Open access peer-reviewed chapter

Heat Transfer Correlations of Supercritical Fluids

Written By

Fangbo Li, Binbin Pei and Bofeng Bai

Submitted: 03 June 2019 Reviewed: 26 August 2019 Published: 03 October 2019

DOI: 10.5772/intechopen.89356

From the Edited Volume

Advanced Supercritical Fluids Technologies

Edited by Igor Pioro

Chapter metrics overview

1,110 Chapter Downloads

View Full Metrics

Abstract

The drastic changes of thermophysical properties in the pseudo-critical region of supercritical fluids bring very big challenges to the traditional Dittus-Boelter-type heat transfer correlations. In this chapter, we will talk about the principles and applications of two kinds of heat transfer correlations of supercritical fluids: the empirical type and the semiempirical type. For the empirical correlations, the modification methods taking into account the variable properties and body force effects will be introduced. We will focus on the proposal of nondimensional parameters describing the buoyancy effect and flow acceleration; while for the semiempirical ones, we mainly talk about the new kind of correlation which is based on the momentum and energy conservations in the mixed convective flow.

Keywords

  • heat transfer correlation
  • empirical type
  • semiempirical type
  • buoyancy effect
  • flow acceleration

1. Introduction

Establishing an accurate heat transfer correlation of supercritical fluid is the prerequisite of the design of the apparatus and devices, such as the supercritical boilers, generation IV reactors, and heat exchangers [1, 2, 3, 4]. However, the drastic change of the thermodynamic properties in the pseudo-critical region and the complex mechanism of turbulent heat transfer lead to difficulties in predicting the heat transfer performance accurately. There are three distinguished heat transfer regimes for supercritical fluid: normal heat transfer (normal HT), improved heat transfer (improved HT), and deteriorated heat transfer (deteriorated HT). The normal HT regime is characterized by the heat transfer coefficients similar with those predicted by the Dittus-Boelter (DB) correlation at subcritical condition; the improved HT regime can be characterized by the higher heat transfer coefficients and lower wall temperatures compared with the normal HT; and the deteriorated HT regime is accompanied with lower heat transfer coefficients than the normal HT; hence, the wall temperature peak can be observed. Figure 1 illustrates the heat transfer performances of supercritical CO2 under different mass fluxes and heat fluxes [5] of the upward flow in a circular tube with an inner diameter of 6.32 mm and heated length of 2.65 m. At relatively lower heat flux/mass flux ratio, the heat transfer will be obviously improved due to the large specific heat in the pseudo-critical region. The heat transfer coefficients will be higher than the ones calculated by the Dittus-Boelter correlation. As the heat flux/mass flux ratio increases, the mixing convective heat transfer occurs, and the wall temperature peak caused by the heat transfer deterioration can be observed. The experiment conducted by Kurganov and Kaptil’ny [6] offered an insight into the mechanism of heat transfer deterioration, the M-shaped velocity profile was observed, and it was considered to be closely associated with the flattened velocity gradient, reduced turbulent shear stress, and turbulent kinetic energy due to the body force (also known as the buoyancy effect). Bae et al. [7, 8] obtained the budgets of the turbulent kinetic energy in the deteriorated heat transfer by using direct numerical simulation, and the mechanism of buoyancy and flow acceleration were analyzed.

Figure 1.

Heat transfer performances of supercritical CO2 of the upward flow under different (a) mass fluxes and (b) heat fluxes in a circular tube with an inner diameter of 6.32 mm and heated length of 2.65 m. Symbols: G, mass fluxes (kg m−2 s−1); P, pressure (MPa); q, wall heat flux (kW m−2); H, enthalpy (kJ kg−1). Data in this figure come from Ref. [1].

The heat transfer correlations are established on the basis of heat transfer mechanism, and a vast majority of correlations are intended for use only for normal HT and improved HT regimes. They can be categorized into two types: the empirical type and the semiempirical type. In the empirical correlations, the correction terms composed of different thermophysical properties (such as the density, thermal conductivity, specific heat, and viscosity) are introduced to the heat transfer correlation of the constant-property fluid (i.e., DB correlation) [9]. However, the performance of the empirical correlations deteriorates with the increasing of the heat flux/mass flux ratio; thus, the empirical correlations considering the nondimensional numbers which reflect the buoyancy and thermally induced flow acceleration effects have also been proposed [10]. It is shown that the predictions on the mixing convection cases can be improved by this method. In order to further improve the performances of the empirical correlations under the strong buoyancy effect, some researchers tried to establish the semiempirical correlations based on the theoretical analysis approaches [11, 12]. The existing semiempirical correlations mainly aim to derive the qualitative relationship between the heat transfer impairment and turbulent shear stress reduction; the coefficients appeared in the correlation are obtained by fitting the experimental data.

In this chapter we mainly talk about the heat transfer correlations of supercritical fluid. Both the empirical type and the semiempirical type will be introduced here. We mainly focus on the principles, applications, and comparisons of these correlations.

