Numerical study on crossflow printed circuit heat exchanger for advanced small modular reactors

Numerical study on crossflow printed circuit heat exchanger for advanced small modular reactors

International Journal of Heat and Mass Transfer 70 (2014) 250–263 Contents lists available at ScienceDirect International Journal of Heat and Mass T...

2MB Sizes 0 Downloads 24 Views

International Journal of Heat and Mass Transfer 70 (2014) 250–263

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer journal homepage: www.elsevier.com/locate/ijhmt

Numerical study on crossflow printed circuit heat exchanger for advanced small modular reactors Su-Jong Yoon a,⇑, Piyush Sabharwall a, Eung-Soo Kim b a b

Idaho National Laboratory, 2525 Fremont Avenue, Idaho Falls, ID 83415-6188, USA Seoul National University, Seoul, Republic of Korea

a r t i c l e

i n f o

Article history: Received 16 July 2013 Received in revised form 30 October 2013 Accepted 31 October 2013

Keywords: Crossflow Heat exchanger PCHE Analytical model Thermal design Cost estimation Advanced SMR

a b s t r a c t Various fluids such as water, gases (helium), molten-salts (FLiNaK, FLiBe) and liquid metal (sodium) are used as a coolant of advanced small modular reactors (SMRs). The printed-circuit heat exchanger (PCHE) has been adopted as the intermediate and/or secondary heat exchanger of SMR systems because this heat exchanger is compact and effective. The size and cost of PCHE can be changed by the coolant type of each SMR. In this study, the crossflow PCHE analysis code for advanced small modular reactor has been developed for the thermal design and cost estimation of the heat exchanger. The analytical solution of singlepass, both unmixed fluids crossflow heat exchanger model was employed to calculate a two-dimensional temperature profile of a crossflow PCHE. The analytical solution of crossflow heat exchanger was simply implemented by using built-in function of the MATLAB program. The effect of fluid property uncertainty on the calculation results was evaluated. In addition, the effect of heat transfer correlations on the calculated temperature profile was analyzed by taking into account possible combinations of primary and secondary coolants in the SMR systems. Size and cost of heat exchanger were evaluated for the given temperature requirement of each SMR. Ó 2013 Elsevier Ltd. All rights reserved.

1. Introduction Small modular reactors (SMR) are being developed to provide heat and electricity to the remote regions where the large nuclear power plants cannot be built. In addition, recently, the SMRs have been considered as good energy sources for the process heat application. Contrary to typical water-cooled SMRs, advanced SMRs use various coolant materials such as gases (helium, supercritical carbon dioxide), molten-salts (FLiNaK, FLiBe and KFZrF4), and liquid metal (sodium). The operating temperature and pressure of an advanced SMR can be varied by the type of reactor coolant. Table 1 describes the typical temperature and pressure conditions of SMRs according to the reactor coolants. Inlet and outlet temperatures of primary and secondary fluids are based on the intermediate heat exchanger (IHX). Thus, the inlet temperature of the primary fluids shown in Table 1 is the outlet temperature of reactor coolant. Fig. 1 shows some of the potential options that the IHX and secondary heat exchanger (SHX) could be configured in the system. It could be linked to production of power system and also could be used for providing the process heat from the reactor for industrial applications. In this case, both IHX and SHX could be either con-

⇑ Corresponding author. Tel.: +1 208 526 5407/357 8858; fax: +1 208 526 2930. E-mail addresses: [email protected], [email protected] (S.-J. Yoon). 0017-9310/$ - see front matter Ó 2013 Elsevier Ltd. All rights reserved. http://dx.doi.org/10.1016/j.ijheatmasstransfer.2013.10.079

ventional-type heat exchanger or compact heat exchanger, such as PCHE. The word ‘‘small’’ in SMRs refers not only to the reactor power, but also to the size of the reactor. In this regard, the compact heat exchanger can be a good design option because the compact heat exchanger has a large heat transfer surface area per unit volume, which can result in the reduced size, weight, and cost. Fig. 2 shows the surface compactness that is a criterion for classifying the heat exchanger. Printed circuit heat exchanger (PCHE) is a compact heat exchanger and has been considered one of the candidate heat exchangers in designs of advanced SMRs. Fine grooves in the plate of the PCHE are made by employing the technique used for making printed circuit boards. The PCHE has been commercially manufactured by Heatric™ [1]. Fig. 3 shows the typical PCHE manufactured by Heatric Ltd. The PCHE can be classified by the flow configurations. The arrangement of flow channels are parallel in the parallel and counter flow PCHEs, whereas it is orthogonal in the crossflow PCHE. Temperature distribution of the crossflow heat exchanger is twodimensional so that it is more complicated to analyze than parallel or counter flow heat exchangers. Due to this complexity, many previous studies have used computational fluid dynamics (CFD) codes such as CFX or FLUENT. However, CFD simulation is not adequate for the design process of the PCHE since CFD simulation requires vast computational time and cost. In the CFD model of PCHE, since

251

S.-J. Yoon et al. / International Journal of Heat and Mass Transfer 70 (2014) 250–263

Nomenclature A C CM Cop cp C⁄ CP D, d f h I0 k L M _ m N NTU Nu OP P p Pe Pr Q R Re T t U X Y

area or overall heat transfer surface area (m2) _ p;j Þ flow stream heat capacity rate ð¼ mc material cost factor (USD/kg) operating cost factor (USD/Wh) specific heat capacity (kJ/kg K) the ratio of heat capacity rate (=Cmin/Cmax) capital cost of PCHE (USD) diameter (m) friction factor convective heat transfer coefficient (W/m2 K) modified Bessel function of the first kind and zero order thermal conductivity (W/m K) length (m) total mass of heat exchanger (kg) mass flow rate (kg/s) number of flow channel number of transfer unit Nusselt number (=hD/k) operating cost of PCHE (USD) channel pitch (m) or Pressure (Pa) the order of scheme Peclet number (=Re  Pr) Prandtl number (=cpl/k) heat duty (W) thermal resistance (m2  K/W) Reynolds number (=qVD/l) temperature (K) thickness (m) overall heat transfer coefficient (W/m2 K) normalized length in x-coordinate (=x/L1NTU1) normalized length in y-coordinate (=y/L2NTU2) or Operation period of heat exchanger (yr)

the structure of PCHE is complicated and the dimensions of flow channel of PCHE are small, the fine structure of mesh is required to obtain accurate results. In this study, we developed the crossflow PCHE analysis code to perform the thermal design and cost estimation of PCHE.

2. Development of crossflow PCHE analysis code In order to perform the thermal design and cost estimation of a crossflow PCHE, the analysis code was developed based on the MATLAB program [8]. The analytical solution of single-pass, both unmixed fluids crossflow heat exchanger was employed to solve the temperature profile of crossflow PCHE. Although typical PCHEs manufactured by Heatric™ have zigzag-shaped flow channel, the straight pipe is assumed in this study. The analytical modeling method, analytical solution, and thermal design procedure are described in following sections.

