Dryout study of a helical coil once-through steam generator integrated in a thermal storage prototype

Dryout study of a helical coil once-through steam generator integrated in a thermal storage prototype

Journal Pre-proofs Dryout study of a helical coil once-through steam generator integrated in a thermal storage prototype E. Rivas, J. Muñoz-Antón PII:...

10MB Sizes 0 Downloads 57 Views

Journal Pre-proofs Dryout study of a helical coil once-through steam generator integrated in a thermal storage prototype E. Rivas, J. Muñoz-Antón PII: DOI: Reference:

S1359-4311(19)37115-7 https://doi.org/10.1016/j.applthermaleng.2020.115013 ATE 115013

To appear in:

Applied Thermal Engineering

Received Date: Revised Date: Accepted Date:

14 October 2019 9 January 2020 25 January 2020

Please cite this article as: E. Rivas, J. Muñoz-Antón, Dryout study of a helical coil once-through steam generator integrated in a thermal storage prototype, Applied Thermal Engineering (2020), doi: https://doi.org/10.1016/ j.applthermaleng.2020.115013

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version will undergo additional copyediting, typesetting and review before it is published in its final form, but we are providing this version to give early visibility of the article. Please note that, during the production process, errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

© 2020 Published by Elsevier Ltd.

DRYOUT STUDY OF A HELICAL COIL ONCE-THROUGH STEAM GENERATOR INTEGRATED IN A THERMAL STORAGE PROTOTYPE E. Rivasa,∗, J. Mu˜ noz-Ant´ona a Escuela

T´ ecnica Superior de Ingenieros Industriales Universidad Polit´ ecnica de Madrid (ETSII-UPM) C/ Jos´ e Guti´ errez Abascal, 2, 28006 Madrid, Spain

Abstract In this work, it is studied several dryout parameters (the dryout onset, first and total dryout, its associated azimuthal position and local steam qualities) in a triple helical coil once-through steam generator under quasi-steady working conditions (inlet pressure, PIN ∈ [37.9, 39.2] bar, inlet subcooling, ∆Tsub ≈ 0 ◦

C and total mass flow rate, m ˙ ws ≈ 0.082 kg·s-1 ) which exchanges heat with

a molten salts counterflow under quasi-steady pressure and temperature conditions (atmospheric pressure and inlet temperature, TIN ∈ [472, 477] ◦ C) but transient mass flow rate, m ˙ ms , depending on the helical coil diameter, D. For this purpose, a CFD model (Computational Fluid Dynamics model) has been used which has been evaluated with the experimental data obtained during a discharge test of the 300 kWth thermal storage prototype with integrated steam generator of the Casaccia Research Center (ENEA, Italy), previous benchmarking of two phase flow model. Results show, within the working conditions range and considering the particular steam generator characteristics, the two-phase flow behavior is uniform. This information has not been reported yet on available literature. In addition, numerical local steam qualities are compared with those calculated by using forced convective boiling correlations suitable for this ∗ Corresponding

author Email address: [email protected] (E. Rivas)

Preprint submitted to Applied Thermal Engineering

January 31, 2020

type of steam generator geometry, concluding that: 270◦ Ruffel’s and Santini’s correlations are the most appropriate during first dryout, and to assume a local steam quality value of 0.97 during total dryout the most suitable approximation. This work can be useful for those who pretend to scale-up such a steam generator for industrial applications requiring process heat and/or power generation. Keywords: scale-up, once-through steam generator, helical coil, dryout, steam quality, Computational Fluid Dynamics (CFD). elsarticle.cls, LATEX, Elsevier, template 2010 MSC: 00-01, 99-00

1

1. Introduction

2

Currently, there are a large number of industrial applications where process

3

steam is required (desalination, sterilization, building or district heating and

4

cooling, etc.). Other clear examples are the CSP plants (Concentrating Solar

5

Power plants), which generate electricity from the transformation of concen-

6

trating solar radiation. In them, the presence of thermal storage systems is

7

very common because it allows to manage its electrical production even in the

8

absence of solar radiation.

9

In general, the thermal storages for commercial CSP plants are based on the

10

double tank system with molten salts. Although the efficiency of these systems

11

is very high (> 93 %) [1], it is necessary to research and develop new thermal

12

storage designs more profitable in order to demonstrate that this technologies

13

have enough potential to replace the more conventional ones, such as those based

14

on the use of fossil fuels.

15

With this aim, the ENEA Research Center (Italy) [2] and the company

16

ANSALDO Nucleare S.p.A. (Italy) [3] have patented the layout of a molten

17

salts thermal storage system based on a single tank with an integrated helical

18

coil steam generator [4], inspired by the SFR technology (Sodium-cooled Fast

19

Reactors) [5]. The patent bases its competitiveness in the simplification of the

20

plant configuration [6]. Subsequently, its technical feasibility has been demon-

2

21

strated on a prototype scale (300 KWth) in ENEA [7, 8, 9]. The next step in

22

its development would be to scale-up it at commercial CSP plant level (> 10

23

MWe).

24

One of the key points in the thermohydraulic design of the system is the

25

calculus of local steam qualities associated to the onset of the dryout because it

26

implies a worsening of the local heat transfer coefficients in the water-steam side

27

(in some cases until one and two orders of magnitude [10]). The difficulty is due

28

to, from the experimental point of view, the ENEA prototype instrumentation

29

is not enough: there are no measurements of local heat fluxes and there are only

30

temperature measurements of a single coil. Also, from the theoretical point of

31

view, the found correlations in the literature for forced convective boiling flows

32

in this type of geometries are not 100 % suitable: either by the heat transfer

33

fluid, the helical coil/tube characteristics or the working conditions.

34

Therefore, the objectives of this work are:

35

• to study the dryout onset in the covered working conditions range depend-

36

ing on the helical coil diameter,

37

• and from this information, to assess the adaptability of the literature

38

correlations, both during the first and the total dryout, to this system.

39

For this purpose, the employed methodology has been the numerical analysis

40

by CFD modeling (software: STAR-CCM+ 9.04.011 [11]), which results has

41

been evaluated, first, against experimental data from the state-of-the-art, and

42

second, against those obtained during a discharge test of the thermal storage

43

prototype with integrated steam generator belonging to ENEA.

44

This work is structured as follows: first of all, it is summarized the state-

45

of-the-art correlations for calculating local steam qualities associated with the

46

dryout onset in forced convective boiling, particularly in vertical helical coils.

47

Then, the ENEA thermal storage prototype and its operating mode during

48

discharge tests is described. Secondly, the CFD pre and post processing are

49

presented, and the model evaluation (both numerical and experimental) is made

50

up. And finally, the dryout study as function of the helical coil diameter and 3

51

the comparison with the found correlations, are carried out. The most relevant

52

conclusions are set out.

53

2. State-of-the-art of the correlations for calculating local steam qual-

54

ities associated with the dryout onset in forced convective boiling

55

flows in vertical helical coils

56

One of the main differences in forced convective boiling flows between vertical

57

straight tubes and vertical helical coils is the dryout onset. While in the first

58

case it happens simultaneously in all tubes at equal height, in the second one,

59

it starts at certain surface points and spreads until all surface points dryout

60

completely at a particular height (Figure 1), i.e., there is a height range where

61

the coils are partially dry and partially wet.

Figure 1: Visualization of dryout onset in a helical coil (Source: CFD research findings).

62

This difference is due to the joint action of frictional, centrifugal and vis-

63

cous forces. These forces induce the two-phase flow undergoes a secondary flow

64

perpendicular to the principal flow [12]. The secondary flow promotes the extra-

65

dos surface wetting until higher steam qualities than in case of straight tubes.

66

Comparative studies carried out by Cumo et al. [13] and Styrikovich et al. [14]

67

between vertical straight tubes and vertical helical coils show such difference.

4

68

The characteristics of employed samples and the working conditions ranges of

69

these studies are shown in Table 1.

70

In Table 1 it can be seen that the main involved factors on found correlations

71

for calculating steam qualities for dryout onset in forced convective boiling flows

72

are pressure (P ), mass flux (G) and heat flux (Q).

73

The pressure and mass flux effects were analyzed by Styrikovich et al. [14].

74

This work analyses that increasing pressure and mass flux, steam quality de-

75

creases accordingly during the total dryout, evidencing its dependency. Among

76

the revised correlations, only Naitoh et al. correlation [15] is mass flux indepen-

77

dent, see Table 1.

78

Berthoud and Jayanti propose three different correlations to determine the

79

steam quality during the first dryout depending on the type of annular flow

80

regime [16] (Table 1). This work is based on the experimental data of Roumy

81

¨ [17], Styrikovich et al. [14], Unal et al. [18, 19], Breus and Belyakov [20] and

82

Carver et al. [21]. To identify the regime type, Berthoud and Jayanti employed

83

a 2D map (Figure 2) in which, on the x-axis, it is represented the dimensionless

84

number x0 given by the expression:

x0 =

G √ ρv gD

(1)

85

where ρv is the steam density, g the gravity and D the helical coil diame-

86

ter. This parameter is a measure of the centrifugal force acting on the vapor

87

phase and the drag force acting on the droplets, which shows up the droplets

88

redeposition process. And, on the y-axis, it is represented the dimensionless

89

number y0 (equivalent to the Reynolds number of the liquid phase) given by the

90

expression:

y0 = 91

G · d∗ µl

where d∗ is a geometric factor given by di ·

(2) q

di ( 0.02 ) (di is the inner tube

92

diameter) and µl the water dynamic viscosity. This parameter shows up the

93

droplets entrainment process through the vapor core. 5

Figure 2: Processes controlling the first dryout in helical coils, 2D map [16].

94

95

On this map, depending on working conditions and samples characteristics, three different zones can be distinguished:

96

• Gravity zone: characterized by high D, low G and high P , where dryout

97

onset happens at low steam qualities. In this case, flows are often very

98

stratified.

99

• Redeposition zone: characterized by low D, high G and low P , where

100

dryout onset happens at high steam qualities. In this case, flows tend to

101

be slightly stratified and the extrados surface remains often well chilled.

102

• Entrainment zone: characterized by very high D and G and standard P ,

103

where dryout onset happens at medium steam qualities. In this case, flows

104

tend to be slightly stratified and the intrados surface remains often well

105

chilled.

106

In this work, the authors also propose a correlation during total dryout, 6

107

noting that it also depends only on the type of annular flow regime [16] (Table 1).

108

Although, in contrast to the previous correlations, it is heat flux independent.

109

110

Therefore, to clarify this issue (heat flux dependence/independence) it is necessary to draw upon other works.

111

¨ While in correlations proposed by Duchatelle et al. [22], Unal et al. [18, 19],

112

Jensen et al. [23, 24, 25, 26] and Ruffell [27] heat flux appears explicitly, in that

113

proposed by Santini et al. [28] heat flux does not appear (see Table 1). First

114

three correlations were obtained from non-uniform wall heatings and fourth and

115

fifth ones were obtained at lab scale with electrical heat (uniform heating):

116

117

118

119

120

121

• 1st : Duchatelle et al. used the data of a real steam generator exchanging heat with a liquid sodium at counterflow ¨ • 2nd : Unal et al. used the data of three concentric helical coils heated by means of an external liquid sodium at counterflow • 3rd : Jensen et al. used the data of three helical coils electrically heated applying a higher flux on the top surface than on the bottom surface.

