature will be limited to values between 40 and 90 �C. Theother values are the same as in the reference case (Table 3). Bothsimple and recuperated, single-pressure cycles are investigated.

Fig. 9 shows the comparison between simple and recuperatedcycles. When the heat-source outlet temperature is limited to a rel-atively low temperature, both type of cycles perform equally wellas explained in Walraven et al. [4]. When the limit is set at a highertemperature, the recuperated cycles perform better than thesimple ones. The internal heat recuperation increases the cycleefficiency, which leads to a higher plant efficiency for a fixedheat-source outlet temperature.

6.5. Influence cooling-fluid inlet temperature

The cooling-fluid inlet temperature is varied between 10 and40 �C, while keeping the other parameters constant (Table 3). Onlythe results for ORCs with all plate heat exchangers are given,because the results in the case of all shell-and-tube heat exchang-ers show the same trend. The exergetic plant efficiency decreaseslinearly as shown in Fig. 10a, but the decrease is less strong thanin the case when ideal components – the results of the ideal caseare shown in Walraven et al. [4] – are used. Fig. 10b shows the

200 500 800 1,100 1,400 1,700 2,00025

30

35

40

45

Cooling fluid mass flow [kg/s]

pla

teη

pla

nt

ex[%

]

Isobutane PropaneR134a R218R227ea R245faR1234yf RC318

(a) Exergetic plant efficiency

200 500 800 1,100 1,400 1,700 2,0000

2

4

6

8

Cooling fluid mass flow [kg/s]

pla

teΔ

TL

ow

min

[◦ C]

Isobutane PropaneR134a R218R227ea R245faR1234yf RC318

(c) Pinch-point-temperature difference in con-

denser

Fig. 11. Exergetic plant efficiency (a), minimum temperature in the condenser (b) and Dfluids.

pinch-point-temperature difference between the working fluidand the cooling fluid. This temperature difference decreases withincreasing cooling–water inlet temperature, so that the condensertemperature remains relatively low.

6.6. Influence cooling-fluid mass flow

In this section the cooling-water mass flow is varied between200 and 2000 kg/s. The other parameters are given in Table 3and only the results for ORCs with all plate heat exchangers areshown; the results for ORCs with all shell-and-tube heat exchang-ers are analogous. Fig. 11a shows the exergetic plant efficiency as afunction of the cooling-water mass flow. At low values of the cool-ing-water mass flow, an increase of this mass flow will lead to anincrease in the plant efficiency. This efficiency reaches a maximumfor a mass flow of about 800 kg/s for ORCs with Propane and for amass flow of about 1100 kg/s for ORCs with Isobutane, R134a, R218and RC318. For the other investigated fluids the maximum isreached at higher mass flows. The same effect is also seen whenshell-and-tube heat exchangers are used, but it is then less strong.The maximum in the exergetic plant efficiency at a certain massflow rate is explained by Fig. 11b, which shows the minimum tem-perature in the condenser. This temperature decreases withincreasing cooling water mass flow, which results in a better plantefficiency. The pressure drop in the cooling water increases on theother hand. The combination of the increasing pressure drop andincreasing mass flow results in an increasing pumping power. Thiseffect becomes more important than the decreasing condensertemperature, which results in a maximum plant efficiency for acertain cooling fluid mass flow rate.

200 500 800 1,100 1,400 1,700 2,00025

30

35

40

45

Cooling fluid mass flow [kg/s]

pla

teT

cond

[◦ C]

Isobutane PropaneR134a R218R227ea R245faR1234yf RC318

(b) Condenser temperature

TLowmin (c) for single-pressure, simple ORCs with all plate heat exchangers for different

Page 9: Energy Conversion and Management - EnergyVille · flow rate on the performance of the power plant are also investi- ... c and the baffle spacing at the inlet L b;i, ... chevron,

Isob

utan

e

Pro

pane

R13

4a

R21

8

R22

7ea

R24

5fa

R12

34yf

RC

318

30

35

40

45

ηpla

nt

ex[%

]

1 pressure shell 2 pressures shell1 pressure plate 2 pressures plate

(a) Exergetic plant efficiency

Isob

utan

e

Pro

pane

R13

4a

R21

8

R22

7ea

R24

5fa

R12

34yf

RC

318

40

50

60

70

Tout[◦ C

]

1 pressure shell 2 pressures shell1 pressure plate 2 pressures plate

(b) Outlet temperature

Fig. 12. Exergetic plant efficiency (a) and heat source outlet temperature (b) for simple ORCs with all plate heat exchangers or all shell-and-tube heat exchangers for differentfluids.