Advertisement

2. Empirical heat transfer correlations

For the improved HT and normal HT regimes, abundant empirical correlations have been proposed, and most of them can give satisfying predictions [13, 14, 15]. Morky et al. [9] and Gupta et al. [16] proposed the correlations based on the Buckingham Π theorem, and the comparison with the experimental results confirmed their capability on the calculating the heat transfer improvement regime. However, as the increasing of the heat flux, the effect of the mixing convection increases. The fluid near the heated wall is accelerated under the effect of the buoyancy force and the flow acceleration, leading to the attenuation of heat transfer [17]. On the basis of this, researchers tried to modify the correlations by considering the nondimensional numbers reflecting the buoyancy force and the flow acceleration. Bae and Kim [10] and Komita et al. [18] analyzed the momentum balance under the mixed convection case and proposed a nondimensional parameter Bu to describe the effect of the buoyancy on the shear stress; then, they proposed a new function to describe the relationship between the Nusselt numbers and Bu based on the experimental data. This heat transfer correlation takes the form of the piecewise function divided by the value of the Bu number. Deev et al. [19] discussed the heat transfer of supercritical water in the channel. Two nondimensional criteria considering the effects of the viscous force and inertial force on heat transfer were proposed, and the weight constant treating the superposition between the forced and natural convection was introduced. Such practices were also conducted by Cheng et al. [20], Yu et al. [21], and Kuang et al. [22]. In their work, the dimensionless numbers such as the buoyancy number and the acceleration number were proposed to correct the deviation of the real heat transfer from that of the normal HT regime. Figure 2 shows the approach establishing the empirical correlations. The dimensionless groups are firstly chosen, and the coefficients in the correlation can be obtained by the linear regression analysis. Through variable transformations the dimensionless groups are reduced, and the final form can be obtained. Some typical empirical correlations are listed in Table 1.

Figure 2.

The approach deriving the empirical correlations.

Ref.FluidParameters rangeCorrelation
Mokry et al. [9]H2OP = 22.8–29.4 MPa
G = 200–1500 kg m−2 s−1
q = 70–1250 kW m−2
d = 3–38 mm
Nub=0.0061Reb0.914Pr¯b0.654ρw/ρb0.518
Pr¯b=cp¯μbλb,cp¯=HwHbTwTb
Gupta et al. [16]H2OP = 24 MPa
G = 200–1500 kg m−2 s−1
q = 70–1250 kW m−2
Nuw=0.004Rew0.923Pr¯w0.773μw/μb0.366ρw/ρb0.186
Jackson [23]H2OAfter review of the existing literatures and dataNub=0.0183Reb0.82Prb0.5ρw/ρb0.3cp¯/cpbn
where the exponent n is taken as
i. n=0.4+0.2Tw/Tpc115Tb/Tpc1 for Tpc < Tb < 1.2Tpc;
ii. n=0.4+0.2Tw/Tpc1 for Tb < Tpc < Tw;
iii. n = 0.4 for Tb < Tw < Tpc and 1.2Tpc < Tb < Tw.
Bae and Kim [10]CO2P = 7.75–8.55 MPa
G < 1200 kg m−2 s−1
qw < 150 kW m−2
Nub=NuvarpfBuNuvarp=0.021Reb0.82Prb0.5ρw/ρb0.3cp¯/cpbn i. For 5 × 10−8 < Bu < 7 × 10−7, fBu=1+108Bu0.032
ii. For 7 × 10−7 < Bu < 1 × 10−6, fBu=0.0185Bu0.435
iii. 1 × 10−6 < Bu < 1 × 10−5, fBu=0.75
iv. For 1 × 10−5 < Bu < 3 × 10−5,fBu=0.0119Bu0.36
v. For 3 × 10−5 < Bu < 1 × 10−4,fBu=32.4Bu0.4
Cheng et al. [20]H2Od = 10, 20 mm
P = 22.5–25 MPa
G < 3500 kg m−2 s−1
qw < 2000 kW m−2
Tb = 300–450°C
Nub=0.023Reb0.8Prb1/3F
F=minF1F2
F1=0.85+0.776103πA2.4
F2=0.48103πA,pc1.55+1.211πAπA,pc
πA=βcpqG
Deev et al. [19]H2Od = 2.67–12.0 mm
P = 22.6–26 MPa
G < 2000 kg m−2 s−1
qw = 200–2250 kW m−2
Nub=0.023Reb0.8Prb0.4ρw/ρb0.25cp¯/cpbnYY=1ξY1+ξY2
Y1 = 1 + a1exp(b1Kh2 + c1Kh)
Y2 = 1 + a2exp(b2Kh2 + c2Kh)
ζ = exp(−0.5(KAm)2)
Kh = KAm(HbHpc)/Hpc
a1 = 1.5, b1 = −30, c1 = 3
a2 = −0.15, b2 = −125, c2 = −25
KAm = (10−5Repc)0.5/(103Kqm)
Kqm = qwβpc/(Gcp,pc)
Yu et al. [21]H2OAfter review of the existing literatures and dataNub=0.01378Reb0.9078Pr¯b0.6171ρw/ρb0.4356Gr0.012q+0.0605Pr¯b=cp¯μbλb,cp¯=HwHbTwTb,Gr=d4qwλbνb2,q+=βqwGcp¯
Kuang et al. [22]H2OAfter review of the existing literatures and dataNub=0.0239Reb0.759Pr¯b0.833ρw/ρb0.31λw/λb0.0863μw/μb0.832Gr0.014q+0.021Pr¯b=cp¯μbλb,cp¯=HwHbTwTb,Gr=d4qwλbνb2,q+=βqwGcp¯

Table 1.

Typical empirical heat transfer correlations of supercritical fluids.

Advertisement

3. Semiempirical heat transfer correlations

As is mentioned before, as the heat flux increases, the heat transfer deterioration phenomenon will happen. Heat transfer deterioration is characterized with lower values of the heat transfer coefficients and very high wall temperatures. A widely used quantitative expression to define the heat transfer deterioration of supercritical fluid [24, 25, 26] is shown by Eq. (1):

h<chDB=c0.023λbdReb0.8Nub0.4E1

where h refers to the heat transfer coefficient of the heat transfer deterioration and hDB is the heat transfer coefficient given by the Dittus-Boelter correlation (i.e., normal HT regime). The coefficient c is taken as 0.3 by Schatte et al. [24]. Unfortunately, under the deteriorated HT regime, most of the empirical correlations will lose their accuracy. When deteriorated HT regime occurs, the dominated mechanism of the turbulent heat transfer is the interaction among the buoyancy, flow acceleration, and variable thermophysical properties. Unfortunately the empirical correlations can hardly reflect the heat transfer mechanism. Thus, their performances are unsatisfying, as is shown in Figure 3.