122

I.e., the employed wall heating methodologies (uniform or non-uniform) are

123

different and, therefore, it is difficult to distinguish the potential heat flux de-

124

pendence/independence.

125

On the other hand, the Jensen’s correlation is not applicable here because its

126

working pressure is far from those of the ENEA prototype (since he employed

127

¨ R113 as working fluid). And, neither the Duchatelle’s and Unal’s correlations

128

are applicable here because they have been derived from very different working

129

conditions ranges. In particular, the first, for large heat and mass fluxes and

130

the second, for very high pressures and inlet subcoolings.

131

Even though in most of the works where it was employed few helical coils par-

132

allel arranged, two-phase flow instabilities were observed, hardly ever included

133

specific studies about their influence on the dryout (in many cases, two-phase

134

flow instabilities were inhibited by using calibrated orifices at the tubes in-

135

¨ lets). About this regard, highlighting the Chapter 7 of Unal [18], where a study

7

136

about the DWOs (Density Wave Oscillations) observed during the operation of

137

a full-scale steam generator is presented, and the Naitoh et al. study [26], who

138

investigated the dryout coupled by DWOs. Naitoh deduced the dryout fluctua-

139

tions had periods between 3 s and 20 s, result consistent with the experimental

140

data of the ENEA prototype. Here it has shown, without inhibiting two-phase

141

flow instabilities experimentally, they do not influence the dryout by using a

142

simulation methodology which takes into account this period.

143

Finally, there are very few authors who have analyzed the effect of the geo-

144

metric characteristics of the samples. According to Ruffell [27], the helix angle,

145

¨ ϕ, strongly influences the first dryout and according to Unal et al. [18, 19],

146

if D/di ≥ 38.9, the geometry effect on the necessary power (Q˙ nec ) to achieve

147

the first dryout is negligible, but not the necessary power to achieve the total

148

dryout. Here it has analyzed both of them: ϕ ∈ [1.61, 2.36]◦ and D/di ∈ [15,

149

22] under different working conditions.

150

Considering this review, the objectives of this study are justified based on

151

the following ideas: it is difficult to distinguish which correlation/s among those

152

found in the literature are applicable to the ENEA prototype working conditions

153

for calculating local steam qualities associated with the dryout onset since not all

154

of them depend on the same factors (Naitoh’s, Berthoud and Jayanti’s, Ruffell’s,

155

and Santini’s. correlations, see expressions and applicability ranges in Table 1).

156

Perhaps, due to the fact that the employed wall heating methodologies are very

157

different. Limitation which, besides, was also detected by Santini et al. [28],

158

¨ Jensen et al. [23, 24, 25, 26], Unal et al. [18, 19] and Ruffell [27] in their works.

8

9

L = 1.709 m; di = 0.010 m; D = 0.136 m; pitch = 0.055 m

(upward or downward)

1984) [14]

Water/R12 (upward)

Water (upward)

Jayanti 1990) [16]

(Santini et al.

2014) [28]

do = 0.00635 m;

(upward)

(Berthoud &

L = 2 m; di = 0.00475 m;

1972) [13]

G ∈ [199, 810] kg·s-1 ·m-2

D = 1 m; pitch = 0.79 m

do = 0.01723 m;

P ∈ [10.7, 60.7] bar

Q ∈ [43.6, 209.3] kW·m-2

1 Helical coil L = 32m; di = 0.01253 m;

P ∈ [11, 200] bar

Q ∈ [10, 1800] kW·m

-2

G ∈ [100, 1900] kg·s-1 ·m-2

P ∈ [65, 240] bar

Q = 120 kW·m-2

G ∈ [500, 1800] kg·s-1 ·m-2

P ∈ [98.1, 176] bar

Q ∈ [100, 1500] kW·m-2

G ∈ [500, 1500] kg·s-1 ·m-2

xT ot

x1 = 103.235 ·

ρl ρv

v

v

Gravity zone −2.378  −1.712  · Gµldi · ρ

v

G √ g D

0.967  −0.740 · GQ ∆H

Redeposition zone 0.098 ! 0.067  −0.430  0.101  −0.785  q G di G · µl · ρ √ · GQ · µl Q∆H · g·(ρlσ−ρv ) ∆H g D

ρv ρl

x1 = 0.44 − 0.0006 · P · 10−5 · G0.114



v

G √ g D

0.08 

Entrainment zone 0.119 −0.267  −0.984  0.950  −0.428  q G · Gµldi · ρ √ · GQ · µl Q∆H · g·(ρlσ−ρv ) ∆H g D

ρl ρv







CORRELATION

Total dryout   1.722  −0.494  2 −0.381  −1.61  ρl = log10 · µµvl · Gµldi · Gσ µdli · ρ ρv



x1 = 3.223 + log10



x1 = 107.068 ·

DRYOUT STEAM QUALITY CONDITIONS

D = [0.08-3.3] m

di = [0.008-0.02] m;

1 Helical coil

D = 0.09 m; pitch = 0.07 m

Straight tube vs. helical coil R12

(Cumo et al.

Straight tube vs. helical coil

Water

(Styrikovich et al.

GEOMETRY

FLUID

AUTHOR

vertical helical coils.

Table 1: Correlations for calculating local steam qualities associated with the dryout onset (first and total) in forced convective boiling flows in

10 Water (upward)

¨ ¨ Unal 1981, Unal

et al. 1981) [18, 19]

R113 (upward)

Jensen & Bergles 1981,

FLUID

1982, 1983) [23, 24, 25, 26]

(Jensen 1980,

AUTHOR

GEOMETRY

xT ot ∈ [0.55, 0.94]

0.1175 m; di /D = 0.0182,

ϕ = 7.77◦ , 8.83◦

pitch = 0.79 m;

D= 1.5 m, 0.7 m, 0.7 m;

xT ot ∈ [0.08, 1.00]

∆Tsub ∈ [35.6, 156.8] ◦ C

P ∈ [147, 202] bar

Q ∈ [41,731] kW·m-2

L = 40.13 m, 35.50 m, 26.67 m; di = 0.018 m;

G ∈ [112, 1829] kg·s-1 ·m-2

4 Helical coils

0.01270 m

pitch = 0.0254 m, 0.01588 m,

ϕ = 1.78◦ , 1.68◦ , 3.87◦ ;

0.0353, 0.0649;

∆Tsub ∈ [0,110] ◦ C

P = 9.4 bar

Q ∈ [54, 800] kW·m-2

G ∈ [570,5470] kg·s-1 ·m-2

CONDITIONS

D = 0.4096 m, 0.2159 m,

do =0.00777 m, 0.00769 m;

di =0.00762 m, 0.00744 m;

heating length);

1.295 m, 1.257 m (wall

L = 0.635 m, 1.27 m,

3 Helical coils





, if G > 950 kg· m-2 · s-1

, if G ≤ 950 kg· m-2 · s-1  di 0.31 D

 di 0.17 D

    PIN IN a1 = 1 + 3.8 · 1 − HHSAT ; a2 = 0.114 − 0.041 · log 1 − P cri  −1/2 L a3 = 1 + 4.59 · deq ; Leq = πAdGQ · (HSAT − HIN + ∆H · x) i n o n o a41 = 1 + 0.44 · exp −0.056 · dDi − exp − 3dD , 1st dryout i  n o n o a42 = 1 + 0.56 · exp −0.011 · dDi − exp − 3dD , Total dryout i  0.32 Leq 2 (d0 −di ) 0.22 a5 = , a6 = di + 28 · F r , a7 = 1 di

   3 ·a41 ·a5 Q = G · ∆H · 0.97 · a1 ·a2a·a6 ·a , 1st dryout 7·    a1 ·a2 ·a3 ·a42 ·a5 Q = G · ∆H · 0.97 · , Total dryout a6 ·a7 ·

 Q = G · ∆H · 1.7126 · 105 · Re−1.143 · x−0.436 · T ot

 Q = G · ∆H · 4.09 · 10−6 · Re0.50 · x−0.460 · T ot

CORRELATION

11

Water (upward)

Water (upward)

Water (upward)

(Naitoh et al.

1974) [15]

(Duchatelle et al.

1975) [22]

(Ruffel et al.

1974) [27]

G ∈ [300, 1800] kg·s-1 ·m-2

3 Helical coils

D = 0.58 m, 1.32 m, 2.36 m

length); di = 0.0125 m;

P up to 180 bar*

Q/G ∈ [0.06, 0.34] J·G-1

P ∈ [45, 175] bar

0.011, 0.007

L = 8.2 m (wall heating

Q ∈ [310, 1500] kW·m-2

G ∈ [375, 3500] kg·s-1 ·m-2

P up to 175 bar

CONDITIONS

di /D = 0.032, 0.025,

4 Helical coils

di /D=0.0165

GEOMETRY

*Although it was also validated for other working conditions by means of modeling.

FLUID

AUTHOR

0.13·(D/(100·di )) (G/103 )1.5

·



Q 100

x1,270◦ = 1 − 0.0004 · Q − 0.0109 ·

x1,0◦ = 1 −

 G 2 103

·

D di

 0.5

0.75·(G/103 )0.5

xT ot = 1.39 · 10−4 · Q0.732 · G−0.209 · exp 2.46 · 10−3 · P · 10−5

xT ot = 1 + (0.139 − 0.071 · (P · 10−5 )0.186 ) · (Q · 0.85986 · 10−5 )

CORRELATION

159

3. ENEA experimental setup

160

ENEA prototype is a 2 m diameter and 2.8 m height thermal storage tank

161

whereby the hot molten salts enter through the upper side of the tank. Inside it,

162

there are 12 tons mixture of inorganic molten salts, commonly known as “solar

163

salt”: NaNO3 and KNO3 (60/40 % by weight), under atmospheric pressure. The

164

integrated steam generator is vertically and non-symmetrically arranged within

165

the tank. It consists of two concentric cylindrical shells, called “downcomer”

166

and “shell”, and three in-line helical coils, referred as #1, #2 and #3 (coil/tube

167

characteristics: see Table 2) comprised in the space between shells (Figure 3).

168

For additional information about the prototype setup consult [29] and [30].

Figure 3: ENEA experimental setup

169

During discharge tests, the pressurized water flowing within the tubes changes

12

170

phase due to the heat transfer with the molten salts at counterflow circulation

171

through the space between shells and tubes. Therefore, the highest thermal

172

hydraulic gradients within the prototype happen here and thus, the involved

173

phenomenology intrinsically depends on the steam generator geometry. Table 2: Design characteristics of the integrated steam generator (Source: ENEA).

GEOMETRIC FEATURE

VALUE (Unit)

External tube diameter, do

0.0127 m

Internal tube diameter, di

0.0094 m

Horizontal pich, Xhor

0.0162 m

Vertical pich, Xver

0.0180 m

Helical coil diameters, D1

0.2038 m

Helical coil diameters: D2

0.1714 m

Helical coil diameters: D3

0.1390 m

174

The shells have a thermal insulating function: the downcomer allows the

175

water to arrive the lower part of the generator without directly exchanging heat

176

with the molten salts (thus, water-steam phase change happens upwards) and

177

the shell separates the molten salts circulating though the generator from the

178

molten salts tank bulk. Finally, it has a window at the top, through which the

179

hot molten salts go into the generator, and a diffuser at the bottom, through

180

which the cold molten salts depart the generator.