V

volume (m3) or Fluid velocity (m/s)

Greek and symbols D difference d plate thickness e effectiveness of PCHE (=Q/Qmax) go surface efficiency h dimensionless temperature (=(TT2,in))/(T1,inT2,in) l dynamic viscosity of fluid (Pa s) _ DP=qÞ P pumping power ð¼ m q density (kg/m3) U approximated exact solution by Richardson extrapolation method / numerical solution Subscript f h HX in j max min L out p tot w x–s

fin grid width or Hydraulic diameter heat exchanger inlet index of fluid (1: primary, 2: secondary) maximum minimum Laplace transformation operation outlet plate total wall cross-sectional

2.1. Analytical modeling and solution of crossflow heat exchanger For single-pass crossflow design, a solution to the problem of both fluids unmixed in the heat exchanger was obtained by Nusselt [9] in the form of analytical series expansions by assuming the longitudinal conduction can be neglected. Since the flow mixing is not occurred in the PCHE, this model is applicable. The analytical model of crossflow PCHE has been developed by the energy balance between the hot and cold fluid sides shown in Fig. 4. The energy balance equations in Fig. 3 are as follows: Fluid 1 (Hot fluid):   @T _ 1 cp;1 T1 _ 1 cp;1 T1 þ 1 dx dm dm dQ |{z} |fflfflfflfflfflfflffl{zfflfflfflfflfflfflffl} @x     |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}  Heat transfer rate ¼ 0   Fluid enthalpy rate Fluid enthalpy rate  from fluid to wall into the C:V: out of the C:V: ð1Þ

Table 1 Typical temperature and pressure conditions of advanced SMRs. Coolant type

Water-cooled [2] Gas-cooled [3] Liquid metal-cooled [4] Molten salt-cooled [5]

Coolants

Operation conditions

Primary/secondary

Inlet/outlet temperature (°C)

Water/Water He/He Na/Na FLiBe/FLiNaK

Primary

Secondary

323/295 950/637 545/390 700/650

200/293 351/925 320/526 600/690

Primary/secondary pressure (MPa)

15.0/5.0 7.0/7.0 0.1/0.1 0.1/0.1

252

S.-J. Yoon et al. / International Journal of Heat and Mass Transfer 70 (2014) 250–263

(a) Option 1 – 1 SHX

IHX

Reactor

SHX

IHX

Power Conversion System

SHX

IHX

Reactor

SHX

(b) Option 2 – 2 SHXs Power Conversion System

1

SHX

IHX

2

2

3 4

(c) Option 3 – 3 SHXs

Power Conversion System

Reactor IHX

3

1

5

Fig. 3. Section of crossflow PCHE [7].

In above energy balance equations, the heat transfer from the hot fluid to the cold fluid can be calculated by the term dQ that can be expressed by using the rate equations for convection and conduction as follows:

6

IHX 6

SHX

g0;1 h1 ðT1  Tw;1Þdxdy

|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}

SHX 4

dQ ¼

5

Fig. 1. Schematic of potential heat exchanger arrangements.

Fluid 2 (Cold fluid):

  @T _ 2 cp;2 T2 þ 2 dy _ 2 cp;2 T2 dm dm dQ |{z} @y |fflfflfflfflfflfflffl{zfflfflfflfflfflfflffl}  |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} ¼ 0      Fluid enthalpy rate þ Heat transfer rate Fluid enthalpy rate from fluid to wall  into the C:V: out of the C:V:

¼ ð2Þ

Convection

!

from hot fluid to wall   Tw;1  Tw;2 dxdy kw dw |fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl} ð Conduction within the wall Þ

Fig. 2. Heat transfer surface area density spectrum of heat exchanger surfaces [6].

g0;2 h2 ðTw;2  T2 Þdxdy

|fflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl{zfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflfflffl}

¼

!

Convection from wall to cold fluid

ð3Þ

253

S.-J. Yoon et al. / International Journal of Heat and Mass Transfer 70 (2014) 250–263

Fig. 4. Energy balance control volume for crossflow heat exchanger [10].

By employing the definition of the overall heat transfer coefficient U ¼ dAdQDT, but neglecting the fouling thermal resistances, the heat transfer rate dQ can be given by:

dQ ¼ UdAðT1  T2 Þ

ð4Þ

where dA = dxdy. By substituting Eq. (4) into Eqs. (1) and (2), and simplifying, the following differential equations for the crossflow heat exchanger are obtained:

@h1 ðX; YÞ þ h1 ðX; YÞ ¼ h2 ðX; YÞ @X

ð5Þ

Since the Laplace transform of the derivative of function is n o L dFðnÞ ¼ sLfFðnÞgn!s  Fð0Þ, Eq. (14) can be expressed as dn n!s

follow:

ph2L þ h2 ðX;0Þ þ h2L ¼ h1L )

By applying Laplace transform for variable X to Eq. (13), the following expression can be obtained with the boundary condition of h1 in Eq. (9):

n n o 1 L L dh dX

o

X!s

@h2 ðX; YÞ þ h2 ðX; YÞ ¼ h1 ðX; YÞ @Y

ð6Þ

UA UA 1 ¼ ¼ _ j cp;j m Cj Cj

Z

UdA ðj ¼ 1; 2Þ

ð7Þ

ð8Þ

If the inlet temperature distributions of primary and secondary sides are assumed to be uniform, following two boundary conditions are accompanied.

h1 ð0; YÞ ¼ 1;

h2 ðX;0Þ ¼ 0

ð9Þ

The solution of systems of partial differential equations can be obtained by implementing the Laplace transform and inverse transform. Exact solutions of this problem have been reported by Nusselt [9,11]. And there are various crossflow heat exchanger models [12–19], but it is proven that those all models are alternative expressions of Nusselt’s model [20]. Thus, in this study, the Nusselt model is employed as the analytical solution of the crossflow heat exchanger. First, the Laplace transforms for each variable can be defined as follows:

hiL ¼ hiL ðX; pÞ ¼ Lfh1 ðX; YÞgY!p ;

i ¼ 1; 2

ð10Þ

~hiL ¼ ~hiL ðs; pÞ ¼ Lfh1L ðX; pÞg ; X!s

i ¼ 1; 2

ð11Þ

LfLffðX; YÞgX!s gY!p ¼ LfLffðX; YÞgY!p gX!s

ð12Þ

To obtain the solution of the partial differential equations, first, the Laplace transform for variable Y is applied to Eqs. (5) and (6).

dh1L þ h1L ¼ h2L dX dh2L þ h2L ¼ h1L dY

þ Lfh1L gX!s ¼ Lfh2L gX!s )

ð16Þ

ðs þ 1Þ~h1L ¼ ~h2L þ 1p

~h1L ¼

The number of heat transfer units, NTU, is defined as a ratio of the overall thermal conductance to the heat capacity rate. NTU is given by:

NTUj ¼

Y!p

From algebraic Eqs. (15) and (16), ~ h1L and ~ h2L are given by:

where 0 6 X (=x/Lx) 6 NTU and 0 6 Y (=y/Ly) 6 C⁄NTU. The dimensionless temperature hj is defined by:

hj ¼ ðTj  T2;in Þ=ðT1;in  T2;in Þ ðj ¼ 1; 2Þ

ð15Þ

ðp þ 1Þh2L  h1L ¼ 0

ð13Þ ð14Þ

!    pþ1 1 1 1 ¼ p p ps þ s þ p p s þ pþ1

ð17Þ

!      1 1 1 1 ¼ p p ps þ s þ p pðp þ 1Þ s þ pþ1 !   1 1 1 ¼  p p p þ 1 s þ pþ1

~h2L ¼

ð18Þ

The following Laplace transform pairs and their relationship are used to obtain the solutions of Eqs. (17) and (18).