Figure 3.

The comparison among the results of the empirical correlations listed in Table 1 and the experimental results from Hu [27]. This case describes the upward flow of supercritical water in a circular tube with an inner diameter of 26 mm and heated length of 2.0 m. Data in this figure come from Ref. [11].

3.1 Correlation by Jackson

Jackson [12] investigated the semiempirical correlation considering the effect of the mixed convective heat transfer in a vertical heated tube. The author proposed an equivalent “buoyancy-free forced convective flow” which has the same heat transfer effect with the real supercritical fluid flow. He tried to establish the relationship between the heat transfer coefficients attenuation and the shear stress reduction based on the equivalent Nusselt and Reynolds numbers of the equivalent flow with the aid of the empirical equation under forced convective condition. Integrating the equations of momentum across the entire flow and the near-wall layer of thickness δt, respectively, then reorganizing the two equations leads to

τwτδt=±δtρbρavegE2

where τw is the wall shear stress, τδt is the shears stress at y = δt, and the integrated average value of the density ρave across the near-wall layer is defined by ρave=0δtρdy/δt. In Eq. (2) the positive sign indicates the upward flow, and the negative one is applied for the downward flow. Then the author assumed that the effect of the buoyancy in modifying the distribution of shear stress and turbulence production has led to such an “equivalent” flow which is not influenced by the buoyancy but has a bulk velocity which is either higher or lower than the real flow. The wall shear stress τw for this “equivalent” buoyancy-free flow can be replaced by the modified shear stress τδt at the location y = δt in a turbulent shear flow. Then the author proposed a model which reflects the shear stress distributions in the supercritical fluid flow:

τδt/τw=Reb/Reb2m1FVP1/FVP1E3

where the FVP1 and FVP1 are the modifications on the variable properties for the “equivalent” buoyancy-free flow and the real flow, respectively. The Nusselt number of the “equivalent” forced convection can be expressed in terms of the Reynolds number, Reb, using the empirical equation:

NuB=K2Rebm2Prbn2FVP2E4

where K is the constant and then

NuF=K2Rebm2Prbn2FVP2E5

Combining Eqs. (4) and (5) follows that

NuB/NuF=Reb/Rebm2FVP2'/FVP2E6

Combining Eqs. (2) and (5) the author obtained

NuB/NuF=FVP2/FVP1m4/FVP2/FVP1m4τδt/τwm4E7

Or equivalently

NuB=K2Rebm2Prbn2FVP2FVP3'/FVP3τδt/τwm4E8

The above equation shows the relationship between the heat transfer and the reduced shear stress across the near-wall layer. Based on this, the author considered three cases of mixing convective heat transfer in a vertical heated tube to fluid flowing either in the upward direction (i.e., the buoyancy-aided case) or the downward direction (i.e., the buoyancy-opposed case). Firstly, a model of the buoyancy-influenced heat transfer was proposed:

NuBNuFFVP3FVP3=1CBBobFVPBNuBNuFFVP3FVP32.10.45E9

where CB is a constant, Bob is the buoyancy number, and FVPB is the modification on the variable properties for the buoyancy-influenced flow. As can be seen from Eq. (9), NuB/NuFFVP3/FVP3 is a unique function of a single parameter CBBobFVPB, and the author referred to this function as ψCBBobFVPB. It can be evaluated by assigning a series of values to the product CBBobFVPB in Eq. (9) and calculating the corresponding values of NuB/NuFFVP3/FVP3. Figure 4(a) shows the effect of the buoyancy on heat transfer predicted by Eq. (9) by the three curves (the red line for the downward flow and the black line for the upward flow). Note that by Figure 4(a), we can see if CBBobFVPB is less than 0.04, the function ψCBBobFVPB will be within about 2% of unity; thus, the heat transfer mode will be the variable-property forced convective heat transfer.

Figure 4.

Effect of the theoretical buoyancy and flow acceleration factors on heat transfer [12], (a) the effect of buoyancy, and (b) the effect of flow acceleration. Data in this figure come from Ref. [12].

The second model proposed by the author was about the heat transfer due to the thermally induced flow acceleration. It shows how the acceleration of a heated flow causes the reduced turbulence production, the deteriorated HT, and the turbulence laminarization. The final form of this model is given as

NuANuFFVP3FVP3=1CAAcbFVPANuBNuFFVP3FVP3'1.10.45E10

where CA is a constant, Acb is the flow acceleration number, and FVPA is the modification on the variable properties for the acceleration-influenced flow. As is shown in Eq. (10), the product NuA/NuFFVP3/FVP3 is a unique function of the parameter CAAcbFVPA, and we will describe this function as ψCAAcbFVPA. The variation of ψACAAcbFVPA with CAAcbFVPA can be readily determined using a similar approach to that described in the buoyancy-influenced heat transfer case. The result is shown in Figure 4(b). The function ψCAAcbFVPA is predicted to fall from unity as CAAcbFVPA increases from zero. This indicates that the heat transfer is attenuated as the flow acceleration increases, and note that this phenomenon is independent on the flow direction. The author argued that with the increase of the flow acceleration, the excess pressure acting on the flow is able to balance the shear stress; thus, the core layer will not experience any shear stress and soon leads to the turbulence laminarization.