181

Its operating mode during discharge tests is as follows: the water flow is

182

established by a pump belonging to an external loop; that is the reason to talk

183

about forced convective boiling in the water-steam side. With this operating

184

mode, the coming out fluid is superheated steam, saturated steam or subcooled

185

water, depending on the thermal conditions of the tank. Molten salts circulate

186

through the existing gaps in the tubes matrix by gravity. Their driving force is

187

gravity when go through the generator, since they increase their density while

188

they cool down because heat exchange with the tubes matrix where the pres-

189

sured water flows at a lower temperature than molten salts. This process causes 13

190

the molten salts tank stratification, i.e., it promotes the development of a cold

191

zone at the tank lower part, coming from the salts that have gone through the

192

generator, a hot zone at the tank upper part, and an intermediate zone, which

193

thickness depends on the tank thermal conditions (see Figure 4). In summary,

194

during discharge tests, there is a continuous natural circulation of molten salts

195

which transports heat from the source (molten salts) to the sink (water-steam).

196

Such molten salts natural circulation will slow down as the discharge evolves

197

due to the gradual decrease of the buoyancy forces. This is the reason to talk

198

about molten salts transient mass flow rate. This gradual decrease is the result

199

of the progressive difference height decreasing between the respective hot molten

200

salts columns from the source and the sink, ∆L, during discharge, i.e, in Figure

201

4, ∆L7050s > ∆L10650s .

Figure 4: Simplified temperature field at two times, t7050s (left) and t10650s (right) during a discharge test [29]. Arrows indicate the molten salts inlet/outlet.

202

The heat transfer between molten salts and water-steam is less efficient as

203

discharge evolves. Consequently, as the discharge evolves, the two-phase flow

204

length on the steam generator increases (in Figure 4, z7050s < z10650s although

205

this two-phase flow length may be different on each helical coil). This fact can

206

affect the dryout onset depending on the helical coil diameter. In order to check

207

this influence, a CFD model has been developed and validated with the different 14

208

variables measured in the prototype at ENEA. This analysis has not yet been

209

reported on the available bibliography and could be of interest for developers of

210

this type of steam generator technology.

211

4. CFD modelling

212

4.1. Geometry and mesh models

213

CFD model is focused on the steam generator. Geometrically is a 3D full-

214

R scale model made by means Autodesk Inventor Professional 2016 (helical coils

215

R design) [31] and STAR-CCM+9.04.011 CAD module (downcomer, shell, win-

216

dow and diffuser design) [11], see Figure 5.

Figure 5: Steam generator geometric model, perspective views: outer view: shell, window and diffuser (top); intermediate view: helical coils (center) and inner view: downcomer (bottom).

15

217

Downcomer, shell, window and diffuser have been assumed un-thickness and,

218

in addition, adiabatic elements. Although window and diffuser are metal ele-

219

ments (with high thermal conductivity), their influence on the water-steam be-

220

haviour in the helical tubes has been considered negligible because they are far

221

enough.

222

The geometric domain is divided into three subdomains: the water-steam

223

side, the molten salts side, and the helical coils walls, which separate both media.

224

Therefore, the model allows to study the thermohydraulic behavior of molten

225

salts circulating by the space existing between helical coils and shells and water-

226

steam flowing within the tubes simultaneously.

227

They have been discretized by means of a polyhedral cells meshing (see

228

Figure 6). The maximum characteristic sizes in molten salts and water-steam

229

sides are 0.0050 m and 0.0025 m respectively, although meshing includes some

230

refinements in the narrower zones with, at least, two polyhedral cells by gap.

231

Both subdomains take account with a prismatic layer, single in the first

232

case and double in the second, on their respective boundaries (i.e. shell and

233

tubes, in molten salts side; and tubes, in water-steam side), which maximum

234

characteristic sizes are 0.00250 m and 0.00125 m respectively (see Figure 6).

235

The transition between prismatic layers (smaller cells) and polyhedral bulks

236

(larger cells) has been carried out gradually (by a 1.5 factor). Also, a double

237

uniform prismatic layer has been included filling the helical coils walls thickness

238

(see Figure 6).

239

The mesh has been made using the automatic tool of STAR-CCM+ 9.04.-

240

R 011 . The total number of cells is around 7.2·106 ; from which 32 % belong to

241

molten salts side, 45 % belong to water-steam side and 23 % belong to helical

242

coils walls.

16

Figure 6: Mesh model detailed view: molten salts side in yellow, helical coils walls in grey and water-steam side in blue.

243

In order to ensure numerical results are mesh model independent, a test

244

varying the maximum characteristic size only in water-steam side has been car-

245

ried out (the molten salts mesh model was already evaluated by Rivas and Rojas

246

[29]). The results are summarized in the following table (Table 3) and they are

247

satisfactory for “coarse” and “medium” mesh models. Table 3: Mesh model test: some numerical results at t¯7050s .

Total number of cells Characteristic size (m)

COARSE

MEDIUM

FINE

Reference Value

≈ 6.9·106 (97 %)

≈ 7.2·106 (100 %)

≈ 8.5·106 (119 %)



0.00375

0.00250

0.00150



PIN,1 (bar)

39.0

39.0

38.5

39.2

PIN,2 (bar)

39.0

39.0

38.7

39.2

PIN,3 (bar)

39.1

39.1

39.0

39.2

zx1 ,3 Y+ water-steam side

1.16

1.16

1.13



50-500

50-300

10-60

Recommended: < 1 or > 30

248

As can be seen, the finest mesh (referred in Table 3 as “fine”) presents

17

249

the worst numerical results of some variables of interest to this work as the

250

inlet pressure to each helical coil, {PIN,# }#=1,2,3 . It is due to the used wall

251

treatment “All Y+”. For this mesh model, the wall treatment dependency

252

with the prismatic layer size on boundaries implies unsuitable values of the Y+

253

in water-steam side (values between 1 and 30, [32, 33]) to assess adequately

254

the turbulent variables (shear stress, turbulence kinetic energy and turbulent

255

dissipation rate), on such boundaries. Since, to make a finer mesh to solve

256

this problem and thus, to obtain suitable values of the Y+ (values less than

257

1, [32, 33]), would be a very time consuming task from the computational cost

258

point of view and coarser meshes present equally suitable inlet pressure results

259

(absolute errors between 0.1 bar and 0.2 bar), they have been selected. Besides,

260

with these coarser meshes (“medium” and “coarse”), similar values of the first

261

dryout z-location in helical coil 3, zx1 ,3 , have been obtained. Therefore, the finer

262

mesh of both (“medium”) has been chosen because its computational cost did

263

not differ too much from the “coarse” mesh (3 %). Colombo [34] drew similar

264

conclusions regarding finer meshes when this type of wall treatment is used.

265

266

In addition, the available experimental data allow to evaluate the chosen mesh model making needless a more exhaustive test.

267

Finally, a mesh topological diagnosis has been conducted using the statistical

268

parameters (volume change, cell quality, face validity y cell and boundary skew-

269

ness angle) established by the software developer [33]. The results are shown in

270

the following table (Table 4). Table 4: Topological diagnosis of “medium” mesh.

MINIMUM VALUE

MAXIMUM VALUE

Cell and boundary skewness angles

0◦

89◦

Face validity

1

1

Cell quality

0.01

1

0.0004

1

Volume change

271

All obtained values are within the ranges which ensure the good quality of

18

272

the mesh, according to the information provided by the software developer [33].

273

All the above following the standards established in Best Practice Guidelines

274

for a suitable use of CFD codes in the industrial field, [35].

275

4.2. Physical models

276

To study the thermohydraulic behavior of the system, a conjugated steady-

277

state heat transfer model has been employed. This model takes into account

278

natural circulation in molten salts, conduction through helical coils walls and

279

forced convective boiling in water-steam.

280

Both external molten salts and internal water-steam flows are considered

281

turbulent flows. Although salts circulate by natural convection through the

282

generator and, therefore, they move slowly, the generator geometry strongly

283

disturbs their motion, generating complex circulations around helical coils which

284

do not allow simulating a laminar flow. However, water-steam flows by forced

285

convective boiling within the tubes, so, during the phase change turbulent is

286

promoted and upstream the flow is fully developed.

287

Both of them are simulated using a “RANS model” (Reynolds-Average

288

NavierStokes) whose closure is made using the “Realizable κ −  Two Layer

289

model” [32, 33], considering the approximations of Xu for molten salts (suitable

290

for buoyancy driven flows, [36, 33]) and Wolfstein for water-steam (suitable for

291

shear driven flows, [37, 33]), together with the hybrid wall treatment “All Y+”.

292

For the turbulent model, constants pre-established by the software developer

293

have been assumed. The flow is solved in a segregated way, and the nexus

294

between mass and momentum equations is made via the “SIMPLE algorithm”

295

with the Rhie-Chow interpolation [32, 33]. To solve the energy equation a segre-

296

gated model has been assumed (for the variables: temperature, in molten salts

297

and enthalpy, in water-steam).

298

Boussinesq approximation has been assumed in molten salts side [29]. The

299

rest of molten salts thermophysical properties have been considered temperature-

300

dependent [38].

301

The helical coils walls properties were considered constant and uniform, and 19

302

the pre-established values by the software developer for a stainless-steel type

303

UNSS316 [33] have been assumed.

304

4.2.1. Two-phase flow model description

305

To simulate the two-phase flow, a homogeneous model is employed. It solves

306

a single set of transport equations assuming the two-phase flow as a single-

307

phase flow with equivalent or mixture thermophysical properties. These ones

308

are calculated from the thermophysical properties of each phase and their corre-

309

sponding volume fractions, based on a homogeneity hypothesis. This two-phase

310

flow treatment is commonly used in industrial applications such as steam gener-

311

ators [33]. Its main advantage versus other models, which simulate each phase

312

separately, is its computational efficiency. The principal assumptions of this

313

model are: a) both phases flow at the same velocity (the slip between phases is

314

negligible) and b) both phases are in thermal equilibrium.

315

316

Two phase flow mass, momentum and energy equations, in integral form, are: ∂ ∂t

Z

Z

Z ρm vm · dS =

ρm dV + V

S

Sm dV

(3)

V

317

∂ ∂t

Z

Z ρm vm dV +

V

S

Z ρm vm ⊗ vm · dS = − P I · dS S Z Z + Tm · dS + (fg + sm ) dV S

(4)

V

318

∂ ∂t

Z

Z

Z (ρm Hm vm + P ) · dS = − Q · dS S ZS Z + Tm · vm · dS + (fg · vm + Sh ) dV

ρm Em dV + V

S

(5)

V

319

where mixture density and mixture velocity, ρm and vm , are defined as

320

function of densities and velocities of liquid phase and vapor phase respectively

321

(ρl , ρv , vl , vv ) as:

ρm = (α ρl + (1 − α) ρl ); vm = 20

(α ρl vl + (1 − α) ρv vv ) ρm

(6)

322

323

being: Sm , sm and Sh mass, momentum and energy sources, fg the gravity force, I the unitary tensor and Tm viscous stress tensor, calculated as:   2 T (∇vm + ∇vm ) − (∇ · vm ) I 3

Tm = µef f 324

where µef f is the mixture effective viscosity, given by:

µef f = µt + µm 325

326

being µt the turbulent viscosity, computed as ρm 0.09

(8) κ2 

and µm the mixture

viscosity, which is defined as:

µm = (α µl + (1 − α) µv ) 327

329

P |vm |2 ; Hm = hm − ρm 2

phase and vapor phase enthalpies respectively (hl , hv ): 1 (α ρl hl + (1 − α) ρv hv ) ρm

332

334

(12)

where kef f,l and kef f,v are the effective thermal conductivities of the liquid phase and vapor phase respectively, given by:

kef f = k − 333

(11)

And the heat flux vector Q is calculated as:

Q = − [α kef f,l + (1 − α)kef f,v ] · ∇T 331

(10)

where hm is the mixture static enthalpy determined as function of liquid

hm = 330

(9)

The mixture energy Em and the mixture enthalpy Hm are defined as:

Em = Hm − 328

(7)

µt CP P rt

(13)

being, k the thermal conductivity, CP the specific heat and P rt the turbulent Prandtl number of such phases.