 9 8 p
ð19Þ

p!Y

V1 ðX; YÞ  1 

Z

X

 pffiffiffiffiffiffi eðuþYÞ I0 2 Yu du ¼ L1

0

 9 8 p
p

;

p!Y

ð20Þ V1 ðX; YÞ  V1;0 ðX; YÞ ¼ 1  V1 ðY; XÞ

ð21Þ

Thus, by applying the inverse Laplace transform of Eqs. (17) and (18), the final solutions are given by:

(

!) ) 1 1 h1 ðX; YÞ ¼ L L p p s þ pþ1 s!X p!Y

  p 1 1 exp  X ¼L ¼ V1 ðX; YÞ p pþ1 p!Y ( ( !) )  1 1 1 1 1 h2 ðX; YÞ ¼ L L  p p p þ 1 s þ pþ1 s!X p!Y

    1 1 p exp  X ¼ L1  p pþ1 pþ1 p!Y 1

(

1

¼ V1 ðX; YÞ  V1;0 ðX; YÞ ¼ 1  V1 ðY; XÞ

ð22Þ

ð23Þ

254

S.-J. Yoon et al. / International Journal of Heat and Mass Transfer 70 (2014) 250–263

Above solutions can be easily implemented by using built-in functions of the MATLAB program. MATLAB provides the modified Bessel function of the first kind and zero order and integration of the function. The temperature solution of the crossflow heat exchanger is obtained simply and quickly by solving Eq. (19). 2.2. Thermal design of PCHE To meet heat requirements of the inlet and outlet temperatures of fluids, the size of heat exchanger is calculated through the thermal design procedure. The outlet temperature was averaged along the perpendicular to the flow direction in each fluid side since the temperature profile of crossflow heat exchanger is two-dimensional distribution. Thermal design process of crossflow PCHE is shown in Fig. 5. In general, the size of the heat exchanger is determined by the fluid temperatures. Fig. 6 shows the schematic diagram of crossflow PCHE. Geometric parameters of PCHE were tabulated in Table 2. In order to determine the size of crossflow PCHE, at first, the sizes of crossflow PCHE (Lx, Ly, and Lz) are assumed. From the assumed geometry information, the overall heat transfer surface area A is calculated as follows:



p

p d þ d Lx N1 Nz ¼ d þ d Ly N2 Nz 2 2

Fig. 6. Schematic diagram of crossflow PCHE.

Table 2 Design parameters of crossflow PCHE.

ð24Þ

The number of flow channels in each fluid side is determined as follows:



Geometric parameters

Value

Channel diameter d, (m) Channel pitch P, (m) Plate thickness tp, (m)

0.003 0.0033 0.00317

p

p ðLy þ tf Þ ðLx þ tf Þ d þ d Lx Lz ¼ d þ d Ly Lz 2 2 2tf ðtf þ dÞ 2tf ðtf þ dÞ

ð28Þ

N1 d þ ðN1  1Þtf ¼ Ly

ð25Þ

The mass flow rate through each fluid side is calculated from the heat duty and the given temperature conditions by using the following heat balance equations:

N2 d þ ðN2  1Þtf ¼ Lx

ð26Þ

_ 1 cp;1 ðT1;in  T1;out Þ ¼ m _ 2 cp;2 ðT2;out  T2;in Þ Q ¼m

Nz ¼ Lz =ð2tp Þ

ð27Þ

Simplifying Eqs. (25)–(27) and substituting them to Eq. (24) yields:

Vj ¼

Calculate overall heat transfer surface area, As

Calculate mass flow rate . and fluid velocity, mf , Vf

Calculate Temperature Distribuon

Calculate heat transfer coefficient, h

Calculate average outlet − temperature Tout

Resize

_j m

−out| <ε |Tout-T Yes Determinaon of HX sizes

Fig. 5. Thermal design process of crossflow PCHE.

ð31Þ

The Reynolds number and Prandtl number of each fluid side are calculated as follows:

Rej ¼

qj Vj Dh;j lj

Prj ¼

cp;j lj kj

ðj ¼ 1; 2Þ

ðj ¼ 1; 2Þ

ð32Þ

ð33Þ

The heat transfer coefficient is calculated by empirical heat transfer correlations. The heat transfer correlations used in this study will be discussed in Section 2.4. When the Nusselt number is determined by empirical correlation, the heat transfer coefficient in each fluid side is calculated as follows:

hj ¼ Nuj No

ð30Þ

2;in Þ

ðj ¼ 1; 2Þ

qj Axs;j

Calculate Number of Transfer Unit, NTU

Calculate Non-dimensional parameters (Re, Pr, Nu)

Calculate overall heat transfer coefficient, U

_ 1 ¼ c ðT QT Þ m p;1 1;in 1;out

The fluid velocity of each fluid side is calculated as follows:

Given Requirements and condions of HX: Q, Dh , Tin , Tout , P, material properes f(P, T)

Set the inial sizes of HX

thus:

_ 2 ¼ c ðT Q T m p;2 2;out

Thermal Design Process

ð29Þ

kj Dh;j

ðj ¼ 1; 2Þ

ð34Þ

Overall heat transfer coefficient of crossflow PCHE is given by:



1 h1

1 1 ¼ þ Rw þ h12 h1 þ kdww þ h1 1 2

ð35Þ

Calculated values in Eqs. (28), (30), and (35) are used to calculate the NTU in Eq. (8). When the NTU is determined, temperature profile of the crossflow heat exchanger can be calculated by the

S.-J. Yoon et al. / International Journal of Heat and Mass Transfer 70 (2014) 250–263

analytical solutions in Eqs. (22) and (23). Dimensionless temperature profiles, hj(X, Y), of crossflow PCHE are converted to Tj(x, y) as follows: Tj ðx;yÞ ¼ ðT1;in  T2;in Þhj ðx;yÞ þ T2;in sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi! # "    Z NTU1 x Lx NTU2 NTU2 ¼ ðT1;in  T2;in Þ 1  exp  y þ u Io 2 yu du þ T2;in Ly Ly 0 ð36Þ

where 0 6 x 6 Lx, 0 6 y 6 Ly. The mean outlet temperatures of primary and secondary fluids can be calculated by averaging each temperature profile function as follows:

T1;out ¼ ¼

T2;out ¼ ¼

1 Ly 1 Ly 1 Lx 1 Lx

Z Z

T1 ðLx ; yÞdy 0



ðT1;in  T2;in Þhj ðLx ; yÞ þ T2;in dy

ð37Þ

0

Z Z

L1

T2 ðx;Ly Þdx 0 Lx



ðT1;in  T2;in Þhj ðx;Ly Þ þ T2;in dx

ð38Þ

0

Above mean outlet temperatures should be corresponded to the required outlet temperature conditions. If the mean outlet temperatures disagree with the required outlet temperatures, then the heat exchanger has to be resized until the calculated mean outlet temperatures are good agreement with the required outlet temperatures. From the calculated outlet temperatures, log mean temperature difference (LMTD) of crossflow PCHE can be obtained by:

DTLMTD ¼

ðT1;in  T2;out Þ  ðT1;out  T2;in Þ   ðT T Þ ln ðT1;in T2;out Þ 1;out

ð39Þ

2;in

In order to estimate the operational cost of heat exchanger, the pressure loss of the heat exchanger has to be calculated. Since the channel shape of PCHE is a semi-circle, the pressure loss calculation using the friction factor correlation for a circular pipe can cause an incorrect result. The fully-developed laminar friction factor of semi-circular straight pipe [21] is given by:

f  Re ¼ 15:78;

channel should be employed. However, the straight pipe is assumed in this study so that the friction factor correlation by Berbish et al. [22] is employed to calculate the turbulent friction factor of semi-circular straight pipe. The friction factor is given by:

f ¼ 0:478Re0:26 ; 8200  Re  58000

ð41Þ

The pressure loss of the cross flow PCHE can be calculated by the pressure loss equation of straight pipe:

DPj ¼ 2f j

Lj q V2 Dh;j j j

ðj ¼ 1; 2Þ

ð42Þ

2.3. Thermo-physical properties of working fluids

L2

Ly

255

Re < 2300

ð40Þ

Hesselgreaves [21] suggests the turbulent friction factor of PCHE of which the flow channel is wavy. If the flow channel is wavy or zigzag-shaped, the correlations for wavy or zigzag-shaped