The third model was presented which describes the heat transfer under the condition where the influence of the buoyancy and the thermally induced bulk flow acceleration are combined together. Using the approach developed earlier to relate the Nusselt number ratio and the shear stress ratio, the author proposed

(NuBANuF)(FVP3FVP3)1.1= [ 1(CBBobFVPB)(NuBANuF)2.1(FVP3FVP3)2.1 r(CBBobFVPB)(NuBANuF)2.1(FVP3FVP3)2.1 ]0.45E11

Here the parameter r indicates the ratio of NuBA/NuF1.1CAAcbFVPA to NuBA/NuF2.1CBBobFVPB (i.e., the ratio of the strengths of the buoyancy effect to the flow acceleration effect). In Figure 5, the author evaluated the parameter NuBA/NuFFVP3/FVP3 for the specified values of CBBobFVPB and r and produced the curves of ϕBACBBobFVPB=NuBA/NuFFVP3/FVP31.1 against CBBobFVPB and r. Thus, for the upward flow, the acceleration influence promotes the buoyancy effect, whereas for the downward flow, the acceleration opposes the buoyancy effect. These trends are illustrated in Figure 5 for three values of r = 0, 0.5, and 0.8.

Figure 5.

Heat transfer for mixed flow with combined influence of flow acceleration and buoyancy effect [12]. Data in this figure come from Ref. [12].

3.2 Correlation by Li and Bai

Li and Bai [11] proposed a semiempirical correlation based on the momentum and energy integration equations in the thermal boundary layer. The authors observed the direct numerical simulation (DNS) data under deteriorated HT regime in Bae et al. [7], Hu [27], Peeters et al. [28], and Zhang et al. [29], and they found that in the turbulent core, the temperature gradient is negligible; therefore, the fluid temperature in the core is approximately the bulk temperature, especially when Tb < Tpc < Tw. While in the thermal boundary layer δ, the fluid temperature increases from Tb to Tw (where δ is the thickness of the thermal boundary layer). This can be explained as follows: the fluid within the pseudo-critical region absorbs considerable energy due to the high cp, leading to the small temperature gradient in the turbulent core [7]. Based on these characteristics, Li and Bai built a “two-layer model” for deteriorated HT (as is shown in Figure 6) [11, 30]. The turbulence in the tube is divided into two layers: the thermal boundary layer and the core layer. In the thermal boundary layer, the buoyancy is strong and leads to the shear stress redistribution. However, in the core layer, the temperature and density gradients are small; hence, the buoyancy effect can be neglected.

Figure 6.

Physical model of the straight tube.

Figure 6 shows the straight tube the authors analyzed, and the flow direction is vertically upward. As is shown, the element abcd with a thickness of δ and a length of dx in the thermal boundary layer is chosen to build up the momentum equation:

dpdxτwδ+τδδρ¯g=1δddx0δρu2dyE12

In the above equation, τw is the shear stress at the wall and ρ¯ is the average density across the thermal boundary layer. The wall shear stress and friction factor here are calculated based on Blasius equation, as is shown in Eqs. (13) and (14):

τw=0.5fbρbub2E13
fb=0.079Reb0.25E14

Assume that the pressure gradient on the cross section keeps constant. In a similar way, for the bulk flow, we could get

dpdx2τwRρbg=ρubdubdx=GdubdxE15

Subtracting Eq. (15) and Eq. (12), the pressure gradient could be eliminated and the following result could be obtained:

2τwRτwτδδ+ρbρ¯g=1δddx0δρu2dyGdubdxE16

It is obvious that for the vertically upward flow, the radial part of the velocity (ur) is small enough to be neglected. Thus, the term ρu in the integration sign equals to the total mass flux G. Then Eq. (16) can be simplified as

2τwRτwτδδ+ρbρ¯g=Gδddx0δudyGdubdxE17

Note that the thickness of the thermal boundary layer δ is negligible compared with the radius of the tube R. Concerning with this fact

τwτδδ=ρbρ¯g+GdubdxGδddx0δudyE18

The effects of the buoyancy and flow acceleration on the shear stress redistribution are reflected by the two terms at the right side of Eq. (18), respectively. According to Negoescu et al. [31], in the upward flow, the influence of the flow acceleration is much weaker compared with the buoyancy. In order to simplify Eq. (18), the authors adopted the assumption that the axial velocity ub linearly varies within the thermal boundary layer thickness δ. This leads to

τwτδδ=ρbρ¯g+GdubdxG2δubdδdx+δdubdxE19

The gradient of the bulk velocity can be obtained by the analysis on the energy balance of the bulk fluid:

dubdx=Gd1/ρbdTbdTbdx=GρbβbcpbdHbdx=4qwβbρbcpbdE20

The authors made further assumption that δ(dub/dx) in Eq. (19) is much smaller than ub(dδ/dx). This can be interpreted as follows: the axial thickness variation of the thermal boundary layer (δ) is negligible compared with the axial bulk velocity (ub) variation. Then Eq. (19) could be simplified as

τwτδτw=2ρbρ¯fbρbub2+δfbubdubdxE21

Thus, the drop of the shear stress across the thermal boundary layer can be rewritten as the following pattern:

τwτδτw=2Gr¯bReb2+4kTRebPrbδfbdE22

where Grb¯ is the average Grashof number and kT is a nondimensional parameter reflecting the expansion of the fluid:

Gr¯b=ρbρbρ¯gd3μb2,ρ¯=12ρw+ρb for Tw<TpcorTb>Tpc1TwTbρbTpcTb+ρwTwTpc for TbTpcTwE23
kT=βbqwdλbE24