335

The distribution of phases, that is, the volume fraction, α, is not obtained

336

by solving a transport equation, but from the distribution of static enthalpy 21

337

under the assumption of thermodynamic equilibrium. In this case, the steam

338

quality, x, is given by the next function: 



hm − hl,sat x = max 0, min 1, hv,sat − hl,sat 339

340

341

342

phase, respectively, at the saturation temperature Tsat . And finally, considering the above approaches, the steam volume fraction α can be defined as: x x + (1 − x)

344

(14)

where hl,sat and hv,sat are the static enthalpies of liquid phase and vapor

α=

343



ρv,sat ρl,sat

(15)

where ρl,sat and ρv,sat are the densities of liquid phase and vapor phase, respectively, at the saturation temperature Tsat .

345

The Rohsenow model [39, 33] is used to simulate the heat transfer through

346

the helical coils walls. This model is based on the following hypothesis: during

347

the phase change heat transfer through the walls takes place, either by nucleated

348

boiling or by film boiling.

349

Nucleated boiling occurs when walls temperature is slightly above the satu-

350

rated liquid temperature and involves the bubble nucleation at walls.

351

Here it is assumed nucleated boiling is present even within the annular liquid film

352

regime (this assumption is appropriate at relative low flow rates and sufficient

353

wall superheats, [40]). The bubbles nucleation depends on the walls roughness.

354

In this case, smoothed walls have been assumed. A wall finish type 2B (EN

355

10088-2, [41]) has been considered and, in the working conditions range (3 ·

356

104 < Re < 2 · 105 ), we have checked the flow behavior is hydraulically smooth

357

(although this check is outside of the scope of this work).

358

However, film boiling involves the creation of a continuous vapor film cover-

359

ing the helical coils walls when the critical heat flux is exceeded. Vapor film has

360

a low thermal conductivity, which isolates the walls decreasing the heat transfer

361

process considerably.

22

362

363

To calculate the heat flux during nucleated boiling, the model employs the Rohsenow correlation [39, 33]: r Q = µl hlat

g(ρl − ρv ) σ



CP,l (Tw − Tsat ) n Cqw hlat P rl p

3.03 (16)

364

In equation: µl , hlat , CP,l , ρl , y P rl are the viscosity, boiling enthalpy,

365

specific heat, density and Prandtl number of liquid phase respectively, np is a

366

model coefficient which depends on the working fluid (in this case, it is assumed

367

1, [42]), g is the gravity, ρv the vapor phase density, σ the surface tension on

368

water-steam interface, Tw the walls temperature, Tsat the saturation tempera-

369

ture and Cqw is a model coefficient which depends on the combination working

370

fluid-helical coils walls (it is assumed 0.013, [42]).

371

This correlation is applicable to any type of geometry since it does not

372

depend on it (neither on its orientation). It is based on the hypothesis the

373

agitation of the bubbles is the heat transfer mechanism controlling the nucleated

374

boiling, and it is formulated as a single-phase flow in forced convection.

375

To calculate the heat flux during film boiling, the model uses the same

376

equations as in the case of a single-phase flow of superheated steam. The model

377

coefficient which controls the transition between both regions is αf ilmBoiling .

378

If it is not possible to have a fine enough mesh to solve the steam film thick-

379

ness (as happens in this case), αf ilmBoiling should be explicitly given. Here,

380

a αf ilmBoiling = 1 has been assumed because, by means a sensibility analy-

381

sis (comparing the obtained results for αf ilmBoiling = 0.8, 0.95, 0.99, 1), we

382

have checked this value provided the best numerical results of some variables of

383

interest to this work.

384

Finally, for calculating water-steam thermophysical properties, the IAPWS-

385

IF97 models [43] have been used. These ones provide the fundamental equation

386

for the specific Gibbs free energy from which, using appropriate combinations

387

and their derivatives, it is possible to deduce the specific volume, internal energy,

388

entropy, enthalpy, specific heat and the sound velocity of water-steam.

23

389

4.2.2. Two-phase flow model justification

390

Currently, the two -phase flow mechanistic approaches (those which first

391

determine the existing two-phase flow pattern for certain given conditions and

392

then formulate separate hydrodynamic models for each of them) are the most

393

commonly used approaches according to the CFD codes state-of-the-art, but

394

some limitations and shortcomings have been found for its applicability to this

395

work as:

396

• they need a minimal number of coefficients which require calibration against

397

experimental data [44]. For example, in StarCCM+ it is possible imple-

398

ment different models, but all of them depend on specific submodels to

399

correctly capture: the nucleation site density, bubble departure frequency,

400

bubble departure diameter, quenching heat transfer coefficient, etc. [33].

401

Furthermore, solving this set of equations entails assuming higher compu-

402

tational costs and,

403

• they have an intrinsic dependence on the near-wall mesh [45, 46]. This

404

is due to these submodels were originally formulated for one-dimensional

405

thermo-hydraulic models in terms of the mean-flow variables. The imple-

406

mentation of these submodels as CFD wall boundary conditions assumes

407

replacing the non-local flow quantities by the near-wall local flow quanti-

408

ties. And, it can only be done for extremely coarse meshes, with the first

409

near-wall cell covering the whole boundary layer thickness. With a near-

410

wall mesh, which is adequate for the CFD turbulent wall treatment, such

411

approximation can significantly overestimate the vapor generation [47].

412

For these reasons and because:

413

• from the experimental point of view, the Roshenow correlation has been

414

tested for a wide range of working fluids on cylindrical surfaces and ma-

415

terials [48] and,

416

• from the numerical point of view, the homogeneous boiling model together

417

with the Roshenow correlation implemented in StarCCM+, have been 24

418

already successfully used in other cases to predict the forced convection

419

boiling flows behavior, for example [49], [50], and [51],

420

in this work, this semi-empiric approach has been chosen to simulate the two

421

-phase flow.

422

4.2.3. Two-phase flow model benchmarking

423

Since the experimental setup does not have other direct measurements in

424

water-steam side than inlet water and and outlet steam temperatures to each

425

helical coil ({TI 8 20#}#=1,2,3 and {TI 8 21#}#=1,2,3 respectively, Figure 3),

426

and no adequate experiments have been found in the literature about vertical

427

helical tubes to assess the two-phase flow model performance along the coils,

428

reference experiments on vertical straight tubes have been used here, as some

429

authors have already done in their works about vertical helical tubes (for exam-

430

ple [52]).

431

The experiments are those carried out by Bartolomei et al. [53], explained

432

and analyzed in [54], and Bartolomei and Chanturiya [55]. These tests, classified

433

in the literature as “surface boiling experiments”, are commonly used for boiling

434

models benchmarking [56].

435

Bartolomei et al. measured the average void fractions at different heights

436

along vertical straight tubes, in which the water-steam phase change happened

437

upwards. The test sections walls were uniformly heated and all the experiments

438

were carried out with subcooled water at the tube inlets.

439

Here, two test sections have been considered: the first one with tube diameter

440

of 12.03 mm and total tube length of 1.4 m (heated length of 1 m) and the second

441

one with diameter of 24 mm and total tube length of 2.4 m (heated length of 2

442

m). And two experiments in each of them.

443

The working conditions of these experiments are summarized in Table 5.

25

Table 5: Bartolomei et al. experiments working conditions.

Ref.

Experiment

Q (MW·m-2 )

m ˙ ws (kg·s-1 )

∆Tsub (K)

POU T (bar)

1

1.2

0.1705

63

68.9

2

0.8

0.1705

39

68.9

3

7.9

0.4026

48

45

4

7.9

0.4026

50

30

[53, 54]

[56]

444

As can be seen, they combine different values of boundary heat flux (Q),

445

mass flow rate (m ˙ ws ), inlet subcooling (∆Tsub ) and the outlet pressure (POU T ).

446

Using the mesh model described in section 4.1, a sensitivity study has been

447

carried out once again in order to ensure numerical results are mesh model

448

independent.

449

First, a benchmarking between the two-phase flow model and a mechanis-

450

tic type model implemented in STAR-CCM+, [56], is carries out. Results of

451

experiments 1 and 2 are employed for it.

452

453

The comparison between observed and simulated average void fractions along the first test section are shown in Figure 7.

Figure 7: Average void fraction as function of distance along the first test section.

454

The benchmarking is calculated by means of the statistical parameters: cor-

455

relation coefficient (R), normalized mean square error (N M SE), fractional bias

456

(F B) and fraction of prediction within a factor of two of observations (F AC2) 26

457

(see Appendix for definitions). A perfect model would have R, F AC2 = 1 and

458

N M SE, F B = 0.

459

460

In Table 6, these parameters for the average void fractions along test sections are shown. Table 6: Statistical parameters comparing the average void fractions models results with the measured data by Bartolomei et al. All of them are dimensionless.

EXPERIMENT 1

EXPERIMENT 2

NMSE

FB

FAC2

R

NMSE

FB

FAC2

R

SIMULATION

0.063

-0.041

0.750

0.931

0.149

-0.217

0.917

0.952

MECH.MODEL

0.075

0.027

0.750

0.960

0.092

0.122

0.917

0,974

461

In general, there is a good agreement for both cases. The N M SE and F B

462

have the same order of magnitude (note that the negative F B values indicate

463

a general model overestimation), the F AC2 are similar and the R are high for

464

both cases. Although the mechanistic model shows slightly better results.

465

466

Hence, it could be assumed that the approximations made on the watersteam side are a good compromise for this case.

467

Second, a comparison between the Bartolomei et al. experimental results,

468

which working pressures are closer to those of the ENEA prototype, and the nu-

469

merical results obtained with the two-phase flow model is is addressed. Results

470

of experiments 3 and 4 are used for it.

471

472

In Figure 8, observed and simulated average void fractions along the second test section are shown.

27

Figure 8: Average void fraction as function of distance along the second test section.

473

As can be appreciated, the two-phase flow model:

474

• has reasonably captured the boiling beginning in both cases,

475

• has correctly reproduced the average void fraction profiles: during experi-

476

ment 3 with an average relative error (RE) of 33 % and during experiment

477

4 with an average RE of 22 %, comparable to those showed by some mech-

478

anistic models for the average void fraction calculus, for example [52], [56]

479

and [57].