D. Walraven et al. / Energy Conversion and Management 87 (2014) 227–237 235

Fig. 11c shows the pinch-point-temperature difference in thecondenser. If the cooling fluid mass flow rate increases, the cool-ing-fluid outlet temperature will decrease and the curve of thecooling fluid will become flatter in a temperature-heat diagram.In order to keep the average temperature difference in the con-denser more or less constant (the total amount of available heatexchanger surface is limited), the pinch-point-temperature differ-ence in the condenser has to increase.

6.7. Influence number of pressure levels

The effect on the plant efficiency of adding extra pressure levelsis shown in Fig. 12a. The reference parameters are used (Table 3)and the results for both plate heat exchangers and shell-and-tubeheat exchangers are shown.

Adding an extra pressure level to subcritical cycles (Isobutane,Propane, R134a, R245fa) has a positive effect on the plant perfor-mance. Transcritical cycles do not use the second pressure leveland their plant efficiency remains therefore constant. Double-pressure, subcritical cycles can cool down the heat source morethan single-pressure subcritical cycles (Fig. 12b) and obtain betterplant efficiencies, while the energetic cycle efficiency remainsabout constant when adding a second pressure level.

7. Conclusions

The system optimization of different configurations of ORCswith both plate heat exchangers and shell-and-tube heat exchang-ers is compared in this paper. Models for heat exchangers used insingle-phase flow, evaporation and condensation which are avail-able in the literature are implemented and added to a previousdeveloped ORC-model.

It is shown that ORCs with all plate heat exchangers performmostly better than ORCs with all shell-and-tube heat exchangers.The disadvantage of plate heat exchangers with an equal numberof passes at both sides of the exchanger is that the geometry ofboth sides of the heat exchanger are identical, which can lead toan inefficient heat exchanger when the two fluid streams requirestrongly different channel geometries.

The influence of the maximum allowed heat-exchanger surface,heat-source inlet temperature, constraint on the heat-source outlettemperature, cooling-fluid inlet temperature and mass flow andthe number of pressure levels on the performance of the ORC havebeen investigated. It is shown that the efficiency of an ORCincreases with increasing heat-exchanger surface and that everyfluid has a heat-source inlet temperature for which the plant

efficiency is maximal. Recuperated ORCs are only useful whenthe heat-source outlet temperature is constrained and the bestdouble-pressure subcritical ORCs perform better than the bestsingle-pressure transcritical cycles.

The cooling-fluid inlet temperature and mass flow have a stronginfluence on the performance of the ORC. The next step in the sys-tem optimization is to model the cooling system and to include itin the optimization.

Acknowledgments

Daniël Walraven is supported by a VITO doctoral grant. Thevaluable discussions on optimization with Joris Gillis (KU Leuven)are gratefully acknowledged and highly appreciated.

Appendix A. Heat transfer and pressure drop correlations

The heat-transfer and pressure-drop correlations used in thispaper are given in this appendix.

A.1. Single-phase heat transfer and pressure drop

The model of Martin [13] is used to calculate the pressure dropand heat transfer coefficient in chevron-type plate heat exchangersfor single-phase flows. The frictional-pressure drop for single-phase flow in one plate is given by:

Dpplatesingle

h ifr¼ 4

f platesingleLp

Dh

G2

2q; ðA:1Þ

with f platesingle the Fanning friction factor, G ¼ _m

2aWNpthe mass velocity

and Np the number of channels. The Fanning friction factor is givenby:

1ffiffiffiffiffiffiffiffiffiffif platesingle

q ¼ cos b

0:045 tan bþ 0:09 sin bþ f0cos b

� �0:5 þ1� cos bffiffiffiffiffiffiffiffiffiffiffiffi

3:8f 1

p ; ðA:2Þ

with f0 the Fanning friction coefficient if b ¼ 0� and f1 the Fanningfriction coefficient if b ¼ 90�. These coefficients are given by:

f0 ¼16Re for Re < 2000

ð1:56 ln Re� 3:0Þ�2 for Re P 2000

(; ðA:3Þ

f1 ¼

149:25Re þ 0:9625 for Re < 2000

9:75Re0:289 for Re P 2000

8><>: ; ðA:4Þ

Page 10: Energy Conversion and Management - EnergyVille · flow rate on the performance of the power plant are also investi- ... c and the baffle spacing at the inlet L b;i, ... chevron,

236 D. Walraven et al. / Energy Conversion and Management 87 (2014) 227–237

where Re ¼ GDhl is the Reynolds number.