Therefore, Eq. (22) can also be rewritten as

τwτδτw=2Gr¯bReb2+4βbdqwRebμbcpbδfbdE25

In order to build the connection between the shear stress and heat transfer, the authors analyzed the energy balance of the element abcd shown in Figure 6:

qw+ρbubddx0δcpTdy+cpbTbddx0δρudy=λbdTdyy=δμtPrtHyy=δE26

The physical interpretation of the budgets in the above equation can be given as follows: the first term is the wall heat flux (i.e., energy input from the boundary ad); the second term is the net energy input by the convection through the boundaries ab and cd; the third term is the fluid energy from the core layer through the boundary bc. The axial heat conduction is ignored here. The fourth and fifth terms are the energy output by the heat conduction and turbulence diffusion through the boundary bc, respectively. As is indicated by Kurganov and Kaptil’ny [6], the turbulent viscosity in the thermal boundary layer drastically attenuates under the deteriorated HT regime. A factor Cμ1 is introduced here to reflect the ratio of the turbulent viscosities under the deteriorated HT regime and normal HT regime. Moreover, based on the mixing length assumption, the turbulent viscosity at the boundary bc is

μt=Cμρbl2uyy=δ=Cμρbl2τδμt,μt=Cμ1ρbl2τδ1/2E27

In order to give the qualitative description of μt, the authors applied the Bo* number and ρw/ρb into Eq. (27) to reflect the influence of the buoyancy force and the variable thermophysical properties:

Cμ1=φBobmρwρbn,Bob=Gr¯b/Reb2.7E28

where m and n are the exponents. Then the Nikuras-Van Driest mixing length model was chosen here [32]:

lm/R=0.140.08η20.06η41expy+/26,η=Ry/RE29

The wall distance y+ was taken as the dimensionless thickness of the thermal boundary layer, and it is defined as

y+=ρluτμly=ρlτwμlyE30

For the variable-property turbulence, the local value of the thermophysical properties is more advantageous [33, 34]. Therefore, in Eq. (30) the local definitions of the density and viscosity were adopted. Considering the thermal boundary layer thickness (δ) minuteness, the authors omitted the terms higher than the second order in the Taylor series of the two terms at the right side of Eq. (26):

λbdTdyy=δλbTwTbδ,μtPrtdHdyy=δμtPrtHwHbδE31

The terms on the left side of Eq. (26) can be dealt as follows:

ρbubddx0δcpTdyρbubddxHaveδ=GδdHavedx+HavedδdxE32
cpbTbddx0δρudy=GHbdδdxE33

where Have is the average specific enthalpy across the thermal boundary layer. Then the heat transfer coefficient can be obtained based on Eq. (26):

h=λbδ+μtcp¯PrtδGTwTbδdHavedx+Havedδdx+HbdδdxE34

The μt in Eq. (26) is determined by Eq. (25) and Eqs. (20)(22). The fitting curve among Cμ1, Bo* number, and the density ratio (ρw/ρb) is obtained by the dataset of the vertically upward water flow in the circular tube under the deteriorated HT reported by Mokry et al. [9], Hu [27], Wen [35], and Wang et al. [36]. The total number of the data points is 700, with 495 data points as the fitting set and 205 data points as the assessment set. The parameters range of the dataset is given in Table 2.

Refs.P (MPa)G (kg m−2 s−1)q (kW m−2)d (mm)Hb (kJ kg−1)Heated length (m)
Mokry et al. [9]24.1129–334200–500101450–26004.0
Hu [27]23–30300–500600–900261300–29002.0
Wen [35]23–26332–700446–6007.61300–29002.64
Wang et al. [36]23450–536450–458101300–24002.5

Table 2.

The experimental data sources used in the paper.

In Eq. (26), the thermal boundary layer thickness δ is expressed as

δ=μbρbτwδ+E35

According to the analysis on the temperature profiles based on the data presented in the Ref. [29] which has similar working conditions, the dimensionless excess temperature exceeds 0.9 at δ+ = 300 in the tube turbulence (for details, see Ref. [11]). For simplicity, the authors assumed an approximate value of 300 for δ+, and then the expression of δ can be expressed as

δ=δ+μb/ρbτw=300μb/ρbτwE36

According to Eqs. (36), (13), and (14), it can be found that the thickness of the thermal boundary layer δ=300μb/ρbτw104m is R ≈ 100δ. Thus, the assumption that the thermal boundary layer thickness is negligible compared with the tube radius is reasonable. The order of magnitude of the terms in Eq. (19) is listed as follows based on the dataset:

ub1ms1,G103kgm2s1,δ=300μb/ρbτw104m,dδdx=1321G0.875d0.125μ0.125dμdTdTbdx106,ubdδdx106ms1E37
qw5×105Wm2,ρb5×102kgm3,d102m,β101K1,cp104Jkg1K1,dubdx1s1,δdubdx104E38

The authors got that the ub(dδ/dx) could be ignored compared with the term δ(dub/dx). Therefore, the assumption simplifying Eq. (19) is reasonable. Similarly, the order of magnitude of the parameters in Eqs. (32) and (33) could be analyzed as follows:

G103kgm2s1,qw5×105Wm2,H106Jkg1,d102m,dHavedxdHbdx=4qwGd105Jkg1E39
δ=300μb/ρbτw104m,dδdx=1321G0.875d0.125μ0.125dμdTdTbdx106E40

Therefore,

ρbubddx0δcpTdyGδdHavedx+Havedδdx104Wm2E41
cpbTbddx0δρudy=GHbdδdx103Wm2E42

The second and third terms at the left side of Eq. (26) are far less than the qw in the left side; thus, they could also be ignored. Eq. (34) could be simplified as

h=λbδ+μtcp¯PrtδE43

Plugging Eqs. (25), (27)(30), and (36) into Eq. (43), the final form of the model could be obtained after some arrangements:

Nub=CAReb0.8751+Cμ1CBReb0.875Pr¯bPrtlmd1Cd2Gr¯bReb2.625+4kTReb1.625Prb0.5E44

where CA=6.67×104,CB=0.2,Cd=1.9×104,Prt=0.9. The fitting equation of Cμ1 was obtained as follows:

For 3×106<Bomax<8×105,

Cμ1=1.115Bob0.147ρw/ρb1.365E45

For 8×105<Bomax<3×104,

Cμ1=0.0016Bob0.416ρw/ρb1.325E46

Here the parameter Bomax is defined as the maximum of the buoyancy number (Bo*) under a specified case (i.e., specified P, G, qw, d). The range of the parameters in the new correlation is P = 23–30 MPa, G = 200–900 kg m−2 s−1, q = 129–700 kW m−2, d = 7.6–26 mm, 3 × 10−6 < Bomax < 3 × 10−4.

Figure 7 depicts the results of the model evaluation based on the above assessing procedure. It is clear that all the existing correlations could not obtain the tendencies of the wall temperature under the heat transfer deterioration regime. The semiempirical model proposed by Li and Bai shows comparatively higher accuracy compared with the existing correlations, especially at the peak region of the wall temperature. Most of the existing correlations significantly overestimate the heat transfer coefficients under the heat transfer deterioration regime.

Figure 7.

Comparisons of the predicted results and experimental results. Experimental data come from Hu [27], as is shown in Table 2; (a) and (c) are the wall temperature results; (b) and (d) are the heat transfer coefficients results. Data in this figure come from Ref. [11].

In Figure 8, the error graphs of different heat transfer correlations (here the fitting set and the assessment set are both included) are shown. It is shown that the semiempirical correlations introduced here are correlated well with the heat transfer coefficients dataset. Seventy-two percent and 63% of the data points fall into the error bars of ±30% and ±25%, respectively. In order to further investigate the predicting capacity under the deteriorated HT regime, the authors introduced another two statistical parameters: the mean relative error (MRE) and the root mean square error (RMSE) to evaluate the different correlations performance. Their definitions are given in Eqs. (47)(49), and the comparison results are shown in Table 3. As can be seen, the Li and Bai correlation can give the smallest prediction errors, thus proving its suitability and superiority.

Figure 8.

The errors of the predicted results of different heat transfer correlations. Data in this figure come from Ref. [11]. (a) comparison among Jackson correlation, Bae correlation and Li & Bai correlation; (b) comparison among Morky correlation, Yu correlation, Kuang correlation and Li & Bai correlation; (c) comparison between Cheng correlation and Li & Bai correlation.

CorrelationMRE/%RMSE/%
Mokry et al. [9]4496
Bae et al. [10]63279
Jackson [23]130180
Cheng et al. [20]94148
Yu et al. [21]10163
Kuang et al. [22]65113
Li and Bai [11]2732

Table 3.

The statistics of the present and existing correlations.

REi=EXPiPREiEXPiE47
MRE=1ni=1nREiE48
RMSE=1ni=1nREi2E49
Advertisement

4. Conclusions

The approach to establishing the heat transfer correlations of supercritical fluids is a critical issue since the correlations play very important role in the design and optimization of the systems and devices. In this chapter, we have discussed the principles and applications of the heat transfer correlations of supercritical fluids. The modeling approaches of the correlations of supercritical fluid heat transfer are reviewed, including the nondimensional parameters applied on the modification of the empirical correlations and the “equivalent buoyancy-free flow method” used for the semiempirical correlations.

Then we introduce a new physically based semiempirical correlation which is based on the momentum and energy conservations in the mixing convective flow. Considering the mechanism of heat transfer deterioration, a physical model characterizing the redistribution of the shear stress under the combined effect of buoyancy and flow acceleration was obtained. Then the model about the heat transfer coefficients under the influence of the reduced shear stress was derived by the energy integration equation within the thermal boundary layer. Based on this, a semiempirical heat transfer correlation was proposed and then verified with a wide range of experimental data. Compared with the existing correlations, the prediction accuracy of this newly developed correlation is significantly improved under the heat transfer deterioration regime. The investigation on the different statistical parameters shows that this semiempirical correlation is superior to the empirical ones.

Advertisement

Acknowledgments

This study is supported by the National Key Research and Development Program of China (no. 2016YFB0600100) and the China National Funds for Distinguished Young Scientists (no. 51425603).

Advertisement

Conflict of interest

No conflict.

Advertisement

Abbreviations

DB

Dittus-Bolter

DNS

direct numerical simulation

EXP

experimental results

HT

heat transfer

MRE

mean relative error

PRE

results given by correlations

RE

relative error

RSME

root mean square error

Advertisement

Appendices and nomenclature

Ac

flow acceleration parameter

Bo*

Jackson buoyancy parameter

Bom*

the maximum Jackson buoyancy number under a specified working condition

Bu

Bae and Kim buoyancy parameter

c1, c2, ck

exponents in the empirical correlations (in Figure 2)

cp

specific heat (J kg−1 K−1)

cp¯

average specific heat in the thermal boundary layer (=(HwHb)/(TwTb)) (J kg−1 K−1)

Cμ, Cμ1

the correction coefficients under heat transfer deterioration in the Li and Bai correlation

CA, CB, Cμ1, Cd, m, n

constants in the Li and Bai correlation

d

inside diameter of the circular tube (m)

f

Fanning resistance coefficient

F, F1, F2

constants in the Cheng correlation

FVP1,FVP2,FVP3

the modifications on the variable properties for the “equivalent” buoyancy-free flow