480

Finally, by means CFD simulations the dryout z-location on test sections

481

can be directly determined from steam quality azimuthal distributions on their

482

walls. Here, the simulated dryout z-location along the second test section has

483

been observed, during experiment 3, on zx = 1.8 m and, during experiment 4,

484

on zx = 2 m.

485

Therefore, considering all the above and given the objectives of this work, it

486

is concluded that this model seems suitable to describe the two-phase flow up

487

to the dryout.

488

4.3. Boundary conditions and simulation methodology

489

The model boundary conditions are established:

490

• firstly, from direct and indirect measurements obtained during a discharge

491

test of the thermal storage prototype; 28

492

• secondly, from the obtained results by means of a pressure drop model in

493

water-steam side. The main pressure drop model hypothesis is that the

494

thermohydraulic instabilities strongly evidenced in the mass flow rates

495

during discharge did not affect the dryout and, therefore, an average be-

496

havior of them can be assumed.

497

The discharge test of October 29th , 2012 has been used. This test was devised

498

and carried out by researchers and technicians from ENEA and provided to the

499

OPTS project partners to development different tasks. Briefly, it takes 10400

500

s and comprises three intervals (see Figures 9 and 10): the first one, from the

501

1650 s to 5250 s (left red area), is a transient interval to achieve the test nominal

502

conditions (PIN ≈ 39 bar, ∆Tsub ≈ 0 y m ˙ ws ≈ 0.082 kg/s), the second one, from

503

the 5250 s to 11350 s (central green area), is a quasi-stationary interval, and

504

the third one, from the 11350 s to 12050 s (right red area), is also a transient

505

interval to get the safe shutdown conditions (PIN ≈ 0 bar, ∆Tsub ≈ 0 y m ˙ ws ≈

506

0 kg/s). Figures 9 and 10 show some measurements during the test.

Figure 9: Temperatures (on both sides) and mass flow rate of molten salts during the test of October 29th , 2012 (in red, transient intervals and in green quasi-stationary interval).

29

Figure 10: Individual pressures and total mass flow rate of water-steam during the test of October 29th , 2012 (in red, transient intervals and in green, quasi-stationary interval).

507

In particular, in Figure 9 direct measurements are: inlet water temperature

508

to each helical coil, {TI 8 20#}#=1,2,3 , and molten salts temperature at steam

509

generator inlet, TI 8 111.

510

511

And indirect measures are: molten salts mass flow rate through steam generator, m ˙ ms , which is calculated by the following energy balance: ¯ OU T − H ¯ IN ) m ˙ ws (H m ˙ ms = ¯ CP,ms (TI 8 111 − TI 8 110)

(17)

512

¯ IN and H ¯ OU T , the where m ˙ ws is the total water-steam mass flow rate; H

513

average enthalpies of inlet water and outlet steam respectively and C¯P,ms the

514

molten salts average specific heat in the temperature range.

515

Inlet water average turbulent kinetic energy and inlet water average turbu-

516

lence dissipation rate to each helical coil, {k¯IN# }#=1,2,3 and {¯ εIN# }#=1,2,3 , are

30

517

given by the expressions [32]: 3 k¯IN# = 2

v¯IN#

I¯IN#

2

; ε¯IN# =

Cµ3/4

3/2 k¯IN#

di

(18)

518

where {¯ vIN# }#=1,2,3 is the inlet water average velocity and {I¯IN# }#=1,2,3

519

the inlet water average turbulence intensity to each helical coil. Cµ a turbulence

520

model constant pre-established by the software developer [33]. In this case, an

521

inlet fully developed flow has been assumed to each helical coil, so: ¯ −1/8 I¯IN# = 0.16 Re IN#

522

523

(19)

¯ IN }#=1,2,3 the inlet water average Reynolds number to each hebeing {Re # lical coil, based on di [32].

524

Molten salts average turbulent kinetic energy and molten salts average tur-

525

bulence dissipation rate at steam generator inlet have been considered negligible

526

because they are far enough from the studied area.

527

Finally, obtained measurements from pressure drop model are the outlet

528

steam pressures to each helical coil, {POU T# }#=1,2,3 (Figure 10). Since, the

529

simulation computational cost of the entire discharge interval is very high, the

530

following methodology has been adopted.

531

• Firstly, two times within the quasi-stationary interval has been select,

532

denoted by t7050s and t10650s (see Figures 9 and 10). These, have been se-

533

lected because, on the one hand, they are far enough from the startup/shutdown

534

intervals to assume negligible their transient effects and, on the other hand,

535

from t7050s to t10650s , molten salts mass flow rate decreases up to a 12 %,

536

which allows to study its influence on the dryout. Each selected time

537

belongs to a two-phase flow instability cycle. These instabilities can be

538

seen within the time evolution of inlet water temperatures and pressures,

539

{TI 8 20#}#=1,2,3 and {PIN# }#=1,2,3 and in outlet steam temperatures

540

and pressures, {TI 8 21#}#=1,2,3 and {POU T# }#=1,2,3 (see Figures 9 and

541

10).

31

542

• Secondly, the representative cycles of each selected time, denote by t¯7050s

543

and t¯10650s , have been made and simulated. Taking into account the

544

measured instabilities period is 20 s (compatible with Naitoh et al. results

545

[26]), for example, t¯7050s is made by averaging the measurements between

546

(t7050s − 10 s) and (t7050s + 10 s), and the associated uncertainties have

547

been assumed as the standard deviations in such interval.

548

Finally, the numerical approach provides a second order precision in the

549

spatial resolution of main conservation equations. These have been carried out

550

R R in 12 nodes DELL M630 of 20 cores Intel Xeon [email protected] by

551

node, with processing times about 26 hours by simulation.

552

5. Model evaluation

553

The model evaluation conditions are established from direct and indirect

554

measurements during the test and pressure drop model results, taking into ac-

555

count the methodology described above.

556

In particular, direct measurements are: outlet steam temperature to each

557

helical coil, {TI 8 21#}#=1,2,3 , molten salts temperature at steam generator

558

outlet, TI 8 110 (see Figure 9) and local wall temperatures along helical coil

559

3, {TI 8 3n1}n=0,1,...,9 , corresponding to heights {zn }n=0,1,...,9 (see Figure 11).

560

These last, considering a local reference system in polar coordinates centered in

561

such helical coil, {r3 , θ3 }, are always in the same position: r3 = do and θ3 = 0.

562

˙ is the And, the exchanged power between molten salts and water-steam, Q,

563

indirect magnitude, calculated using the following expression: ¯ OU T − H ¯ IN ) Q˙ = m ˙ ws (H

564

565

566

567

(20)

¯ IN and H ¯ OU T , the where m ˙ ws is the total water-steam mass flow rate; H average enthalpies of inlet water and outlet steam respectively. The obtained magnitudes from pressure drop model are inlet water pressures to each helical coil, {PIN# }#=1,2,3 (Figure 10).

32

Figure 11: Wall thermocouples distribution along the internal helical tube (left) and zoom of the n-th (right)

568

In Table 7, simulated and observed outlet steam temperature to each heli-

569

cal coil, {TI 8 21#}#=1,2,3 , and molten salts temperature at steam generator

570

outlet, TI 8 110, are shown at t¯7050s and t¯10650s , as well as, the corresponding

571

uncertainties. Table 7: Comparison between simulated and observed temperatures.

t¯7050s

t¯10650s

Simulated

Observed

Simulated

Observed

476

475 ± 2

470

469 ± 2

TI 8 212 ( C)

476

470 ± 2

468

459 ± 2

TI 8 213 (◦ C)

475

474 ± 2

466

467 ± 2

305

305 ± 2

277

278 ± 2

TI 8 211 (◦ C) ◦



TI 8 110 ( C)

33

572

As can be seen, simulated temperatures fit well with observed at both times.

573

Nevertheless, there is a slight overestimation in helical coil 2 (possibly, due to a

574

thermocouple placement error). Therefore, it could be said the model correctly

575

captures the global thermal behavior, both the molten salts and the water-

576

steam, within the quasi-stationary interval.

577

In Table 8, simulated and observed exchanged power between molten salts

578

and water-steam are shown at t¯7050s and t¯10650s . In this table, the corresponding

579

uncertainties are not presented due to its calculation difficulty. Table 8: Comparison between simulated and observed exchanged powers.

t¯7050s Q˙ (W)

t¯10650s

Simulated

Observed

Simulated

Observed

189834

192114

189023

189663

580

As can be observed, there is a good agreement between numerical and ex-

581

perimental results at both times, with differences of 2280 W in the first time

582

(equivalent to 1.2 % of percent error) and of 640 W in the second one (equivalent

583

to 0.3 % of percent error). Thus, it could be also claimed the model correctly

584

captures the total heat transfer between molten salts and the water-steam within

585

the quasi-stationary interval.

586

Simulated and observed inlet water pressures to each helical coil, {PIN# }#=1,2,3 ,

587

at t¯7050s and t¯10650s , as well as, the corresponding uncertainties are shown in

588

Table 9. Table 9: Comparison between simulated and observed pressures.

t¯7050s

t¯10650s

Simulated

Observed

Simulated

Observed

PIN1 (bar)

39.0

39.2 ± 0.6

37.3

37.9 ± 0.6

PIN2 (bar)

39.0

39.2 ± 0.6

37.5

37.9 ± 0.6

PIN3 (bar)

39.1

39.2 ± 0.6

37.6

37.9 ± 0.6

34

589

As can be appreciated, simulated pressures fit well with observed ones at

590

both times. So, the model properly captures the water-steam global hydraulic

591

behavior within the quasi-stationary interval.

592

In Figure 12, local wall temperatures along helical coil 3, {TI 8 3n1}n=0,1,...,9 ,

593

observed measurements (filled dots) and simulated values (empty dots) at t¯7050s

594

(in red) and t¯10650s (in blue) are shown, as well as, the corresponding uncer-

595

tainties.

Figure 12: Observed and simulated local wall temperatures (helical coil 3).

596

597

It is shown, simulated local wall temperatures fit well with observed up to heights: z5 at t¯7050s and z3 at t¯10650s .

598

From it, the model overestimates the wall heating. It is probably due to this

599

model is not suitable to completely describe the water-steam phase change in

600

this case, as it had been advanced during the benchmarking when average void

601

fractions were high.

602

As has been mentioned before, the experimental setup does not have other

603

direct measurements in water-steam side than those already used for validation. 35

604

In this case, it is logical because, on the one hand, “traditional” measure-

605

ments generally are invasive techniques (such as the use of thermocouples, etc.)

606

and, given the coils dimensions, they would strongly disturb the two-phase flow

607

masking its local phenomenology and, on the other hand, more “sophisticated”

608

measurements, generally non-invasive techniques (such as the use of X-rays,

609

etc.), would not be techno-economically viable.

610

At the dryout onset, the cooling of the walls drastically modifies because

611

they change from being: liquid wetted walls to steam wetted walls. This causes

612

a sharp decrease of local heat transfer coefficients causing a poor fluid-structure

613

thermal response. It implies a local increase of walls temperature regarding to

614

the downstream walls temperature [10].

615

Thus, using a wall thermocouples distribution along the tubes can be con-

616

sidered a non-invasive technique for indirect measure the dryout z-localization

617

(as it is stated, for example, in [58]).