The working fluids used in advanced SMRs are water, helium, sodium, and molten-salts (FLiNaK, FLiBe). In this study, IAPWS IF97 standard formulation [23] is used to calculate the water properties. The helium property is based on the ‘‘National Institute of Standards Technology (NIST) Chemistry WebBook’’ data [24]. Thermophysical properties of sodium and molten-salts are summarized in Table 3. The unit of temperature, T, in Table 3 is Kelvin. Some material properties of molten-salts are not accurate and have to be investigated in the future. In this study, the effect of uncertainty of material property was investigated. The maximum uncertainty of property is assumed to be ±30%. 2.4. Correlations for heat transfer coefficient The heat transfer correlations have been developed with various fluids under the various conditions. Dittus-Boelter [28] and Gnielinski [29] correlations are well-known and widely-used heat transfer correlations, but it is known that Dittus-Boelter has definite uncertainty. Gnielinski correlation is used as the default heat transfer correlation of the developed code. The heat transfer correlations are summarized in Table 4. Each correlation is valid for limited ranges of temperature and pressure, shape of channel, and fluid material. A heat transfer coefficient calculated by the empirical correlation has an influence on the temperature profile of the crossflow heat exchanger and consequently on the heat exchanger sizing. Therefore, the appropriate selection of a heat transfer correlation is important for accurate heat exchanger sizing. In addition, the empirical heat transfer correlations developed for circular pipe might predict the heat transfer coefficient incorrectly since the flow channel of PCHE is a semi-circle. In this study, the temperature profiles according to

Table 3 Thermophysical properties of sodium, FLiNaK and FLiBe. Sodium [25] Density (kg/m3) Heat capacity (J/(kg K)) Thermal conductivity (W/(m K)) Dynamic viscosity (Pa s)

896.60679 + 0.5161343T  1.8297218  103T2 + 2.2016247  106T3 – 1.3975634  109T4 + 0.44866894  1012T5 – 0.057963628  1015T6 (38.12 – 0.069  106T2 – 19.493  103T + 10.24  106T2)/22.99 99.5 – 39.1  103T exp(6.44060.3958lnT + 556.8/T) FLiNaK (LiF-NaF-KF) [26]

Density (kg/m3) Heat capacity (J/(kg K)) Thermal conductivity (W/(m K)) Dynamic viscosity (Pa s)

2,579.3 – 0.624T 976.78 + 1.0634T 0.36 + 5.6  104T

Density (kg/m3) Heat capacity (J/(kg K)) Thermal conductivity (W/(m K)) Dynamic viscosity (Pa s)

2,413 – 0.488T 2397.73 ( ± 20%) 0.629697 + 0.0005T

2.487  105exp(4,478.62/T) FLiBe (LiF-BeF2) [26]

0.000116exp(3,755/T)

256

S.-J. Yoon et al. / International Journal of Heat and Mass Transfer 70 (2014) 250–263

Table 4 Summary of single-phase heat transfer coefficient correlations. Authors

Channel shape

Working fluid

Berbish [22] Kim et al. [27] Dittus-Boelter [28]

Semi-circular Semi-circular Circular

Air He Any fluid

Gnielinski (NuG) [29]

Circular

Any fluid

Nusselt number correlation 0.0228Re 3.255 + 0.00729(Re350) 0.0243Re0.8Pr0.4 for heating 0.0265Re0.8Pr0.3 for cooling ð8f ÞðRe1000ÞPr pffi

8,200 < Re < 5.83  104 350 < Re < 800, Pr = 0.66 104 < Re < 1.2  105, 0.7 < Pr < 120, L/D > 60

1þ12:7

Lyon [30] Lubarsky and Kaufman [31] Reed [32] Taylor and Kirchgessner [33], Wieland [34] Wu and Little [35] Wang and Peng [36]

Circular Circular, Annular Circular Circular

Liquid Metal Liquid Metal Liquid Metal He

7.0 + 0.025Pe0.8 0.625Pe0.4 3.3 + 0.02Pe0.8 0.021Re0.8Pr0.4

0 6 Pr 6 0.1, 300 < Pe < 104 0 6 Pr 6 0.1, 200 < Pe < 104 100 < Pe 3.2  103 < Re < 6  104, 60 < L/D

Rectangular Rectangular

N2 gas Water, Methanol

0.00222Re1.09Pr0.4 0.00805Re0.8Pr0.33

3000 < Re 2300 < Re

2.5. Cost estimation methodology The total cost of a heat exchanger consists of capital cost and operating cost. The capital cost is related to the material cost of the heat exchanger so that it can be estimated based on its weight. Therefore, the capital cost depends on the structural material and size of the heat exchanger. The total mass of heat exchanger is given by:

MHX ¼ qHX VHX

ð43Þ

The volume of the heat exchanger can be approximated as follows:

VHX ¼ Lx Ly Lz

ð44Þ

The capital cost can be calculated by multiplying the cost factor of material ($/kg) and the total mass of heat exchanger, as follows:

CP ¼ CM MHX

ð45Þ

On the other hand, the operating cost is estimated based on the thermal-hydraulic condition of the heat exchanger. Since the operating cost is defined as the cost to operate the circulation pump of the heat exchanger, it is important to optimize the design of heat exchanger to reduce the pressure loss, in turn reducing the operating cost. To estimate the operating cost, the pumping power of the heat exchanger is used. The pumping powers of the heat exchanger in hot and cold fluid sides can be defined by [37]:

P2 ¼

2300 < Re < 5  106, 0.5 < Pr < 2000

2=3 f 1Þ 8ðPr

the heat transfer correlations are compared to evaluate the effect of the heat transfer correlation.

P1 ¼

Valid ranges

0.8

_ 1 DP1 m

ð46Þ

q1 _ 2 DP2 m

ð47Þ

q2

Thus, the operating cost can be approximated as follows [37]:

OP ¼ COP YðP 1 þ P 2 Þ

ð48Þ

Therefore, the total cost of heat exchanger becomes:

Ctotal ¼ CP þ OP ¼ CM MHX þ COP YðP 1 þ P 2 Þ

ð49Þ

As a structural material of PCHE, the nickel-based alloys such as Alloy 800H, Alloy 617, and Hastelloy N are under development. The thermo-physical properties of structural material also have an influence on the sizing and cost estimation of heat exchanger. According to the previous studies on the cost of heat exchangers [37,38], the material costs of Alloy 800H and Alloy 617 are assumed to be 120 USD/kg conservatively. The material cost of Hastelloy N is approximately 124 USD/kg [39]. The operating cost is 0.0000612 USD/Wh, which is based on the consumer price index average price data [37]. For high temperature condition of 700 °C, the thermo-physical properties and material costs of PCHE structural materials are summarized in Table 5.

3. Crossflow PCHE analysis The fluid combinations of primary and secondary sides are different for each reactor coolant of SMR. In this study, four fluid combinations of primary and secondary side are selected by taking into account the reactor coolant types of SMR. The heat duty of heat exchanger is assumed to be 300 MW. The Reynolds numbers of primary and secondary sides are varied according to the fluid property and heat duty. Test conditions are summarized in Table 6.

3.1. Grid sensitivity test In this study, the temperature profile of crossflow PCHE was obtained by solving the system of differential equations. Twodimensional grid is used to calculate the temperature profile. In developed code, the mean outlet temperature of each fluid side is calculated by averaging the temperature at the outlet. Thus, the number of grids can influence the mean outlet temperature. In order to investigate the grid effect, the grid sensitivity test was carried out. Tested numbers of grid and results are summarized in Table 7. In grid sensitivity test, x and y-axial dimensions of heat exchanger, NTUs of primary and secondary sides are assumed to be a unity. The numbers of reference grid was 40  40.

Table 5 Thermo-physical properties and cost of structural materials of PCHE. Structural material

Density (kg/m3)

Thermal conductivity (W/(m K))

Heat capacity (J/(kg K))

Cost (USD/kg)

Alloy 617 [37] Alloy 800H [38] Hastelloy N [39]

8,360 7,940 8,860

23.9 22.8 23.6

586 460 523

120 120 124

257

S.-J. Yoon et al. / International Journal of Heat and Mass Transfer 70 (2014) 250–263 Table 6 Test matrix for crossflow PCHE analysis. Coolants (primary/secondary)

Inlet temperature (°C) (primary/secondary)

Pressure (MPa) (primary/secondary)

Heat duty (MW)

Re

Pr

Water/Water Helium/Helium Sodium/Sodium FLiBe/FLiNaK

320/200 950/350 545/320 700/600

15/15 7.0/7.0 0.1/0.1 0.1/0.1

300

104–105 104–105 104–105 103–104

1 0.66 0.005 10

Table 7 Grid sensitivity test result. Dimensionless mean outlet temperature Numbers of grids Primary side Secondary side

55 0.52124 0.47876

10  10 0.52251 0.47749

20  20 0.52314 0.47686

The grid sensitivity analysis by the Richardson extrapolation [40] was performed to estimate the numerical error due to the grid number. The approximation of exact solution U is given by:

U ¼ /h þ

/h  /2h 2p  1

ð50Þ

where / is a numerical solution and the order of the scheme p is defined by:

log p¼



/2h /4h /h /2h

ð51Þ

log2

Fig. 7 shows the dimensionless mean outlet temperatures of the primary and secondary sides. The dimensionless temperature in the grid sensitivity test was defined in Eq. (7). The Richardson solutions of primary and secondary fluids were 0.5238 and 0.4762, respectively. The maximum relative errors between the reference grid (40  40) and Richardson solution of primary and secondary fluids were 0.48% and 0.53%, respectively. Consequently, the grid sensitivity test results show that the effect of grid number is negligible. 3.2. Effectiveness-NTU (e-NTU) method analysis The effectiveness of heat exchanger, e, is the ratio of the actual heat transfer rate to the maximum possible heat transfer rate of the exchanger and is defined by:



Q ðT1;in  T1;out Þ ¼ Q max ðT1;in  T2;in Þ

ð52Þ

40  40 0.52346 0.47654

80  80 0.52362 0.47638

160  160 0.52370 0.47630

320  320 0.52374 0.47626

The e-NTU method [10] is one of general methods applied for the thermal design of heat exchanger. In this method, the effectiveness of crossflow heat exchanger is a function of NTU. The relationship between the effectiveness and NTU of both unmixed fluids crossflow heat exchanger [41] is given by:



e ¼ 1  exp

 1 0:22 0:78  ðNTUÞ ðexp½C ðNTUÞ   1Þ C

ð53Þ

where C⁄( = Cmin/Cmax) is the ratio of minimum to maximum heat capacity rate of fluid. The effectiveness of crossflow PCHE calculated by the developed code is compared with that of the e-NTU method. Fig. 8 shows the effectiveness-NTU plots by the developed code and e-NTU method. The effectiveness of a heat exchanger increases as the NTU increases. The effectiveness increases rapidly as the ratio of heat capacity rate decreases. The maximum relative error of effectiveness between the code and e-NTU method was 3.95%. 3.3. Effect of fluid property uncertainty The uncertainty of fluid property can be propagated into the calculated temperature profile of the heat exchanger. To evaluate the effect of uncertainty of fluid property, the uncertainty of fluid property was assumed to be ±30%. In this calculation, the heat duty, the temperatures and pressures in Table 1 were employed. Hastelloy N is employed as the structural material. Gnielinski heat transfer correlation was used to calculate the heat transfer coefficient. The

Φprimary=0.5238

0.523

0.482

0.522

0.481

0.521

Primary Secondary

0.520 0.519

0.480 0.479 0.478

0.518

0.477

Φsecondary=0.4762

0.517

0.476 0

40

80

120

160

200

240

280

320

Number of grids Fig. 7. Effect of grid number on dimensionless mean temperature profile of crossflow heat exchanger.

1.0 0.9 0.8

Effectiveness

0.483

Dimensionless mean outlet temperature of secondary fluid

Dimensionless mean outlet temperature of primary fluid

1.1 0.524

0.7 0.6 0.5

ε-NTU method PCHE analysis code C*=0.1 C*=0.1 C*=0.3 C*=0.3 C*=0.5 C*=0.5 C*=0.7 C*=0.7 C*=1 C*=1

0.4 0.3 0.2 0.1 0.0 0

1

2

3

4

5

6

7

8

9

10

NTU Fig. 8. Effectiveness-NTU comparison between e-NTU method and crossflow PCHE analysis code.

S.-J. Yoon et al. / International Journal of Heat and Mass Transfer 70 (2014) 250–263

Averaged temperature (oC)

320

k (-30%) cp(-30%) μ (-30%) Reference k (+30%) cp(+30%) μ (+30%)

310

300

290

800

+1.05% 750 o

dimensions in x, y, and z axial directions of heat exchanger were assumed to be 0.9897 m, 0.9897 m, and 0.634 m, respectively. Corresponding numbers of flow channels are 300, 300, and 100, respectively. Figs. 9–16 show the average temperature profiles according to the uncertainty of thermo-physical property for each coolant. The

Averaged temperature ( C)

258

700

-1.65%

650 600

Reference k (-30%) cp(-30%) μ (-30%) k (+30%) cp(+30%) μ (+30%)

550 500 450 400 350 0.0

0.2

0.4

0.6

0.8

1.0

Distance from the inlet (m) 280

+0.75%

270

Fig. 12. Average temperature profile of secondary fluid according to uncertainty of material property (Helium).

-0.47% 260 0.0

560 0.2

0.4

0.6

0.8

540

250

-0.77%

500 480 460 440 420

+0.7%

400 380

230

360 0.0

-0.47% k (-30%) cp(-30%) μ (-30%) Reference k (+30%) cp(+30%) μ (+30%)

220

210

200 0.0

0.2

0.4

0.6

0.8

Fig. 10. Average temperature profile of secondary fluid according to uncertainty of material property (Water).

950

k (-30%) cp(-30%) μ (-30%) Reference k (+30%) cp(+30%) μ (+30%)

900 850 800 750

0.4

0.6

0.8

1.0

Fig. 13. Average temperature profile in primary side according to uncertainty of material property (Sodium).

500

+0.37% 480

-0.54%

460 440 420

k (-30%) cp(-30%) μ (-30%) Reference k (+30%) cp(+30%) μ (+30%)

400 380 360 340 320 0.0

700

0.2

0.4

0.6

0.8

1.0

Distance from the inlet (m)

650

+2.38%

600 550

-1.52% 500 0.0

0.2

Distance from the inlet (m)

1.0

Distance from the inlet (m)

Averaged temperature (oC)

520

240

Averaged temperature (oC)

Averaged temperature (oC)

+0.48%

o

Fig. 9. Average temperature profile of primary fluid according to uncertainty of material property (Water).

Averaged temperature ( C)

Distance from the inlet (m)

260

k (-30%) cp(-30%) μ (-30%) Reference k (+30%) cp(+30%) μ (+30%)

1.0

0.2

0.4

0.6

0.8

1.0

Distance from the inlet (m) Fig. 11. Average temperature profile of primary fluid according to uncertainty of material property (Helium).