FVP1,FVP2,FVP3

the modifications on the variable properties for the real flow

FVPA

the modification on the variable properties for the acceleration-influenced flow

FVPB

the modification on the variable properties for the buoyancy-influenced flow

g

gravitational acceleration (m s−2)

G

mass flux (kg m−2 s−1)

Gr

Grashof number

h

local heat transfer coefficient (kW m−2 K−1)

H

enthalpy (kJ kg−1)

kT

the dimensionless number reflecting the expansion of the fluid

lm

the mixing length (m)

m1, m2, m4, n2

the exponents in the Jackson’s semiempirical model

Nu

Nusselt number

P

pressure (MPa)

Pr

Prandtl number

Prb¯

average Prandtl number across the thermal boundary layer

Prt

turbulent Prandtl number

q

wall heat flux (kW m−2)

r

radial coordinate (m)

R

tube radius (m)

Re

Reynolds number

T

temperature (K)

X, Y

nondimensional parameters in the empirical correlation (in Figure 2)

Y, Y1, Y2, Kh, KAm, Kqm

constants in the Deev correlation

y

distance from the wall (m)

y+

nondimensional distance from the wall

Greek symbolsλ

thermal conductivity (W m−1 K−1)

β

thermal expansion rate (K−1)

ρ

density (kg m−3)

μ

dynamic viscosity (Pa s)

δ

thermal boundary layer thickness (m)

δ+

nondimensional thermal boundary layer thickness in wall units

θ

nondimensional excess temperature

φ

the constant in the Li and Bai correlation

ψ

the functional expression for the acceleration-influenced or buoyancy-influenced heat transfer

τ

shear stress (Pa)