618

619

Here, from the helical coil 3 wall thermocouples distribution, the dryout onset in such helical coil has been experimentally determined.

620

As can be observed in Figure 9, the first increase of walls temperature is

621

between heights: z6 y z5 (TI 8 361 − TI 8 351 = 71 ◦ C) at t¯7050s and, z4 y z3

622

(TI 8 341 − TI 8 331 = 69 ◦ C) at t¯10650s .

623

However, by means CFD simulations the dryout z-location, both first dryout,

624

zx1 ,3 , and total dryout, zxT ot ,3 , on such helical coil can be directly determined

625

from steam quality azimuthal distributions on walls.

626

627

In Table 10, the simulated and observed dryout z-location along helical coil 3, are shown at t¯7050s and t¯10650s , as well as, the corresponding uncertainties. Table 10: Comparison between simulated and observed dryout z-location in helical coil 3.

t¯7050s

zx1 ,3 /zxT ot ,3 (m)

628

t¯10650s

Simulated

Observed

Simulated

Observed

1.16/1.17

1.13 ± 0.063

1.35/1.36

1.38 ± 0.063

Taking into account the problem uncertainties, these results reveal, simu36

629

lated dryout z-locations along helical coil 3 fit well the observations at both

630

times.

631

Hence, it can be assumed the model performance is good enough to capture

632

the two-phase flow local behavior, at least up to the dryout onset, within the

633

quasi-stationary interval. As it had been stated before during the benchmarking.

634

In summary, this model is considered appropriate to deal with the work

635

objectives.

636

6. Dryout study

637

Hereafter, the dryout onset (first and total dryout) is studied in the covered

638

working conditions range, i.e. it is associated azimuthal position and local steam

639

qualities, depending on the helical coil diameter. And finally, the comparison

640

with the found literature correlations is carried out.

641

6.1. Azimuthal position and local steam qualities

642

From numerical results, dryout z-location, both first and total, on each he-

643

lical coil tube are obtained. Results at t¯7050s and t¯10650s are shown in Table

644

11. Table 11: Dryout z-location: first/total.

zx1 ,3 /zxT ot ,3 (m)

zx1 ,2 /zxT ot ,2 (m)

zx1 ,1 /zxT ot ,1 (m)

t¯7050s

1.16 / 1.17

1.14 / 1.15

1.11 / 1.12

t¯10650s

1.35 / 1.36

1.33 / 1.34

1.30 / 1.31

645

From them it follows that under practically the same working conditions of

646

in water-steam side (PIN ∈ [37.9, 39.2] bar, ∆Tsub ≈ 0 and m ˙ ws ≈ 0.082 kg·s-1 )

647

and molten salts side (atmospheric pressure and TIN ∈ [472, 477] ◦ C) but a 12

648

% mass flow rate decrease, both first and total dryout:

649

650

• shift along the steam generator almost 11 turns upwards (0.19 m), i.e., regardless on the helical coil diameter, D,

37

651

• happen primarily on helical coil 1, almost 2 turns upwards on helical coil

652

2 and almost 3 turns upwards on helical coil 3, regardless on molten salts

653

mass flow rate, m ˙ ms .

654

These results quantify that it was already knew: as discharge evolves, heat

655

transfer is increasingly less effective, so a greater heat transfer area along the

656

steam generator is needed to achieve the dryout, being the outer tube (helical

657

coil 1), in which, dryout, first and total, primarily happens. This, at equal z-

658

location height, has a greater heat transfer area than the others (helical coils 2

659

and 3).

660

Once determined dryout z-locations on each tube, the respective steam qual-

661

ity azimuthal distributions on walls during first dryout, {x1,# }#=1,2,3 , as well

662

as the steam quality bulk values both during first and total dryout, x1,bulk and

663

xT ot,bulk respectively, are plotted on the cross sections to the flow containing

664

such z-locations at t¯7050s and t¯10650s . The obtained results are presented in

665

Figure 13.

Figure 13: Steam quality azimuthal distributions on walls and bulk values, during first and total dryout at t¯7050s (left) and t¯10650s (right).

666

From them, it follows that:

667

• The first dryout azimuthal positions, {θx1 ,# }#=1,2,3 , remain constant re38

668

gardless on D and m ˙ ms (θx1 ≈ 170◦ , on the intrados). This is also true

669

for the lowest steam quality azimuthal positions on walls (θx1 ≈ 350◦ , on

670

the extrados). This result can be explained due to the secondary flow

671

presence.

672

On a two-phase flow in a helical coil, the centrifugal and gravity forces act

673

trying to separate the flow due to the density differences between both phases.

674

In addition, there is a secondary flow, superimposed to the primary flow, which

675

also acts on the two-phase flow. This is perpendicular to the main flow and

676

follows lines forming loops, dragging the fluid from the extrados to the intrados,

677

counteracting the separation effects between both phases [12]. The separation

678

effects depend on m ˙ ws since the buoyancy forces have higher influence at low

679

m ˙ ws due to the existing lower turbulence.

680

To visualize the secondary flow presence in this case, in Figure 14 two-phase

681

flow distribution obtained by means the line integral convolution of the vector

682

velocity field (removing the normal component and projecting the tangential

683

component on the cross section to the flow) at z6 and t¯7050s is shown as example.

Figure 14: Line integral convolution of tangential vector velocity field in helical coils at z6 and t¯7050s field (steam generator axis placed on the left side of the figure).

684

As can be seen, the streamlines shape two vortices coincident with first

685

dryout and lowest steam quality on walls azimuthal positions, which represent

686

the loops dragging the fluid from the extrados to the intrados. They go through

687

the walls perimeters and then close through the bulks, as it is shown in Figure

688

15. 39

Figure 15: Tangential vector velocity field in helical coils at z6 and t¯7050s (steam generator axis placed on the left side of the figure).

689

This phenomenology agree with that described by other authors as Pointer

690

et al. [59], Colombo [34], Jo et al. [52] and Cioncolini et al. [60], who also

691

employed CFD techniques to describe forced convective boiling flows in heli-

692

cal coils but using different working conditions, wall heating methodologies or

693

geometric characteristics than here.

694

By other hand, from steam quality azimuthal distributions on walls, the

695

best/worst chilled surface points during the first dryout can be deducted. In

696

Figure 16, temperature azimuthal distributions on walls during the first dryout,

697

{Tx1 ,# }#=1,2,3 , as well as the respective bulk values, Tx1 ,bulk , are plotted on the

698

cross sections to the flow at t¯7050s and t¯10650s .

40

Figure 16: Temperature azimuthal distributions on walls and bulk values during 1st dryout at t¯7050s (left) and t¯10650s (right).

699

It is observed the temperature difference between first dryout and lowest

700

steam quality on walls azimuthal position is ≈ 20 ◦ C. This lack uniformity on

701

the walls thermal behavior at equal z-location (even promoted by the two-phase

702

flow instabilities) could affect the steam generator safety. Although it is outside

703

the scope of this work, it is recommended the steam generator thermomechanical

704

behaviour will be also analyzed by those who pretend to scale-up this system.

705

706

• Both x1,bulk and xT ot,bulk remain constant regardless on D and m ˙ ms (x1,bulk ≈ 0.94 and xT ot,bulk ≈ 0.97).

707

These values can be understood by means the 2D map of Berthoud and

708

Jayanti, which helps to identify the process controlling the first dryout in helical

709

coils. In Figure 17, dimensionless numbers x0 and y0 of each helical coil are

710

represented, both at t¯7050s and t¯10650s .

41

Figure 17: B&J 2D map during the first dryout at t¯7050s and t¯10650s .

711

As can be observed, the study case belongs to the “Redeposition zone”, char-

712

acterized by high steam qualities during the first dryout. In addition, according

713

to the obtained vales of x0 and y0 (very close to each other), it can be deduced

714

the steam quality distributions within the helical coils follow the same pattern,

715

as it is shown in Figure 18.

Figure 18: Two-phase flow patterns at z6 and t¯7050s (steam generator axis placed on the left side of the figure).

716

This pattern is typical from the “Redeposition zone” and has been also ob-

42

717

served in the experiments conducted by Murai et al. [61] using a computed

718

tomography technique, so this model reproduces available experiments in bibli-

719

ography.

720

In summary, it can be said that, in the covered working conditions range

721

and considering this steam generator geometry, although D1 > D2 > D3 , the

722

two-phase flow behavior is uniform. Besides, that what happens at t¯7050s is

723

reproduced, without loss of generality, at t¯10650s , except a certain heights dif-

724

ference. And, therefore, the prototype operation seems not to influence on the

725

dryout.

726

6.2. Simulation results and literature correlations comparison

727

Once defined local steam qualities as {¯ x1,# }#=1,2,3 or {¯ xT ot,# }#=1,2,3 , it is

728

proceed to compare simulated values, {¯ x1,sim,# }#=1,2,3 or {¯ xT ot,sim,# }#=1,2,3 ,

729

with those calculated by using literature correlations applicable to the ENEA

730

prototype working condition for both first dryout, {¯ x1,correlac,# }#=1,2,3 (cor-

731

relations of: Berthoud and Jayanti, Ruffell and Santini et al., see expressions

732

and applicability ranges in Table 1) and total dryout, {¯ xT ot,correlac,# }#=1,2,3

733

(correlations of: Naitoh et al. and Berthoud and Jayanti, see expressions and

734

applicability ranges in Table 1), where simulated values have been calculated as

735

follows:

x ¯1orT ot,sim,# 736

737

738

1 = S

Z x1orT ot,# (r# , θ# ) dS

(21)

S

being {x1orT ot,# }#=1,2,3 the steam quality distributions on the cross sections to the flow during first or total dryout respectively. Note that one of the main handicaps assessing {¯ x1,# }#=1,2,3 or {¯ xT ot,# }#=1,2,3

739

by means experimental correlations is the boundary heat flux calculus (Q). This

740

happens mainly in those cases in which the wall heating is not uniform, as in this

741

case. Nevertheless, in correlations where Q is taken into account (correlation

742

of Berthoud and Jayanti for first dryout, and correlations of Ruffell and Naitoh

743

et al. for total dryout), the necessary power, Q˙ nec , to achieve the dryout, here

43

744

is uniformly distributed on the walls. Considering this hypothesis, obtained

745

results for {¯ x1,# }#=1,2,3 are shown in Figure 19.

Figure 19: Comparison between simulated and calculated results for first dryout.

746

As can be seen, obtained results using Ruffel0 s correlation fit well simulated

747

results, while those obtained using Santini et al. correlation underestimate

748

slightly (around 20 %) local steam qualities during first dryout. Finally, those

749

obtained using Berthoud and Jayanti correlation overestimate them up to 60

750

%. This is due to, in the first case, the correlation (considering 270◦ Ruffel0 s

751

correlation, which corresponds to 180◦ in the local polar reference systems used

752

here) was validated for water as working fluid in a set of helical coils whose

753

working conditions and characteristics were not far away from that established

754

in this work.

755

Regarding Santini et al. correlation, although the considered working fluid

756

is water and the covered working conditions range is supported here, the helical

757

coils characteristics are different: greater vertical pich, Xver , and diameters of

758

helical coil, D, and tube, di .