Fig. 14. Average temperature profile in secondary side according to uncertainty of material property (Sodium).

density change did not have an effect on the temperature profile. In the developed code, when the heat duty is given, the mass flow rate is determined from Eq. (30). Then, the density is used to calculate the fluid velocity and Reynolds number. Since the product of density and fluid velocity is constant by Eq. (31), the Reynolds number is not changed by the density. Consequently,

259

S.-J. Yoon et al. / International Journal of Heat and Mass Transfer 70 (2014) 250–263

660

650 0.0

k (-30%) cp(-30%) μ (-30%) Reference k (+30%) cp(+30%) μ (+30%) 0.2

+0.62%

-0.51% 0.4

0.6

0.8

1.0

Distance from the inlet (m) Fig. 15. Average temperature profile in primary side according to uncertainty of material property (FLiBe).

650

Averaged temperature (oC)

+0.53% 640

630

-0.64% Reference k (-30%) cp(-30%) μ (-30%) k (+30%) cp(+30%) μ (+30%)

620

610

600 0.0

0.2

0.4

0.6

0.8

1.0

Distance from the inlet (m) Fig. 16. Average temperature profile in secondary side according to uncertainty of material property (FLiNaK).

the uncertainty of fluid density does not have an effect on the temperature. Uncertainties of heat capacity and dynamic viscosity resulted in the same temperature profile. The temperature profile of the crossflow analytical model is determined by the size of heat exchanger and NTU value. In this crossflow model, NTU can be approximated as a function of the product of heat capacity and dynamic viscosity so that the changes in these two parameters resulted in the same degree of change in NTU. Uncertainty of thermal conductivity led to the maximum deviation of temperature profile, except the sodium case. The effect of thermal conductivity was relatively smaller than heat capacity and dynamic viscosity for the fluid of low Prandtl number. Overall, the relative deviation of the temperature profile according to the uncertainty of fluid property was very small. Maximum deviation was 2.38% that occurred by the thermal conductivity in the primary side of helium flow. In conclusion, the uncertainty of fluid property was negligible in the thermal design of crossflow PCHE. 3.4. Effect of heat transfer coefficient correlations Comparative analyses were carried out to evaluate the effect of heat transfer correlations on the temperature profile. In this analysis, the heat duty, the temperatures, and pressures in Table 1 were employed. Hastelloy N was employed as the structural material. The dimensions of the x, y, and z axial directions of the heat exchanger were assumed to be 0.9897 m, 0.9897 m, and 0.634 m, respectively.

360

o

670

Average temperatre ( C)

680

Gnielinski Berbish Dittus-Boelter Kim et al Lyon Lubarsky and Kaufman Reed Taylor and Kirchgessner, Wieland Wu and Little Wang and Peng

330

Lubarsky and Kaufman

300 Wang and Peng

ΔT=39 K

270 Wu and Little, Kim et al.

240 0.0

0.2

0.4

0.6

0.8

1.0

Distrance from the Inlet (m) Fig. 17. Temperature profile in primary side by the heat transfer correlations (Water).

280

o

690

Figs. 17 and 18 show the temperature profiles in the primary and secondary side of the water flows, respectively. The Lubarsky and Kaufman’s correlation showed the highest temperature of primary side and lowest temperature of secondary side. Wu and Little’s and Kim et al. correlations showed the lowest temperature of primary side and highest temperature of secondary side. These two correlations predicted the heat transfer coefficients similarly for Prandtl number in the range of 0.5–1.0. Thus, the same temperature profile trend was observed in the water and helium flows of which Prandtl numbers were approximately 1.0 and 0.66, respectively. The maximum deviation among the averaged outlet temperature at the outlet was 39.45 K, which was caused by the Lubarsky and Kaufman’s correlation and Kim et al. correlation. Figs. 19 and 20 show the temperature profiles of helium flows. In the helium flow, the similar trends of temperature profiles were observed as the water flow. The maximum deviation among the averaged outlet temperature was approximately 196.01 K which was caused by the Lubarsky and Kaufman’s correlation and Kim et al. correlation. Figs. 21 and 22 show the temperature profiles of sodium flows. In this case, the effect of heat transfer correlation was smaller than the other cases. For the fluids with very low Prandtl numbers, most of empirical correlations predicted low heat transfer coefficient and the difference among them was also small. Maximum deviation of averaged outlet temperature was 15.34 K, which was caused by Gnielinski’s correlation and Kim et al. correlation. Figs. 23 and 24 shows the temperature profiles of molten salt flows. The Prandtl

Average temperatre ( C)

Averaged temperature (oC)

700

260

Gnielinski Berbish Dittus-Boelter Kim et al Lyon Lubarsky and Kaufman Reed Taylor and Kirchgessner, Wieland Wu and Little Wang and Peng

Wu and Little, Kim et al.

ΔT=39 K

240 Wang and Peng

220 Lubarsky and Kaufman

200 0.0

0.2

0.4

0.6

0.8

1.0

Distrance from the Inlet (m) Fig. 18. Temperature profile in secondary side by the heat transfer correlations (Water).

260

S.-J. Yoon et al. / International Journal of Heat and Mass Transfer 70 (2014) 250–263 1000

550

800 Wang and Peng

700

500

400 0.0

Gnielinski Berbish Dittus-Boelter Kim et al Lyon Lubarsky and Kaufman Reed Taylor and Kirchgessner, Wieland Wu and Little Wang and Peng

0.2

0.4

ΔT=196 K Wu and Little Kim et al.

0.6

0.8

500

o

Lubarsky and Kaufman

600

Average temperatre ( C)

o

Average temperatre ( C)

900

450

350

0.2

Fig. 19. Temperature profile in primary side by the heat transfer correlations (Helium).

700

700 Wang and Peng

o

ΔT=196 K

Lubarsky and Kaufman

680 ΔT=31 K

660

640

400 0.2

0.4

0.6

0.8

620 0.0

1.0

Gnielinski Berbish Dittus-Boelter Kim et al Lyon Lubarsky and Kaufman Reed Taylor and Kirchgessner, Wieland Wu and Little Wang and Peng

0.2

Distrance from the Inlet (m)

Gnielinski

400

0.8

ΔT=31 K

620 Wang and Peng

600

Gnielinski Berbish Dittus-Boelter Kim et al Lyon Lubarsky and Kaufman Reed Taylor and Kirchgessner, Wieland Wu and Little Wang and Peng

580

ΔT=15 K

Berbish, Wu and Little, and Kim et al.

0.6

1.0

Lyon

640

o

Average temperatre ( C)

o

Average temperatre ( C)

450

0.4

0.8

660

Gnielinski Berbish Dittus-Boelter Kim et al Lyon Lubarsky and Kaufman Reed Taylor and Kirchgessner, Wieland Wu and Little Wang and Peng

0.2

0.6

Fig. 23. Temperature profile in primary side by the heat transfer correlations (Molten Salt).

550

350 0.0

0.4

Lyon

Distrance from the Inlet (m)

Fig. 20. Temperature profile in secondary side by the heat transfer correlations (Helium).

500

1.0

Kim et al. Wu and Little

Wang and Peng

0.0

0.8

720

600

500

0.6

Fig. 22. Temperature profile in secondary side by the heat transfer correlations (Sodium).

Average temperatre ( C)

o

Average temperatre ( C)

800

Gnielinski Berbish Dittus-Boelter Kim et al Lyon Lubarsky and Kaufman Reed Taylor and Kirchgessner, Wieland Wu and Little Wang and Peng

0.4

Distrance from the Inlet (m)

Distrance from the Inlet (m)

900

ΔT=15 K

400

300 0.0

1.0

Gnielinski Berbish Dittus-Boelter Kim et al Lyon Berbish, Wu and Little, and Kim et al. Lubarsky and Kaufman Reed Taylor and Kirchgessner, Wieland Wu and Little Wang and Peng Gnielinski

1.0

Distrance from the Inlet (m) Fig. 21. Temperature profile in primary side by the heat transfer correlations (Sodium).

number of molten salt flows was approximately 10.0. Contrary to the sodium flow, the molten salt is a fluid with a high Prandtl number. Not only the empirical correlations developed for liquid metals but also Dittus-Boelter correlation and Gnielinski’s correlations predicted high-heat transfer coefficient. Consequently, these

560 0.0

0.2

0.4

0.6

0.8

1.0

Distrance from the Inlet (m) Fig. 24. Temperature profile in secondary side by the heat transfer correlations (Molten Salt).

correlations predicted similar temperature profiles. Maximum deviation of averaged outlet temperature was 31.04 K, which was caused by Wang and Peng’s correlation and Lyon’s correlation. By comparing heat transfer correlations for each fluid, we found that the effect of heat transfer correlations was largest in the

261

S.-J. Yoon et al. / International Journal of Heat and Mass Transfer 70 (2014) 250–263 Table 8 Sizing and cost estimation results of crossflow PCHEs. Coolants

Heat exchanger size (m)

Average outlet temperature (°C)

Cost (USD) (Operation period: 20 year)

Lx

Ly

Lz

Primary

Secondary

Structural material

Capital

Operation

Water

0.396

1.518

0.634

295.3

293.6

Alloy 617 Alloy 800H Hastelloy N

3.8197  105 3.6278  105 4.1831  105

1.4049  109 1.4420  109 1.4146  109

Helium

1.128

2.095

637.8

925.3

Alloy 617 Alloy 800H Hastelloy N

1.5036  106 1.4280  106 1.6466  106

9.4799  109 9.4901  109 9.4826  109

Sodium

1.214

1.613

390.2

526.0

Alloy 617 Alloy 800H Hastelloy N

1.2459  106 1.1833  106 1.3644  106

1.2703  108 1.2770  108 1.2720  108

Molten Salt

6.352

11.880

651.0

690.2

Alloy 617 Alloy 800H Hastelloy N

4.7996  107 4.5585  107 5.2562  107

9.1158  106 9.1174  106 9.1163  106

In this crossflow PCHE analysis code, the size of heat exchanger is determined by iterative calculation to meet the given heat duty and temperature requirements. The heat duty and temperature requirements in Table 1 and three nickel-based alloys in Table 5 were used in the economic analysis. The maximum operating period of heat exchanger is assumed to be 20 years. In this analysis, the height of heat exchanger, Lz, was fixed to 0.634 m to reduce the number of cases. If the height of the heat exchanger can be varied, the length and width of the heat exchanger are determined for each height so that various heat exchanger sizes can be generated. Economic analysis results are summarized in Table 8. The sizes of heat exchangers using water, helium, and sodium coolants were similar to each other, whereas large heat exchanger size was required to meet the temperature requirements for molten salt coolants. The predicted size of crossflow PCHE for molten salt is deemed inappropriate for molten salt cooled SMR. To reduce the size of the heat exchanger, the heat duty per heat exchanger has to be reduced by increasing the number of heat exchangers. 7.4E6

Alloy 617 Alloy 800H Hastelloy N

Total cost per year (USD/yr)

7.35E6 7.3E6

Total cost per year (USD/yr)

3.5. Cost estimation results

4.755E8

Alloy 617 Alloy 800H Hastelloy N

4.752E8 4.749E8 4.746E8 4.743E8 4.74E8 4.737E8

3

6

9

12

15

18

21

Operation period (yr) Fig. 26. Total cost per year of crossflow PCHE (Helium/Helium).

6.8E6

Total cost per year (USD/yr)

helium flow. The deviation of temperature profiles by the heat transfer correlation decreased for very low or very high Prandtl number fluids. It was not easy to determine a heat transfer correlation that can be applied generally irrespective of coolant types. Consequently, the experimental and/or computational validations are deemed to be required to improve the validity and accuracy of crossflow PCHE analysis code. In addition, the appropriate heat transfer correlation for each fluid type should be applied.

Alloy 617 Alloy 800H Hastelloy N

6.7E6

6.6E6

6.5E6

6.4E6

6.3E6 3

6

9

12

15

18

21

Operation period (yr) 7.25E6

Fig. 27. Total cost per year of crossflow PCHE (Sodium/Sodium). 7.2E6 7.15E6 7.1E6 7.05E6 7E6 3

6

9

12

15

18

21

Operation period (yr) Fig. 25. Total cost per year of crossflow PCHE (Water/Water).

The capital cost of crossflow PCHE was relatively smaller than the operation cost except for molten salt heat exchanger. Due to the large size of the molten salt heat exchanger, its capital cost was estimated to be larger than operation cost. The capital cost of the molten salt heat exchanger was higher than the others, but the operation cost was much lower. The operation cost of the helium heat exchanger was the most expensive. Figs. 25–28 show the total cost per year of the heat exchangers according to the operation period. Operation cost per year decreased as the

262

S.-J. Yoon et al. / International Journal of Heat and Mass Transfer 70 (2014) 250–263

(Grant code: 2012-052255), under DOE Idaho Operations Office Contract DE-AC07-05ID14517.

1.2E7

Alloy 617 Alloy 800H Hastelloy N

Total cost per year (USD/yr)

1E7 8E6

References 6E6

4E6

2E6 3

6

9

12

15

18

21

Operation period (yr) Fig. 28. Total cost per year of crossflow PCHE (FLiBe/FLiNaK).

operation period increased. Total cost of the heat exchanger using Alloy 617 was cheaper than the other structural materials except for the molten salt heat exchanger. In the molten salt heat exchanger, the total cost was minimized by using Alloy 800H. The total cost per year of molten salt heat exchanger decreased rapidly as the operation period increased. 4. Conclusion In this study, the crossflow PCHE analysis code has been developed to evaluate the size and cost of heat exchangers by implementing the analytical solution of single pass, both unmixed fluids crossflow heat exchanger model. Two-dimensional temperature distribution of crossflow PHCE was calculated by the analytical solution. General methods for thermal design and cost estimation of heat exchangers were employed to determine the size and cost of heat exchanger, respectively. The information provided by this code can be used for the optimal design of advanced SMR system. The grid sensitivity test of the code showed that the effect of grid on the temperature profile was negligible. The effectiveness calculated by the code showed a good agreement with that by well-known e-NTU method. The uncertainty of fluid property was propagated into the temperature distribution, but its effect was small enough to be inconsequential. The heat transfer correlations had a considerable influence on the temperature profile. The temperature profile of crossflow PCHE using helium gases showed the largest deviation by the heat transfer correlations. The deviation of temperature profiles by the heat transfer correlation decreased for very low or very high Prandtl number fluids. However, it was not easy to determine the most accurate heat transfer correlation for the heat exchangers using various combinations of coolants. Consequently, experimental and/or computational validations are required to determine the best heat transfer correlation for each SMR. Costs of crossflow PCHE for the advanced SMR designs were investigated. Capital costs of PCHE for water, helium, and sodium flows were lower than their operation costs, whereas capital cost of molten salt heat exchanger was higher than operation cost. Total cost of heat exchanger using Alloy 617 was cheaper than the other structural materials except for the molten salt heat exchanger. In the molten salt heat exchanger, the total cost was minimized by using Alloy 800H. Acknowledgements This work was supported by the National Research Foundation of Korea (NRF) and grant funded by the Korean government (MSIP)

[1] Heatric™ Homepage, 2013. Available online at . [2] Lee, Won-Jae, 2010. The SMART Reactor, 4th Annual Asian-Pacific Nuclear Energy Forum, June 18–19. [3] M. Richards, A. Shenoy, K. Schultz, L. Brown, E. Harvego, M. Jean-PhillippeCoupey, S.M. Moshin Reza, F. Okamoto, H2-MHR conceptual designs based on the SI process and HTE, in: 3rd Information Exchange Meeting on Nuclear Production of Hydrogen and Second HTTR Workshop on Hydrogen Production Technologies, October 5–7, 2005, Japan Atomic Energy Research Institute, Oarai, Japan. [4] Hahn, Do.-Hee., Conceptual design of the sodium-cooled fast reactor KALIMER-600, Nucl. Eng. Technol. 39 (3) (2007) 193–206. [5] Greene, S.R., et al., 2010. Pre-Conceptual Design of a Fluoride-Salt-Cooled Small Modular Advanced High-Temperature Reactor (SmAHTR), ORNL Report, ORNL/ TM-2010/199, 2010. [6] R.K. Shah, A.J. Willmott, Thermal Energy Storage and Regeneration, Hemisphere/McGraw-Hill, Washington, DC, 1981. [7] Southall, D., Diffusion Bonding in Compact Heat Exchangers, Proceeding of ICAPP’09, Tokyo, Japan, May 2009, pp. 10–14. [8] The MathWorks Inc., MATLAB User’s Guide, The MathWork Inc., Natick, MA, 2012. [9] W. Nusselt, Der Warmeiibergang im Kreuzstrom, Z. Ver.dt Ing. 55 (1911) 2021–2024. [10] R.K. Shah, D.P. Sekulic, Fundamentals of Heat Exchanger Design, John Wiley & Sons Inc., Hoboken, New Jersey, 2003. [11] W. Nusselt, Eine neue Formel fur den Warmedurchgangim Kreuzstrom, Tech. Mech. Thermo-Dynam., Berl. 1 (1930) 17–422. [12] D.M. Smith, Mean temperature-difference in cross flow, Engineering 138 (1934) 479–481. 606–607. [13] A.M. Binnie, E.G.C. Poole, The theory of the single pass cross-flow heat interchanger, Proc. Cambr. Phil. Soc. 33 (1937) 403–411. [14] Mason, J.L., Heat Transfer in Crossflow, in: Proc. Second U.S. National Congress of Applied Mechanics, American Society of Mechanical Engineers, New York, 1955, pp. 801–803. [15] H. Kiihl, Probleme des Kreuzstrom-Warmeaustauschers, Springer, Berlin, 1959. [16] Schedwill, H., 1968. Termische Auslegung von Kreuzstromwarmeaustauchern, Fortsch.-Ber. VDI-Z, Reihe 6, Nr.19. [17] T. Ishimaru, N. Kokubo, R. Izumi, Performance of crossflow heat exchanger, Bull. Jup. Soc. mech. Engrs 19 (1976) 1336–1343. [18] F.E. Romie, Transient response of gas-to-gas crossflow heat exchangers with neither gas mixed, Trans. Am. Soc. Mech. Eng. J. Heat Transfer 105 (1983) 563– 570. [19] J. Lach, The exact solution of the Nusselt’s model of the cross-flow recuperator, Int. J. Heat Mass Transfer 26 (1983) 1597–1601. [20] B.S. Baclic, P.J. Heggs, On the search for new solutions of the single-pass crossflow heat exchanger problem, Int. J. Heat Mass Transfer. 28 (10) (1985) 1965–1976. [21] J.E. Hesselgreaves, Compact Heat Exchangers: Selection Design and Operation, Pergamon Press, 2001. [22] N.S. Berbish, M. Moawed, M. Ammar, R.I. Afifi, Heat transfer and friction factor of turbulent flow through a horizontal semi-circular duct, Heat Mass Transfer 47 (2011) 377–384. [23] Holmgren, M., 2003. X Steam for MATLAB, [24] National Institute of Standards and Technology, 2011. NIST Chemistry WebBook: NIST Standard Reference Database Number 69, . [25] International Atomic Energy Agency, 2008. Thermophysical Properties of Materials for Nuclear Engineering: A Tutorial and collection of Data, IAEATHPH, ISBN 978–92–0–106508–7. [26] Sohal, M.S., Ebner, M.A., Sabharwall, P., and Sharpe, P., 2010. Engineering Database of Liquid Salt Thermo-physical and Thermo-chemical Properties, INL/EXT-10-18297. [27] I.H. Kim, H.C. No, J.I. Lee, B.G. Jeon, Thermal hydraulic performance analysis of the printed circuit heat exchanger using a helium test facility and cfd simulations, Nucl. Eng. Des. 239 (2009) 2399–2408. [28] R.H.S. Winterton, Where did the Dittus-Boelter equation come from?, Int J. Heat Mass Transfer 41 (4-5) (1998) 809–810. [29] V. Gnielinski, New equation for heat and mass transfer in turbulent pipe and channel flow, Int. Chem. Eng. 16 (1976) 359–368. [30] R.N. Lyon, Liquid metal heat transfer coefficients, Chem. Eng. Prog. 47 (1951) 75–79. [31] B. Lubarsky, S.J. Kaufman, Review of Experimental Investigations of Liquid Metal Heat Transfer, NACA-Report 1270, Lewis Flight Propulsion Laboratory, 1956. [32] S. Kakac, R.K. Shah, W. Aung, Handbook of Single-Phase Convective Heat Transfer, John Wiley & Sons, 1987. [33] Taylor, M.F., Kirchgessner, T.A., 1959. Measurements of Heat Transfer and Friction Coefficients for Helium Flowing in a Tube at Surface Temperatures up to 5900oR, NACA TN D-133.

S.-J. Yoon et al. / International Journal of Heat and Mass Transfer 70 (2014) 250–263 [34] Wieland, W.F., 1962. Measurement of Local Heat Transfer Coefficients for Flow of Hydrogen and Helium in a Smooth Tube at High Surface to Fluid Bulk Temperature Ratios, Paper Presented at AIChE Symposium on Nuclear Engineering Heat Transfer, Chicago, Ill., December. [35] P. Wu, W.A. Little, Measurement of the heat transfer characteristics of gas flow in fine channel heat exchanger used for microminiature refrigerators, Cryogenics 24 (8) (1984) 415–420. [36] B.X. Wang, X.F. Peng, Experimental investigation on liquid forced-convection heat transfer through microchannels, Int. J.Heat Mass Transfer 37 (Suppl. 1) (1994) 73–82. [37] E.S. Kim, C.H. Oh, S. Sherman, Simplified optimum sizing and cost analysis for compact heat exchanger in VHTR, Nucl. Eng. Des. 238 (2008) 2635– 2647.

263

[38] Kim, I.H., 2012. Experimental and numerical investigations of thermalhydraulic characteristics for the design of a Printed Circuit Heat Exchanger (PCHE) in HTGRs, Ph.D. Thesis, KAIST. [39] Sabharwall, P., Kim, E.S., Siahpush, A., Anderson, N., Glazoff, M., Phoenix, B., Mizia, R., Clark, D., McKellar, M., Patterson, M.W., 2011. Feasibility Study of Secondary Heat Exchanger Concepts for the Advanced High Temperature Reactor, INL/EXT-11-23076. [40] L.F. Richardson, The Approximate Arithmetical Solution by Finite Differences of Physical Problems Involving Differential Equations, with an Application to the Stresses in a Masonry Dam, Philos. Trans. Roy. Soc. London Ser. A 210 (1910) 07–357. [41] A. Triboix, Exact and approximate formulas for cross flow heat exchangers with unmixed fluids, Int. Commun. Heat Mass Transfer 36 (2009) 121–124.