SupercriptsT

transpose matrix

SubscriptsA

flow acceleration effect

ave

average value over the thermal boundary layer

B

buoyancy effect

BA

combined effect of the buoyancy and flow acceleration

b

bulk value

max

maximum

F, f

forced convection

l

definition based on the local values

t

turbulence

pc

pseudo-critical point

w

values at the wall

References

  1. 1. Zhang Q, Li HX, Zhang WQ, Li LX, Lei XL. Experimental study on heat transfer to the supercritical water upward flow in a vertical tube with internal helical ribs. International Journal of Heat and Mass Transfer. 2015;89:1044-1053. DOI: 10.1016/j.ijheatmasstransfer.2015.05.109
  2. 2. Guo LJ, Jin H, Lu YJ. Supercritical water gasification research and development in China. Journal of Supercritical Fluids. 2015;96:144-150. DOI: 10.1016/j.supflu.2014.09.023
  3. 3. Zhao Z, Su S, Si N, Hu S, Wang Y, Xu J, et al. Exergy analysis of the turbine system in a 1000 MW double reheat ultra-supercritical power plant. Energy. 2017;119:540-548. DOI: 10.1016/j.energy.2016.12.072
  4. 4. Li ZH, Wu YX, Tang GL. Comparison between heat transfer to supercritical water in a smooth tube and in an internally ribbed tube. International Journal of Heat and Mass Transfer. 2015;84:529-541. DOI: 10.1016/j.ijheatmasstransfer.2015.01.047
  5. 5. Bae YY, Kim HY, Kang DJ. Forced and mixed convection heat transfer to supercritical CO2 vertically flowing in a uniformly-heated circular tube. Experimental Thermal and Fluid Science. 2010;34:1295-1308. DOI: 10.1016/j. expthermflusci. 2010. 06. 001
  6. 6. Kurganov VA, Kaptil'ny AG. Velocity and enthalpy fields and eddy diffusivities in a heated supercritical fluid flow. Experimental Thermal and Fluid Science. 1992;5:465-478. DOI: 10.1016/0894-1777(92)90033-2
  7. 7. Bae JH, Yoo JY, McEligot DM. Direct numerical simulation of heated CO2 flows at supercritical pressure in a vertical annulus at Re = 8900. Physics of Fluids. 2008;20:055108. DOI: 10.1063/1.2927488
  8. 8. Bae JH, Yoo JY, Choi H. Direct numerical simulation of turbulent supercritical flows with heat transfer. Physics of Fluids. 2005;20:105104. DOI: 10.1063/1.2047588
  9. 9. Mokry S, Pioro IL, Farah A, King K, Gupta S, Peiman W, et al. Development of supercritical water heat-transfer correlation for vertical bare tubes. Nuclear Engineering and Design. 2011;241:1126-1136. DOI: 10.1016/j.nucengdes.2010.06.012
  10. 10. Bae YY, Kim HY. Convective heat transfer to CO2 at a supercritical pressure flowing vertically upward in tubes and an annular channel. Experimental Thermal and Fluid Science. 2009;33:329-339. DOI: 10.1016/j.expthermflusci.2008.10.002
  11. 11. Li FB, Bai BF. A model of heat transfer coefficient for supercritical water considering the effect of heat transfer deterioration. International Journal of Heat and Mass Transfer. 2019;133:316-329. DOI: 10.1016/j.ijheatmasstransfer.2018.12.121
  12. 12. Jackson JD. Models of heat transfer to fluids at supercritical pressure with influences of buoyancy and acceleration. Applied Thermal Engineering. 2017;124:1481-1491. DOI: 10.1016/j.applthermaleng.2017.03.146
  13. 13. Yamagata K, Nishikawa K, Hasegawa S, Fujii T, Yoshida S. Forced convective heat transfer to supercritical water flowing in tubes. International Journal of Heat and Mass Transfer. 1972;15:2575-2593. DOI: 10.1016/0017-9310(72)90148-2
  14. 14. Vikhrev YV, Barulin YD, Kon’Kov AS. A study of heat transfer in vertical tubes at supercritical pressures. Thermal Engineering. 1967;14(9):116-119
  15. 15. Pioro IL, Khartabil HF, Duffey RB. Heat transfer to supercritical fluids flowing in channels-empirical correlations (survey). Nuclear Engineering and Design. 2004;230:69-91. DOI: 10.1016/j.nucengdes.2003.10.010
  16. 16. Gupta S, Farah A, King K, Mokry S, Pioro I. Developing new heat-transfer correlation for supercritical-water flow in vertical bare tubes. In: Proceedings of ICONE-18; 17–21 May 2010; Xi’an, China
  17. 17. Yoo JY. The turbulent flows of supercritical fluids with heat transfer. Annual Review of Fluid Mechanics. 2013;45:495-525. DOI: 10.1146/annurev-fluid-120710-101234
  18. 18. Komita H, Morooka S, Yoshida S, Mori H. Study on the heat transfer to the supercritical pressure fluid for supercritical water cooled power reactor development. In: Proceedings of the NURETH-10, 2005; Seoul, Korea
  19. 19. Deev VI, Kharitonov VS, Baisov AM, Churkin AN. Universal dependencies for the description of heat transfer regimes in turbulent flow of supercritical fluids in channels of various geometries. Journal of Supercritical Fluids. 2018;135:160-167. DOI: 10.1016/j.supflu.2018.01.019
  20. 20. Cheng X, Yang YH, Huang SF. A simplified method for heat transfer prediction of supercritical fluids in circular tubes. Annals of Nuclear Energy. 2009;36:1120-1128. DOI: 10.1016/j.anucene.2009.04.016
  21. 21. Yu J, Jia B, Wu D, Wang D. Optimization of heat transfer coefficient correlation at supercritical pressure using genetic algorithms. Heat and Mass Transfer. 2009;45:757-766. DOI: 10.1007/s00231-008-0475-4
  22. 22. Kuang B, Zhang Y, Cheng X. A wide-ranged heat transfer correlation of water at supercritical pressures in vertical upward tubes. In: Proceedings of the China-Canada Joint Workshop on SCWR, 2008; Shanghai
  23. 23. Jackson JD. Fluid flow and convective heat transfer to fluids at supercritical pressure. Nuclear Engineering and Design. 2013;264:24-40. DOI: 10.1016/j.nucengdes.2012.09.040
  24. 24. Schatte GA, Kohlhepp A, Wieland C, Spliethoff H. Development of a new empirical correlation for the prediction of the onset of the deterioration of heat transfer to supercritical water in vertical tubes. International Journal of Heat and Mass Transfer. 2016;102:133-141. DOI: 10.1016/j.ijheatmasstransfer.2016.06.007
  25. 25. Koshizuka S, Takano N, Oka Y. Numerical analysis of deterioration phenomena in heat transfer to supercritical water. Transactions of the JSME. 1995;38:3077-3084. DOI: 10.1016/0017-9310(95)00008-W
  26. 26. Cheng X, Kuang B, Yang YH. Numerical analysis of heat transfer in supercritical water cooled flow channels. Nuclear Engineering and Design. 2007;237:240-252. DOI: 10.1016/j.nucengdes.2006.06.011
  27. 27. Hu ZH. Heat transfer characteristics of vertical upflow and inclined tube in the supercritical pressure and near-critical pressure region [thesis]. Xi’an: Xi’an Jiaotong University; 2001
  28. 28. Peeters JW, Pecnik R, Rohde M, et al. Turbulence attenuation in simultaneously heated and cooled annular flows at supercritical pressure. Journal of Fluid Mechanics. 2016;799:505-540. DOI: 10.1017/jfm.2016.383
  29. 29. Zhang G, Zhang H, Gu H, Yang Y, Cheng X. Experimental and numerical investigation of turbulent convective heat transfer deterioration of supercritical water in vertical tube. Nuclear Engineering and Design. 2012;248:226-237. DOI: 10.1016/j.nucengdes.2012.03.026
  30. 30. Kim DE, Kim MH. Two layer heat transfer model for supercritical fluid flow in a vertical tube. Journal of Supercritical Fluids. 2011;58:15-25. DOI: 10.1016/j.supflu.2011.014
  31. 31. Negoescu CC, Li Y, Al-Duri B, Ding Y. Heat transfer behavior of supercritical nitrogen in the large specific heat region flowing in a vertical tube. Energy. 2017;134:1096-1106. DOI: 10.1016/j.energy.2017.04.047
  32. 32. Patankar SV, Ivanovic M, Sparrow EM. Analysis of turbulent flow and heat transfer in internally finned tubes and annuli. Journal of Heat Transfer-Transactions of the ASME. 1979;101:29-42. DOI: 10.1115/1.3450925
  33. 33. Chu X, Laurien E, McEligot DM. Direct numerical simulation of strongly heated air flow in a vertical pipe. International Journal of Heat and Mass Transfer. 2016;101:1163-1176. DOI: 10.1016/j.ijheatmasstransfer.2016.05.038
  34. 34. Huang PG, Coleman GN, Bradshaw P. Compressible turbulent channel flows: DNS results and modelling. Journal of Fluid Mechanics. 1995;305:185-218. DOI: 10.1017/S0022112095004599
  35. 35. Wen QL. Experimental study of flow and heat transfer characteristics at SCWR typical conditions [thesis]. Shanghai, Shanghai Jiao Tong University; 2015
  36. 36. Wang F, Yang J, Gu H, Zhao M, Li H, Lu D. Experimental research on heat transfer performance of supercritical water in vertical tube. Atomic Energy Science and Technology. 2013;47:933-939

Written By

Fangbo Li, Binbin Pei and Bofeng Bai

Submitted: 03 June 2019 Reviewed: 26 August 2019 Published: 03 October 2019