44

759

To a better understanding of Santini et al. results, it can be used the

760

Berthoud and Jayanti 2D map [16]. If dimensionless numbers x0 and y0 for

761

a helical coil with Santini et al. characteristics but ENEA prototype work-

762

ing conditions are represented on this map, obtained points during first dryout

763

would be shifted to the “Gravity zone”, characterized by lower steam qualities

764

than the “Redeposition Zone”. Taking into account Santini et al. helical coil

765

diameter is between 5 and 7 times larger than those of the ENEA prototype, is

766

clear the effect of centrifugal forces decreases against the gravity force.

767

In the Berthoud and Jayanti correlation case, notwithstanding working con-

768

ditions and helical coils characteristics comprise those of this case, obtained

769

results does not fit adequately, probably because this correlation was based on

770

the experimental data of five researchers (Breus and Belyakov [20], Carver et

771

¨ al. [21], Roumy [17], Styrikovich et al. [14] and Unal [18, 19]), who employed

772

different working fluids (and therefore, working conditions).

773

Figure 20 shows obtained results for {¯ xT ot,# }#=1,2,3 .

Figure 20: Comparison between simulated and calculated results for total dryout.

45

774

As can be observed, obtained results using Naitoh et al. correlation fits

775

quite well simulated results, while those obtained using Berthoud and Jayanti

776

correlation overestimate them up to 60 % the local steam qualities during total

777

dryout as it was expected. Despite the good fit in the first case, the explicit

778

independence of this correlation with the mass flux allows to question it. Be-

779

sides, there are big differences between Naitoh at al. helical coils and ENEA

780

prototype characteristics.

781

7. Conclusions

782

• In this work, the dryout onset (first and total dryout) in a triple helical

783

coil once-through steam generator has been experimentally and numeri-

784

cally studied. For this purpose, a CFD modelling based on steady-state

785

simulations and a conjugate heat transfer model has been carried out.

786

Besides, a numerical methodology based on the two-phase flow average

787

behavior in each thermohydraulic instability cycle has been employed.

788

• From a benchmarking between the two-phase flow model and a mechanis-

789

tic model, it is deduced that the used semi-empiric model can be applied

790

to the simulation of the forced convection boiling flow in the helical coils.

791

• Numerical results have been evaluated, first, against experimental data

792

from the state-of-the-art, and second, against acquired measurements dur-

793

ing a discharge test of the thermal energy storage prototype with inte-

794

grated steam generator belonging to ENEA and the obtained results with

795

a pressure drop model developed ad hoc for this case, getting good fits up

796

to the dryout.

797

• By means two phase flow tangential vector velocity field on cross section

798

to the flow has been possible to show the secondary flow presence in the

799

helical coils. And, by means steam quality distributions has been possible

800

to characterize the dryout onset, azimuthal position and associated local

801

steam qualities, as function of helical coil diameter. In this case, within the 46

802

working conditions range and considering these particular steam generator

803

characteristics, the two-phase flow behavior is uniform. And, therefore,

804

the prototype operation seems not to influence on the dryout. This phys-

805

ical phenomenology and analysis have not been reported yet on available

806

literature.

807

• Besides, the derived two-phase flow pattern is typical of “Redeposition

808

Zone” within the 2D map of Berthoud and Jayanti [16] and it is like

809

those observed in other works using, both experimental and numerical

810

methodologies, to describe forced convective boiling flows in helical coils

811

at high steam qualities.

812

• Finally, the numerical local steam qualities have been compared with those

813

calculated by using appropriate experimental correlations. From the ob-

814

tained scatter-plots it is concluded, the 270◦ Ruffel0 s correlation is the

815

most adequate expression to determine the local steam quality value dur-

816

ing the first dryout.

817

• Also, assuming a 0.97 value during the total dryout is a good approxi-

818

mation, since although it is the most conservative choice, it matches the

819

value recommended by Mazufri [62] for the thermohydraulic design of a

820

helical steam generator. On the other hand, the Santini et al. correlation

821

offers suitable results (within a 20 % error) being this one, moreover, sim-

822

pler than others because it does not depend on walls heat flux. So, the

823

obtained steam quality macro values are validated, and it is justified the

824

interest of a comprehensive study of them.

825

ACKNOWLEDGEMENTS

826

The authors would like to acknowledge the EU through the 7th Framework

827

Program for the financial support of this work under the OPTS project with

828

contract number: 283138, the UTRINN-STD from Casaccia Research Center

829

(ENEA) for providing the experimental data and the information concerning 47

830

to the experimental setup and the Extremadura Research Centre for Advances

831

Technologies (CETA-CIEMAT) for providing the necessary computing resources

832

for simulations. Technical discussions with members of ATYCOS-CIEMAT were

833

essential for this work.

834

Appendix A. Statistical parameters The relative error (RE) is defined as: RE[%] =

|Cp − Co | Co

The correlation coefficient (R), the normalized mean square error (N M SE), the fractional bias (F B) and the fraction of prediction within a factor of two of observations (F AC2), are given by: R=

(Co − Co )(Cp − Cp ) (Co − Cp )2 (Co − Cp ) ; N M SE = ; FB = σCp · σCp Co · Cp 0.5(Co + Cp ) F AC2 = fraction of data that satisfy 0.5 ≤

Cp ≤ 2.0 Co

835

where: Cp are model predictions; Co observations; and C and σC the corre-

836

sponding average and standard deviation, respectively, over the data-set.

837

Note that since F B measures only the systematic bias of a model, it is

838

possible for a model to have predictions completely out of phase of observations

839

and still have F B = 0.0 because of cancelling errors.

48

NOMENCLATURE Tube diameter (m)

Q

Heat flux (W·m-2 )

f

Force (N)



Power (W)

g

Gravity (9.806 65 m·s-2 )

Re

Reynolds number

h

Static enthalpy (J·kg-1 )

S

Surface (m2 )

d

-1

-1

k

Thermal conductivity (W·m ·K )

T

Temperature (K)

m ˙

Mass flow rate (kg·s-1 )

V

Volume (m3 )

t

Time (s)

X

Pitch (m)

{x, y, z}

Cartesian coordinates (m)

α

Vapor volume fraction

-1

v

Velocity (m·s )

x

Steam quality

x ¯

Local steam quality

{x0 , y0 }

Dimensionless numbers (B&J 2D Map)



Turbulent dissipation rate (m2 ·s-3 )

-1

CP

Specific heat (J·kg ·K )

κ

Turbulent kinetic energy (J·s-1 )

D

Helical coil diameter (m)

µ

Dynamic viscosity (Pa·s)

E

Energy (J)

ρ

Density (kg·m-3 )

-1

-1

-2

G

Mass flux (kg·s ·m )

σ

Surface tension (N·m-1 )

H

Enthalpy (J·kg-1 )

ϕ

Helix angle (◦ )

L

Length (m)



Increment

P

Pressure (bar)

{r, θ}

Polar coordinates

Pr

Prandtl number

{r, θ, z}

Cylindrical coordinates

SUBSCRIPTS 1

First

l

Liquid

sat

Saturated

bulk

Bulk

lat

Latent

simu

Simulated

correlac

Correlated

m

Mixture

sub

Subcooled

cri

Critic

max

Maximum

t

Turbulent

ef f

Effective

min

Minimum

T ot

Total

g

Gravity

ms

Molten salts

v

Vapor

hor

Horizontal

nec

Necessary

ver

Vertical

i

Inner

o

Outer

w

Wall

IN

Inlet

OU T

Outlet

ws

Water-steam

49

ACRONYMS CFD

Computational Fluid Dynamics

RANS

Reynolds Averaged Navier-Stokes

CSP

Concentrating Solar Power

SFR

Sodium-cooled Fast Reactor

DWO

Density Wave Oscillations

[1] E. Zarza, Principales retos de I+D y potencial contribuci´on a un mercado sostenible, in: Gu´ıa T´ecnica de La Energ´ıa Solar Termoel´ectrica, Fundaci´on de la Energ´ıa de la Comunidad de Madrid (FENERCOM), 2012, Ch. 3, pp. 71–122. [2] ENEA. URL http://www.enea.it/it [3] ANSALDO Nucleare S.p.A. URL https://www.ansaldoenergia.com/ansaldo-nucleare/home [4] L. Rinaldi, F. Fabrizi, A. Alemberti, W. Gaggioli, S. Aliotta, A. Barbensi, P. Tarquini, Thermal Energy Storage Tank With Integrated Steam Generator, EP 2652272 B1 (2015-10-28). [5] F. Fabrizi, W. Gaggioli, L. Rinaldi, C. Rocca, R. Surace, Progettazione e realizzazione di uno scambiatore di calore prototipico per reattori refrigerati a sodio con circuito intermedio a sali fusi, Tech. Rep. NNFISS LP3 - 027, ENEA (2011). [6] F. Fabrizi, Trough Molten Salt HTF Field Test Experience, in: NREL Parabolic Trough Technology Workshop, 2007, p. 59. [7] W. Gaggioli, P. Di Ascenzi, L. Rinaldi, P. Tarquini, F. Fabrizi, Effects assessment of 10 functioning years on the main components of the molten salt PCS experimental facility of ENEA, in: AIP Conference Proceedings, Vol. 1734, AIP Publishing, 2016, p. 070009. doi:10.1063/1.4949156. [8] W. Gaggioli, F. Fabrizi, P. Tarquini, L. Rinaldi, Experimental validation of the innovative thermal energy storage based on an integrated system 50

‘storage tank/steam generator’, Energy Procedia 69 (2015) 822–831. doi: 10.1016/j.egypro.2015.03.091. [9] W. Gaggioli, F. Fabrizi, F. Fontana, L. Rinaldi, P. Tarquini, An innovative concept of a thermal energy storage system based on a single tank configuration using stratifying molten salts as both heat storage medium and heat transfer fluid, and with an integrated steam generator, Energy Procedia 49 (2014) 780–789. doi:10.1016/j.egypro.2014.03.085. [10] J. G. Collier, J. R. Thome, Convective boiling and condensation, Clarendon Press, 1994. [11] CD-adapco, Star-CCM+ 9.04.011 (2014). URL https://mdx.plm.automation.siemens.com/star-ccm-plus [12] A. Owhadi, K. J. Bell, B. Crain Jr, Forced convection boiling inside helically-coiled tubes, International Journal of Heat and Mass Transfer 11 (12) (1968) 1779–1793. doi:10.1016/0017-9310(68)90021-5. [13] M. Cumo, G. E. Farello, G. Ferrari, The influence of curvature in post dryout heat transfer, International Journal of Heat and Mass Transfer 15 (11) (1972) 2045–2062. doi:10.1016/0017-9310(72)90030-0. [14] M. A. Styrikovich, V. S. Polonsky, V. V. Reshetov, Experimental investigation of the critical heat flux and post-dryout temperature regime of helical coils, International Journal of Heat and Mass Transfer 27 (8) (1984) 1245–1250. doi:10.1016/0017-9310(84)90052-8. [15] M. Naitoh, A. Nakamura, H. Ogasawara, Dryout in helically coiled tube of sodium heated steam generator, ASME Paper 74-WA/HT-48 (1974). [16] G. Berthoud, S. Jayanti, Characterization of dryout in helical coils, International Journal of Heat and Mass Transfer 33 (7) (1990) 1451–1463. doi:10.1016/0017-9310(90)90042-S.