The hydrostatic-pressure drop is neglected because the fluidsflow alternately up and down and the hydrostatic pressure dropis therefore alternately positive and negative. Both are about equaland can therefore be assumed to neutralize each other.

The heat-transfer coefficient is calculated as:

h ¼ kDh

0:205Pr1=3 lm

lw

� �1=6

ðf platesingleRe2 sin 2bÞ

0:374; ðA:5Þ

with lm the viscosity of the bulk fluid and lw the viscosity of thefluid at the wall. The difference between these viscosities isneglected. Pr is the Prandtl number.

A.2. Heat transfer and pressure drop while evaporating

Han et al. [16] developed correlations for the pressure drop andheat transfer in chevron-type plate heat exchangers during evapo-ration. The frictional pressure drop in one plate during evaporationis:

Dpplateevap

h ifr¼ 4

f plateevap Lp

Dh

G2eq

2ql; ðA:6Þ

with Geq the equivalent mass velocity, defined as:

Geq ¼ G 1� xþ x ql=qv� �1=2

h i; ðA:7Þ

where ql and qv are the densities of saturated liquid and saturatedvapor, respectively.

The correlation for the Fanning friction factor is given by:

f plateevap ¼ c1Rec2

eq; ðA:8Þ

with

c1 ¼ 64710KDh

� ��5:27

b�3:03; ðA:9Þ

c2 ¼ �1:314KDh

� ��0:62

b�0:47; ðA:10Þ

Reeq ¼GeqDh

ll: ðA:11Þ

The acceleration pressure drop is given by:

acDpplateevap ¼

G2eqx

ql � qv

!out

�G2

eqxql � qv

!in

: ðA:12Þ

The heat transfer coefficient is correlated as:

h ¼ kDh

c3Rec4eqBo0:3

eq Pr0:4l ; ðA:13Þ

with:

c3 ¼ 2:81KDh

� ��0:041

b�2:83; ðA:14Þ

c4 ¼ 0:746KDh

� ��0:082

b0:61; ðA:15Þ

Boeq ¼_q

Geqðhv � hlÞ: ðA:16Þ

Boeq is the equivalent boiling number.The experiments performed by Han et al. [16] were done for

b = 45�, 55� and 70�, G = 13–34 kg/m2 s and a heat flux of2.5–8.5 kW/m2, while using R22 and R410a as working fluids.

A.3. Heat transfer and pressure drop while condensing

The correlations for the pressure drop and the heat transfercoefficient while condensing are given by Han et al. [17] and areanalogous to the ones for the evaporation, albeit with differentlydefined coefficients ci and h:

Dpplatecond

h ifr¼ 4

f platecond Lp

Dh

G2eq

2ql; ðA:17Þ

f platecond ¼ c5Rec6

eq; ðA:18Þ

c5 ¼ 3521:1KDh

� �4:17

b�7:75; ðA:19Þ

c6 ¼ �1:024KDh

� �0:0925

b�1:3; ðA:20Þ

h ¼ kDh

c7Rec8eqPr1=3

l ; ðA:21Þ

c7 ¼ 11:22KDh

� ��2:83

b�4:5; ðA:22Þ

c8 ¼ 0:35KDh

� �0:23

b1:48: ðA:23Þ

The experiments performed by Han et al. [17] were done for b = 45�,55� and 70� and G = 13–34 kg/m2 s, while using R22 and R410a asworking fluids.

References

[1] Tester J, Anderson B, Batchelor A, Blackwell D, DiPippo R, Drake E, et al. Thefuture of geothermal energy: impact of enhanced geothermal systems (EGS) onthe United States in the 21st century. Tech. Rep., Massachusetts Institute ofTechnology, Massachusetts, USA; 2006.

[2] Dai Y, Wang J, Gao L. Parametric optimization and comparative study oforganic Rankine cycle (ORC) for low grade waste heat recovery. EnergyConvers Manag 2009;50(3):576–82.

[3] Saleh B, Koglbauer G, Wendland M, Fischer J. Working fluids for low-temperature organic Rankine cycles. Energy 2007;32(7):1210–21.

[4] Walraven D, Laenen B, D’haeseleer W. Comparison of thermodynamic cyclesfor power production from low-temperature geothermal heat sources. EnergyConvers Manag 2013;66:220–33.

[5] Walraven D, Laenen B, D’haeseleer W. Optimum configuration of shell-and-tube heat exchangers for the use in low-temperature organic Rankine cycles.Energy Convers Manag 2014;83:177–87.

[6] Madhawa Hettiarachchi H, Golubovic M, Worek WM, Ikegami Y. Optimumdesign criteria for an organic Rankine cycle using low-temperature geothermalheat sources. Energy 2007;32(9):1698–706.

[7] Quoilin S, Lemort V, Lebrun J. Experimental study and modeling of an organicRankine cycle using scroll expander. Appl Energy 2010;87(4):1260–8.

[8] Shengjun Z, Huaixin W, Tao G. Performance comparison and parametricoptimization of subcritical Organic Rankine Cycle (ORC) and transcriticalpower cycle system for low-temperature geothermal power generation. ApplEnergy 2011;88(8):2740–54.

[9] Domingues A, Santos H, Costa M. Analysis of vehicle exhaust waste heatrecovery potential using a Rankine cycle. Energy 2013;49:71–85.

[10] Franco A, Villani M. Optimal design of binary cycle power plants for water-dominated, medium-temperature geothermal fields. Geothermics2009;38(4):379–91.

[11] Hewitt GF. Hemisphere handbook of heat exchanger design. NewYork: Hemisphere Publishing Corporation; 1990.

[12] Shah RK, Sekulic DP. Fundamentals of heat exchanger design. John Wiley andSons, Inc.; 2003.

[13] Martin H. A theoretical approach to predict the performance of chevron-typeplate heat exchangers. Chem Eng Process 1996;35(4):301–10.

[14] Ayub ZH. Plate heat exchanger literature survey and new heat transfer andpressure drop correlations for refrigerant evaporators. Heat Transfer Eng2003;24(5):3–16.

[15] Djordjevic E, Kabelac S. Flow boiling of R134a and ammonia in a plate heatexchanger. Int J Heat Mass Transfer 2008;51(25):6235–42.

[16] Han D-H, Lee K-J, Kim Y-H. Experiments on the characteristics of evaporationof R410A in brazed plate heat exchangers with different geometricconfigurations. Appl Therm Eng 2003;23(10):1209–25.

[17] Han D-H, Lee K-J, Kim Y-H. The characteristics of condensation in brazed plateheat exchangers with different chevron angles. J Korean Phys Soc2003;43:66–73.

[18] Jokar A, Hosni MH, Eckels SJ. Dimensional analysis on the evaporation andcondensation of refrigerant R-134a in minichannel plate heat exchangers. ApplTherm Eng 2006;26(17):2287–300.

Page 11: Energy Conversion and Management - EnergyVille · flow rate on the performance of the power plant are also investi- ... c and the baffle spacing at the inlet L b;i, ... chevron,

D. Walraven et al. / Energy Conversion and Management 87 (2014) 227–237 237

[19] Longo GA. R410A condensation inside a commercial brazed plate heatexchanger. Exp Therm Fluid Sci 2009;33(2):284–91.

[20] Thonon B, Bontemps A. Condensation of pure and mixture of hydrocarbons in acompact heat exchanger: experiments and modelling. Heat Transfer Eng2002;23(6):3–17.

[21] Lemmon E, Huber M, Mclinden M. NIST reference fluid thermodynamic andtransport properties? REFPROP, The National Institute of Standards andTechnology (NIST), version 8.0; 2007.

[22] Martins JR, Sturdza P, Alonso JJ. The complex-step derivative approximation.ACM Trans Math Softw (TOMS) 2003;29(3):245–62.

[23] Peterson P. F2PY: a tool for connecting Fortran and Python programs. Int JComput Sci Eng 2009;4(4):296–305.

[24] Andersson J, Åkesson J, Diehl M. CasADi – a symbolic package for automaticdifferentiation and optimal control. In: Forth S, Hovland P, Phipps E, Utke J,Walther A, editors. Recent advances in algorithmic differentiation, Lecturenotes in computational science and engineering, vol. 87. Berlin,Heidelberg: Springer; 2012. p. 297–307.

[25] Büskens C, Wassel D. The ESA NLP Solver WORHP. In: Modeling andoptimization in space engineering. Springer; 2013. p. 85–110.