51

[17] R. Roumy, Dryout in helically coiled tubes with boiling Freon-12, European Two-Phase Flow Group Meeting, Risp. (1971). ¨ [18] H. C. Unal, Some aspects of two-phase flow, heat transfer and dynamic instabilities in medium and high pressure steam generators, Ph.D. thesis, Delft University of Technology (1981). ¨ [19] H. C. Unal, M. Van Gasselt, P. Van’t Verlaat, Dryout and two-phase flow pressure drop in sodium heated helically coiled steam generator tubes at elevated pressures, International Journal of Heat and Mass Transfer 24 (2) (1981) 285–298. doi:10.1016/0017-9310(81)90036-3. [20] V. I. Breus, I. I. Belyakov, Burnout in helical coils at high pressures, Thermal Engineering 30 (10) (1983) 592–593. [21] J. R. Carver, C. R. Kakarala, J. S. Slotnik, Heat transfer in coiled tubes with two-phase flow, Tech. Rep. TID-20983, Babcock and Wilcox Co. Research Center, Alliance, Ohio (1964). [22] L. Duchatelle, L. d. Nucheze, M. G. Robin, Heat transfer in helical tube sodium heated steam generators, Tech. Rep. CEA-CONF–3204, CEA Centre d’Etudes Nucleaires de Saclay (1975). [23] M. K. Jensen, A. E. Bergles, Erratum:‘Critical Heat Flux in Helically Coiled Tubes’(Journal of Heat Transfer, 1981, 103, pp. 660–666), Journal of Heat Transfer 105 (1) (1983) 216–216. doi:10.1115/1.3245553. [24] M. K. Jensen, A. E. Bergles, Critical heat flux in helical coils with a circumferential heat flux tilt toward the outside surface, International Journal of Heat and Mass Transfer 25 (9) (1982) 1383–1395. doi:10.1016/ 0017-9310(82)90132-6. [25] M. K. Jensen, A. E. Bergles, Critical heat flux in helically coiled tubes, Journal of Heat Transfer 103 (4) (1981) 660–666. doi:10.1115/1.3244523.

52

[26] M. K. Jensen, Boiling heat transfer and critical heat flux in helical coils, Ph.D. thesis, Iowa State University (1980). [27] A. E. Ruffell, The application of heat transfer and pressure drop data to the design of helical coil once-through boilers, Multiphase Flow Systems Meet., Glasgow. [28] L. Santini, A. Cioncolini, C. Lombardi, M. Ricotti, Dryout occurrence in a helically coiled steam generator for nuclear power application, in: EPJ Web of Conferences, Vol. 67, EDP Sciences, 2014, p. 02102. doi:10.1051/ epjconf/20146702102. [29] E. Rivas, E. Rojas, Heat transfer correlation between Molten Salts and helical-coil tube bundle Steam Generator, International Journal of Heat and Mass Transfer 93 (2016) 500–512. doi:10.1016/j.ijheatmasstransfer. 2015.10.013. [30] E. Rivas, E. Rojas, R. Bay´on, W. Gaggioli, L. Rinaldi, F. Fabrizi, CFD model of a molten salt tank with integrated steam generator, Energy Procedia 49 (2014) 956–964. doi:10.1016/j.egypro.2014.03.103. [31] AutoDesk, AD. Inventor Professional 2016 (2016). URL https://www.autodesk.es/products/inventor/overview [32] H. K. Versteeg, W. Malalasekera, An introduction to computational fluid dynamics: the finite volume method, Pearson Education, 2007. [33] CD-adapco, Star-CCM+ User’s Guide - Version 9.04.011, CD-adapco, 2014, (web book not open-access). [34] M. Colombo, Experimental investigation and numerical simulation of the two phase flow in the helical coil steam generator, Ph.D. thesis, Politecnico di Milano (2013). [35] M. Sommerfeld, B. Van Wachem, R. Oliemans, Best practice guidelines for computational fluid dynamics of dispersed multi-phase flows, European Re53

search Community on Flow, Turbulence and Combustion (ERCOFTAC), 2008. [36] W. Xu, Q. Chen, F. T. M. Nieuwstadt, A new turbulence model for nearwall natural convection, International Journal of Heat and Mass Transfer 41 (21) (1998) 3161–3176. doi:10.1016/s0017-9310(98)00081-7. [37] M. Wolfshtein, The velocity and temperature distribution in onedimensional flow with turbulence augmentation and pressure gradient, International Journal of Heat and Mass Transfer 12 (3) (1969) 301–318. doi:10.1016/0017-9310(69)90012-x. [38] R. W. Bradshaw, R. W. Carling, A review of the chemical and physical properties of molten alkali nitrate salts and their effect on materials used for solar central receivers, ECS Proceedings Volumes 1987 (1987) 959–969. [39] W. M. Rohsenow, A method of correlating heat transfer data for surface boiling of liquids, Tech. Rep. MIT-Heat Transfer Laboratory no. 5., Cambridge, Mass.: MIT Division of Industrial Cooporation (1951). [40] J. R. Thome, Wolverine engineering databook III, Wolverine Industries, 2006, (web book updated yearly). [41] E. Inox, Roughness measurements of stainless steel surfaces (2014). [42] Y. A. Cengel, Heat Transfer: A Practical Approach, McGraw-Hill series in mechanical engineering, McGraw-Hill, 2003. [43] J. R. Cooper, R. B. Dooley, The international association for the properties of water and steam revised release on the IAPWS industrial formulation 1997 for the thermodynamic properties of water and steam, International Association for the Properties of Water and Steam 35 (2007) 3–49. [44] M. Z. Podowski, Toward mechanistic modeling of boiling heat transfer, Nuclear engineering and technology 44 (8) (2012) 889–896. doi:10.5516/ NET.02.2012.720. 54

[45] E. Krepper, B. Konˇcar, Y. Egorov, CFD modelling of subcooled boilingconcept, validation and application to fuel assembly design, Nuclear Engineering and Design 237 (7) (2007) 716–731. doi:10.1016/j.nucengdes. 2006.10.023. [46] B. Konˇcar, E. Krepper, Y. Egorov, CFD modeling of subcooled low boiling for nuclear engineering applications, in: Proc. Int. Conference Nuclear Energy for New Europe 2005, 2005, pp. 140.1–140.14. [47] Y. Egorov, F. Menter, Experimental implementation of the rpi wall boiling model in cfx-5.6, Tech. rep. ANSYS/TR-04-10, ANSYS Germany GmbH (2004). [48] J. Jabardo, E. Silva, G. Ribatski, S. F. de Barros, Evaluation of the rohsenow correlation through experimental pool boiling of halocarbon refrigerants on cylindrical surfaces, Journal of the Brazilian Society of Mechanical Sciences and Engineering 26 (2) (2004) 218–230. doi:10.1590/ S1678-58782004000200015. [49] P. Domalapally, E. Rizzo, L. S. Richard, F. Subba, R. Zanino, CFD analysis of flow boiling in the iter first wall, Fusion Engineering and Design 87 (5-6) (2012) 556–560. doi:10.1016/j.fusengdes.2012.01.024. [50] D. H. Lob´ on, E. Baglietto, L. Valenzuela, E. Zarza, Modeling direct steam generation in solar collectors with multiphase CFD, Applied Energy 113 (2014) 1338–1348. doi:10.1016/j.apenergy.2013.08.046. [51] E. Baglietto, STAR-CCM+ Boiling Model Validation. URL

https://mdx2.plm.automation.siemens.com/sites/default/

files/technical_document/pdf/Homogeneous_Boiling_STARCCM%2B. pdf [52] J. C. Jo, W. S. Kim, C.-Y. Choi, Y. K. Lee, Numerical simulation of subcooled flow boiling heat transfer in helical tubes, Journal of Pressure Vessel Technology 131 (1) (2009) 011305. doi:10.1115/1.3028022. 55

[53] G. Bartolomei, G. Batashova, V. Brantov, et al., Teplomassoobmen-iv, vol. 5 (1980). [54] A. A. Avdeev, V. Pekhterev, Vapor condensation in non-equilibrium bubble flows, Teplofizika Vysokikh Temperatur 24 (6) (1986) 1125–1131. [55] G. Bartolomei, V. Chanturiya, Experimental study of true void fraction when boiling subcooled water in vertical tubes, Thermal Engineering 14 (2) (1967) 123–128. [56] V. Ustinenko, M. Samigulin, A. Ioilev, S. Lo, A. Tentner, A. Lychagin, A. Razin, V. Girin, Y. Vanyukov, Validation of CFD-BWR, a new two-phase computational fluid dynamics model for boiling water reactor analysis, Nuclear Engineering and Design 238 (3) (2008) 660–670. doi:10.1016/j.nucengdes.2007.02.046. [57] S. Lo, Progress in modeling boiling two-phase flows in boiling water reactor fuel assemblies, in: Proc. of Workshop on Modeling and Measurements of Two-Phase Flows and Heat Transfer in Nuclear Fuel Assemblies, Citeseer, 2006, pp. 10–11. [58] K. Moriyama, A. Inoue, H. Ohira, The thermohydraulic characteristics of two-phase flow in extremely narrow channels (the frictional pressure drop and heat transfer boiling two-phase flow, analytical model), Heat TransferJapanese Research 21 (8) (1993) 838–856. [59] W. D. Pointer, D. Shaver, Y. Liu, P. Vegendla, A. Tentner, Evaluation of CFD Methods for Simulation of Two-Phase Boiling Flow Phenomena in a Helical Coil Steam Generator, Tech. Rep. ORNL/TM-2016/612 NT0512000; NENT026; TRN: US1700827, Oak Ridge National Laboratory (ORNL) (2016). doi:10.2172/1335356. [60] A. Cioncolini, A. Cammi, C. Lombardi, L. Luzzi, M. Ricotti, L. Cinotti, G. Castelli, Thermal hydraulic analysis of iris reactor coiled tube steam

56

generator, in: American Nucl. Society Topical Meeting in Mathematics and Computations, 2003, p. 9. [61] Y. Murai, H. Oiwa, T. Sasaki, K. Kondou, S. Yoshikawa, F. Yamamoto, Backlight imaging tomography for gas-liquid two-phase flow in a helically coiled tube, Measurement Science and Technology 16 (7) (2005) 1459. doi: 10.1088/0957-0233/16/7/008. [62] C. M. Mazufri, Dise˜ no Thermohidraulico del Generador de Vapor de un Paso Tipo Helicoidal Modular, in: AATN 2000: 27 Reunion anual de la Asociacion Argentina de Tecnologia Nuclear (AATN), 2000, p. 12.

57

Highlights



CFD modelling of dryout onset in a triple helical coil once-through steam generator.



Benchmarking of two-phase flow model against experimental data from the state-ofthe-art.



Evaluation of numerical results with experimental data of a thermal energy storage prototype with integrated steam generator during a discharge test.



Azimuthal distributions analysis of the two-phase flow as function of helical coils characteristics and working conditions.



Comparison of numerical results with those obtained from adequate correlations for forced convective boiling flows in vertical helical coils.

Declaration of interests ☒ The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper. ☐The authors declare the following financial interests/personal relationships which may be considered as potential competing interests: