nuclear engineering and technology - core · was made to follow the best practice guidelines for...

11
Original Article Validation of Computational Fluid Dynamics Calculation Using Rossendorf Coolant Mixing Model Flow Measurements in Primary Loop of Coolant in a Pressurized Water Reactor Model Istvan Farkas a , Ezddin Hutli a,b,* , Tatiana Farkas a , Antal Tak acs a , Attila Guba a , and Iv an T oth a a Department of Thermohydraulics, Centre for Energy Research, Hungarian Academy of Sciences, Konkoly-Thege Mikl os u ´ t 29-33, 1121 Budapest, Hungary b Budapest University of Technology and Economics, Institute of Nuclear Techniques (BME-INT), M} uegyetem rakpart 9, 1111 Budapest, Hungary article info Article history: Received 10 April 2015 Received in revised form 5 January 2016 Accepted 28 February 2016 Available online 18 March 2016 Keywords: Cold Leg Downcomer Mixing Thermal Shocks Thermal Fatigue Temperature Velocity abstract The aim of this work is to simulate the thermohydraulic consequences of a main steam line break and to compare the obtained results with Rossendorf Coolant Mixing Model (ROCOM) 1.1 experimental results. The objective is to utilize data from steady-state mixing experi- ments and computational fluid dynamics (CFD) calculations to determine the flow distri- bution and the effect of thermal mixing phenomena in the primary loops for the improvement of normal operation conditions and structural integrity assessment of pres- surized water reactors. The numerical model of ROCOM was developed using the FLUENT code. The positions of the inlet and outlet boundary conditions and the distribution of detailed velocity/turbulence parameters were determined by preliminary calculations. The temperature fields of transient calculation were averaged in time and compared with time- averaged experimental data. The perforated barrel under the core inlet homogenizes the flow, and therefore, a uniform temperature distribution is formed in the pressure vessel bottom. The calculated and measured values of lowest temperature were equal. The inlet temperature is an essential parameter for safety assessment. The calculation predicts pre- cisely the experimental results at the core inlet central region. CFD results showed a good agreement (both qualitatively and quantitatively) with experimental results. Copyright © 2016, Published by Elsevier Korea LLC on behalf of Korean Nuclear Society. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/ licenses/by-nc-nd/4.0/). * Corresponding author. E-mail address: [email protected] (E. Hutli). Available online at ScienceDirect Nuclear Engineering and Technology journal homepage: www.elsevier.com/locate/net Nuclear Engineering and Technology 48 (2016) 941 e951 http://dx.doi.org/10.1016/j.net.2016.02.017 1738-5733/Copyright © 2016, Published by Elsevier Korea LLC on behalf of Korean Nuclear Society. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).

Upload: others

Post on 26-Apr-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Nuclear Engineering and Technology - CORE · was made to follow the best practice guidelines for the use of CFD in nuclear safety applications [34]. The generated mesh consisted of

eDirect

Nu c l e a r E n g i n e e r i n g a n d T e c h n o l o g y 4 8 ( 2 0 1 6 ) 9 4 1e9 5 1

Available online at Scienc

Nuclear Engineering and Technology

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

Original Article

Validation of Computational Fluid DynamicsCalculation Using Rossendorf Coolant MixingModel Flow Measurements in Primary Loop ofCoolant in a Pressurized Water Reactor Model

Istvan Farkas a, Ezddin Hutli a,b,*, Tatiana Farkas a, Antal Tak�acs a,Attila Guba a, and Iv�an T�oth a

a Department of Thermohydraulics, Centre for Energy Research, Hungarian Academy of Sciences, Konkoly-Thege

Mikl�os ut 29-33, 1121 Budapest, Hungaryb Budapest University of Technology and Economics, Institute of Nuclear Techniques (BME-INT), M}uegyetem rakpart

9, 1111 Budapest, Hungary

a r t i c l e i n f o

Article history:

Received 10 April 2015

Received in revised form

5 January 2016

Accepted 28 February 2016

Available online 18 March 2016

Keywords:

Cold Leg

Downcomer

Mixing

Thermal Shocks

Thermal Fatigue

Temperature

Velocity

* Corresponding author.E-mail address: [email protected]

http://dx.doi.org/10.1016/j.net.2016.02.0171738-5733/Copyright © 2016, Published by Elthe CC BY-NC-ND license (http://creativecom

a b s t r a c t

The aim of this work is to simulate the thermohydraulic consequences of a main steam line

break and to compare the obtained results with Rossendorf Coolant Mixing Model (ROCOM)

1.1 experimental results. The objective is to utilize data from steady-state mixing experi-

ments and computational fluid dynamics (CFD) calculations to determine the flow distri-

bution and the effect of thermal mixing phenomena in the primary loops for the

improvement of normal operation conditions and structural integrity assessment of pres-

surized water reactors. The numerical model of ROCOM was developed using the FLUENT

code. The positions of the inlet and outlet boundary conditions and the distribution of

detailed velocity/turbulence parameters were determined by preliminary calculations. The

temperature fields of transient calculation were averaged in time and compared with time-

averaged experimental data. The perforated barrel under the core inlet homogenizes the

flow, and therefore, a uniform temperature distribution is formed in the pressure vessel

bottom. The calculated and measured values of lowest temperature were equal. The inlet

temperature is an essential parameter for safety assessment. The calculation predicts pre-

cisely the experimental results at the core inlet central region. CFD results showed a good

agreement (both qualitatively and quantitatively) with experimental results.

Copyright © 2016, Published by Elsevier Korea LLC on behalf of Korean Nuclear Society. This

is an open access article under the CC BY-NC-ND license (http://creativecommons.org/

licenses/by-nc-nd/4.0/).

(E. Hutli).

sevier Korea LLC on behalf of Korean Nuclear Society. This is an open access article undermons.org/licenses/by-nc-nd/4.0/).

Page 2: Nuclear Engineering and Technology - CORE · was made to follow the best practice guidelines for the use of CFD in nuclear safety applications [34]. The generated mesh consisted of

Fig. 1 e Cold leg (CL-1)/downcomer model used for inlet

Nu c l e a r E n g i n e e r i n g a n d T e c h n o l o g y 4 8 ( 2 0 1 6 ) 9 4 1e9 5 1942

1. Introduction

Currently, many nuclear reactors worldwide are nearing their

originally prescribed lifetime of approximately 40 years. In

response, several initiatives have been undertaken to address

the main problems in this regard, so that the existing power

plants can be safely kept in operation for a longer period (up to

60 years) and new power plants can be designed for 80 years or

longer. The two biggest problems affecting the stability of

existing power plants are thermal fatigue (TF) and pressurized

thermal shocks (PTSs). TF is defined as a state (failure) caused

by cyclical thermal stress. PTS is characterized by rapid cool-

ing (i.e., thermal shock) of the downcomer and internal

reactor pressure vessel (RPV) surface, which is sometimes

followed by repressurization of the RPV. Thus, a PTS event

poses a potentially significant challenge to the structural

integrity of the RPV in a pressurized water reactor (PWR) and

water-cooled and water-moderated energy reactor. It is an

important phenomenon in nuclear power plants manage-

ment and safety assessment. The key mixing and flow dis-

tribution phenomena are important factors to be considered

while performing safety analysis of nuclear reactors. The

mixing process is important to completely or partially over-

come the consequences of accidents that may happen during

reactor operation such as reactivity insertion, overcooling

transients, thermal stress, or pressure thermal stress. The

mixing process is of relevance even for normal reactor oper-

ation, (e.g., for determination of the coolant temperature

distribution at the core inlet in the case of partially switched

off main circulation pumps) [1]. When the mixing process

takes place between two streams with a strong temperature

difference (e.g., residual heat removal in a reactor), a strong

density gradient also exists. This indicates how the density

differences affect the mixing of coolant streams and why it is

assumed to be an important parameter for predicting areas

susceptible to TF [2,3]. When two fluid streams of significantly

different temperatures mix before reaching homogeneity,

they can expose a section of pipe wall to periodic fluctuations

of temperature and potentially result in fatigue cracking. The

oscillating temperature also leads to pressure thermal stress

[3]. After several TF events that have recently occurred in

various nuclear power plants, the focus of thermal striping

studies included both fast breeder reactors (FBRs) and light-

water reactors [4e6]. Some examples of TF events in nuclear

power plants are those of the French FBR in 1992 and PWR in

1998, as well as the Japanese PWR in 1999 and in 2003 [4,6,7].

Mixing as a process is relevant for structural integrity and

nuclear safety. The TF and PTS phenomena could be limited

by improving the mixing process quality in the system [8].

Because of the lack of thermocouples to detect the high degree

of turbulence (mixing) effects, many researchers have inves-

tigated the TF and PTS phenomena numerically and/or

experimentally, based on the fluidmixing phenomena in the T

junction and in the complex hydraulic system of a power

plant [9e16]. Nonintrusive techniques such as particle image

velocimetry and laser-induced fluorescence associated with

the thermocouples network were also used in this field and

useful results have been presented [5,17e21]. The perfor-

mance of mixing fluid for enhancing the safety of nuclear

reactors is still a subject for investigation both experimentally

and numerically. As an example, in the last several years,

extensive experimental work has been done using Rossendorf

Coolant Mixing Model (ROCOM) facility, supplementing the

existing database about the mixing phenomena in PWRs with

important new data. In addition, there are a number of vali-

dation works on the applicability of computational fluid dy-

namics (CFD) methods for simulating ROCOM mixing

experiments [22e33].

This paper describes the flow field and the temperature

differences near the downcomer cold-leg inlets (T junction)

that have been measured, illustrated, and quantified by nu-

merical approaches on the experimentmodel used by ROCOM

Test 1.1, performed within the Organisation for Economic Co-

operation and Development PKL2 Project (common project),

with the participation of the Center for Energy Research,

Budapest. The Standard k-ε (FLUENT 6.3) and Reynolds stress

model (RSM) turbulence models, applied in preliminary cal-

culations, showed the same tendencies. For the simulation of

ROCOM experiment 1.1, an RSM was chosen because of the

complex geometry of ROCOM facility. Themixing process was

investigated in many cross sections in the pressure vessel.

Two cross sections before the T junction, between the cold leg

and pressure vessel (the cross sections are marked as “a” and

“b” in Fig. 1), were selected because it is a common component

in the cooling systems ofmost nuclear power plants that has a

high exposure to TF [12]. The comparison has been done be-

tween the results obtained by ROCOM mixing experiments

(ROCOM 1.1) and CFD.

1.1. Objective and the motivation

The objective of this work is not only to validate the CFD

calculation using ROCOM 1.1 measurements but also to create

and optimize the numerical model (mesh) in FLUENT, which

would minimize the computational efforts of simulating

mixing processes in reactor vessel for given conditions. The

aim of the ROCOM 1.1 experiments was to investigate the ef-

fects of water density differences between the primary

coolant loop and the emergency core cooling (ECC) systems on

the mixing in the downcomer. ROCOM 1.1 experiments

matrices were used for the calculation. The density difference

of 12% between one of the Emergency Core Cooling Systems

(ECCs) and the primary loop was used in the calculation.

The ROCOM test facility was used to perform numerous

experiments with different flow conditions (e.g., flow rates

velocity profile determination (cross section a and b).

Page 3: Nuclear Engineering and Technology - CORE · was made to follow the best practice guidelines for the use of CFD in nuclear safety applications [34]. The generated mesh consisted of

Fig. 2 e Streamlines of flow from cold leg 1 (CL-1) colored according to velocity amplitudes. (A) Calculation with constant

density (kg/m3). (B) Calculation with changing density (kg/m3).

Nu c l e a r E n g i n e e r i n g a n d T e c h n o l o g y 4 8 ( 2 0 1 6 ) 9 4 1e9 5 1 943

and density differences). The parameters of ROCOM test 1.1

were selected from the result of a steam collector transient

analysis. In this case, the water coming from the damaged

loop is overcooled, which could cause thermal stress in the

reactor vessel. This was a motivation to investigate the flow

characteristics in the interrogation area.

To study this process, it is necessary to check and validate

the capability of our numerical model created with FLUENT

code. The expected results from this calculationwork could be

useful for developing different reactor vessels’ numerical

models for similar conditions.

2. Methods

2.1. Numerical mesh design and performance ofpreliminary calculations

Because of the complexity and the extent of the geometry of

the ROCOM facility, the meshing was a key task during the

model design and development. To simulate ROCOM experi-

ment 1.1, the numerical mesh was built up gradually by per-

forming test calculations. In this way, we could verify the

correctness of this complex numerical model. Special effort

was made to follow the best practice guidelines for the use of

CFD in nuclear safety applications [34]. The generated mesh

consisted of around 4.6 million hexahedral cells. Attention

was given to components that significantly influence the ve-

locity field (i.e., the mixing process), such as the core barrel

with the lower core support plate and core simulator, the

perforated drum in the lower plenum, and inlet nozzles. In the

initial phase, the calculation’s domain and modeling were

taken for the downcomer and the connecting cold leg nozzles.

The inlet velocity boundary condition was defined at cold leg

nozzles, whereas an outlet pressure boundary condition was

defined at the downcomer bottom. Dimensionless wall dis-

tance (yþ) was studied based on test calculation results. It was

found that when themesh size near the wall was about 2mm,

the maximum yþ value of the mesh was compatible with the

validity range of wall law [34]. More details about the ROCOM

experimental facility, the accessories, and experimental pro-

cedures can be found in related publications [25,35]. The

simulation of mixing in the CFD calculation is taken into ac-

count by temperature differences instead of the concentration

differences, which were used in experimental work. The

reason for using the FLUENT code is that this code contains

the broad physical modeling capabilities needed to model

flow, turbulence, heat transfer, and reactions for industrial

applications. The applied physical models, such as the tur-

bulence model, wall function, and porosity model were

selected based on preliminary test calculations.

2.2. Influence of fluid density on the velocity field

Figs. 2A and 2B show the calculation results using the ROCOM

downcomer model while investigating the density effect (the

calculation was done with and without density difference). In

the case of isothermal flow in the calculation, both tempera-

ture and density were kept constant. The results show that

the streamline structures from cold leg 1 (CL-1) are deter-

mined exclusively by inertial forces, and thus the fluid

arriving from the nozzle creates a butterfly-like flow picture

hitting the internal wall of the downcomer (Fig. 2A). The ve-

locity is zero in the downcomer because the downcomer

diameter is considerably bigger than the diameter of cold leg.

In the second case (i.e., the nonisothermal flow), when the

simulation was done with a significant density difference

(12%) between the liquid of the downcomer and the liquid

injected fromCL-1, the results showed that the gravity force is

dominant and controls the flow characteristic, and therefore,

streamlines turn downward like a plume (Fig. 2B). The

streamlines shown are colored according to velocity ampli-

tude values. In Fig. 2B the driving force is gravity, and there-

fore, it is expected that in the lowest section of the

downcomer bottom there is no flow, that is, the flow will

arrive at zero value (i.e., stagnation area). Onemay notice that

velocity of incoming water in downcomer along streamlines

Page 4: Nuclear Engineering and Technology - CORE · was made to follow the best practice guidelines for the use of CFD in nuclear safety applications [34]. The generated mesh consisted of

Fig. 3 e The perforated barrel.

Nu c l e a r E n g i n e e r i n g a n d T e c h n o l o g y 4 8 ( 2 0 1 6 ) 9 4 1e9 5 1944

in Fig. 2B is greater than that in Fig. 2A. These results show the

importance of physical properties of the fluid in the determi-

nation of its flow characteristics.

2.3. Influence of perforated drum, core support plate,and fuel rods

To investigate the effects of components that significantly

influence the velocity field, the calculation domain was

improved by including inlet nozzles, downcomer, lower

plenum, the reactor core, and inlet and outlet nozzles. The

flow compensator-perforated barrel (perforated drum) in the

lower plenum of the pressure vessel was modeled as a porous

medium (Fig. 3). Calculating a model of this barrel (410 holes)

would drastically increase the resolution of calculation

meshes and computational demand, which was avoided. The

geometry of the numerical model is in agreement with the

original barrel geometry shown in Fig. 3. It consists of three

cylindrical elements separated from each other by solid rings.

The porosity of rings was specified by three different values in

the calculation (Fig. 3).

The calculation results are presented in Figs. 4A and 4B.

The velocity vector fields are shown for a vertical plane cross

section of ROCOM. Fig. 4A shows the results when the perfo-

rated barrel is not introduced into the calculation; therefore,

the fluid coming from the downcomer gets to the vessel bot-

tomwithout deceleration and flows to the opposite side along

Fig. 4 e Velocity vector field in the lower part of the numerical m

the perforated barrel. (B) Without the perforated barrel.

the pressure vessel wall. Fig. 4B shows the result when the

barrel wasmodeled as a porous zone. The 193 holes in the core

support plate and the 193 tubes representing the core of

ROCOM were also included. As can be clearly seen, the flow

coming from the downcomer drastically slows down when

reaching the barrel, due to the hydraulic resistance, and a

homogeneous velocity field develops below the core inside the

barrel.

To see the sensitivity of output results for the applied

boundary conditions, further test calculations have been

performed. Extending the core model length by 20 cm showed

that the position of outlet boundary does not havemuch effect

on the results, so for the final simulation, the original shorter

core was applied. For the CL-1 region, the inlet boundary

conditions showed stronger influence on the obtained results.

To reduce this effect, the inlet boundary was shifted further

from the vessel (Figs. 1 and 5).

The computational domain was developed with the

Gambit geometry and mesh processing software. The applied

computational grid consists of 4,581,836 hexahedral nodes. A

quality test was run on this grid where the normalized

skewness of meshing was studied. This value can vary be-

tween 0 and 1, where 0 corresponds to the ideal case (the angle

of the edges is 90�) and 1means an inside-out mesh (the angle

of the edges is 0�). According to these studies, the mesh

skewness is between 0.0 and 0.1 for 47% of the elements,

whereas the maximum value is 0.65, which is under the 0.778

upper limit proposed in a previous study [34].

2.4. Set up of numerical model initialization

The steady-state calculation is performed based on the

ROCOMexperiment 1.1 database. Density differences between

the primary coolant loop and the ECC water in the CFD

calculation are achieved by temperature differences rather

than the concentration differences, as used in ROCOM

experiment 1.1. According to volumetric flow rates, which

were adjusted according to Table 1, flow velocities have been

calculated based on experimental flow rates and specified to

be constant at pressure vessel nozzles with velocity of

odel (m/s) in the vertical plane of the downcomer. (A) With

Page 5: Nuclear Engineering and Technology - CORE · was made to follow the best practice guidelines for the use of CFD in nuclear safety applications [34]. The generated mesh consisted of

Fig. 5 e Computational model of the experimental facility. (A) Side view and (B) Top view.

Table 1 e Boundary conditions of ROCOM experiment 1.1.

Cold leg Normalizedvolume

flow rate (%)

Volumeflow rate

(L/sec) (nominalvalue: 185 m3/hr)

Relativedensity

1 12.21 6.27 1.12

2 3.15 1.62 1.00

3 3.15 1.62 1.00

4 3.15 1.62 1.00

ROCOM, Rossendorf Coolant Mixing Model.

Nu c l e a r E n g i n e e r i n g a n d T e c h n o l o g y 4 8 ( 2 0 1 6 ) 9 4 1e9 5 1 945

0.3548m/s and 0.0917m/s for CL-1 and CL-2e4, respectively. It

is well-known that the bends in the cold leg pipes affect the

flow profile [20], and so the asymmetry of velocity profiles at

the pressure vessel inletmust be considered in the finalmodel

[34]. Inlet velocity profiles have been determinedwith the help

Fig. 6 e Effect of gravitational force on the velocity profile of wa

means Cross section).

of auxiliary simulations. For calculations, the cold leg/down-

comer model has been used as shown in Fig. 1. Velocity inlet

boundary conditions were applied to the modeled cold leg

inlet and the nozzles of other cold legs. Pressure outlet

boundary conditions were applied to the bottom of the

downcomer. In the modeled pipe, the velocity profile was

saved in the cross section, which corresponded to the final

model inlet (“a” denotes this section in Fig. 1). Because CL-2, 3,

and 4 are not considered in this case, the flow rates from them

are constant and identical; the velocity fields in CL-2 and CL-4

are identical, whereas the velocity profile in CL-3 is affected by

their reflection (CL-2 and CL-4).

Fig. 6 shows the primary results of the case of the cold leg

concerned (CL-1) with both isothermal and nonisothermal

flows. To avoid the inaccuracies thatmay come frommodeling

simplifications (e.g., the outlet boundary was defined at the

bottomof the downcomer), velocity profilewas recorded inCL-

1 at 10 cm away from the downcomer, assuming that the hot

ter entering reactor pressure vessel from CL-1 (m/s). (Cut

Page 6: Nuclear Engineering and Technology - CORE · was made to follow the best practice guidelines for the use of CFD in nuclear safety applications [34]. The generated mesh consisted of

Fig. 7 e Surface and lines in downcomer chosen to evaluate

results. The different colors represent the temperature

distribution: the highest temperature is red, whereas the

lowest temperature is dark blue.

Nu c l e a r E n g i n e e r i n g a n d T e c h n o l o g y 4 8 ( 2 0 1 6 ) 9 4 1e9 5 1946

water of nozzle does not affect the velocity field (see cross

section “b” in Fig. 1). The results in Fig. 6 reveal that the average

flow velocity in the nonisothermal flow case is greater than

that in the case of isothermal flow. At the nozzle inlet cross

section, the asymmetry of vertical velocity occurs together

with a temperature difference, which induces horizontal

asymmetryvia thedifference indensity.Hotwater comes from

the vessel into the upper part of CL-1 (close to the vessel). Its

effect extends to Section a (Cut a in Fig. 6). In general, the result

shows that there is an agreement between the measurement

results and the velocity profile distortion caused by tempera-

ture stratification. One can notice that the velocity profiles are

different in the two cases. In Section a, horizontal asymmetry

is induced by gravity. The velocity profile saved in Section b is

similar to that in Section a, so gravity still does not play an

important role here, although a temperature difference exists.

Therefore, velocity profiles are determined in auxiliary simu-

lations for ROCOM inlets. Pressure outlet boundary conditions

have been assigned to the 193 pipes representing the core.

Fig. 8 e Temperature distributions in the do

2.5. Test calculation of ROCOM experiment 1.1

The simulation process was carried out for quasi-steady-state

condition with a period of 73e83 seconds, according to the

ROCOM experiment 1.1 conditions [25]. The steady solver of

FLUENTcodewasused for calculations. Becauseof theabsence

of precise initial conditions (e.g., for the pressure field near the

perforated barrel), the obtained calculation results diverged.

Increasing fluctuations of pressure and velocity field occurred

in the calculation starting from estimated values. To avoid

thesefluctuations, the flow resistanceof the porouslymodeled

perforated barrel (drum) was decreased and then gradually

increased in calculation up to its final value (applying the

suggestions of the FLUENT handbook [34]). In such a way the

pressure and velocity fieldswere stabilized, but the calculation

still did not converge to a steady state. Rather, it fluctuated

between some limit values. This behavior of simulation can be

explained by the presence of processes on a different time

scale, for example, presence of the mixing process of the

coolant, which comes from cold legs with different velocities/

temperatures, and the coolant, which is already stagnating in

the downcomer. To model these phenomena at different time

scales, the transient calculation was performed for further

studies. For the initial state, the results of the previously

reached steady-state calculation have been applied. A time

step of 0.001 seconds was used in the beginning while the

number of iterations for one time step significantly decreased.

For this purpose, nearly 600 time steps were needed, and then

the time step was increased to five times the initial value. The

first 0.6 seconds of the time-dependent simulation was calcu-

lated with 0.001-second time steps. The remaining simulation

was calculated with 0.005-second time steps. To achieve a

quasi-stationary condition, it was necessary to simulate 110

seconds. This was the case in this experiment as well. The

water was circulating in the loops before the beginning of the

measurement to achieve a quasi-stationary condition.

3. Comparison of CFD and measurementresults

A quasi-steady density field calculation in the reactor vessel

has been developed according to ROCOM 1.1 [25]. Time

wncomer. (A) Measured. (B) Calculated.

Page 7: Nuclear Engineering and Technology - CORE · was made to follow the best practice guidelines for the use of CFD in nuclear safety applications [34]. The generated mesh consisted of

Fig. 9 e Temperature distribution along a vertical line of

the downcomer chosen between cold leg 1 and cold leg 2

sectors as in Fig. 7. CFD, computational fluid dynamics;

Exp., Experiments.

Nu c l e a r E n g i n e e r i n g a n d T e c h n o l o g y 4 8 ( 2 0 1 6 ) 9 4 1e9 5 1 947

averaging was done on measured data between 73 seconds

and 83 seconds, which were used for comparison with the

calculation results. The positions of the calculation grid built

in the downcomer are indicated with a colored surface in

Fig. 7. Azimuthal reference lines are characterized by their

different z coordinates and the axial line is indicated by its

angle. The numbers in Fig. 7 show the positions of the cold

legs. The measured and calculated results have been

compared qualitatively along the surface of pressure vessel.

Further quantitative analysis was done along black reference

lines. The temperature field of the chosen reference surface is

shown in Figs. 8A and 8B. Fig. 8A displays the measurement

results. These results show a similar flow profile: a hot water

layer develops in the upper part of the downcomer due to

buoyancy that is fed by the slow hot coolant coming from CL-

2, CL-3, and CL-4. The plume of cold water of higher density

arriving fromCL-1 vertically flows through the upper hot layer

toward the lower plenum, while entraining some hot water.

These two phenomena balance the thickness of the upper

layer. There are some differences between temperature fields

in Figs. 8A and 8B. According to measurement results, this hot

layer extends deeper in the downcomer than in the simulation

Fig. 10 e Temperature distribution along downcomer perimeter

Z¼e0.3275 m. (B) Temperature distribution along the perimete

cold leg nozzles. CFD, computational fluid dynamics; Exp., Expe

(as presented in Fig. 9); at the same time, the cold plume ex-

tends further radially as presented in Figs. 10B, 11A, and 11B,

but axially it is shorter than in the calculation. Based on these

results, the quantity of hot water entrained by the cold plume

differs in the two cases. In the numerical model, the mixing

zone around the plume was predicted to be wider and longer.

The interpretation can be given by a faster down flow of cold

water than in the measurement. Fig. 12 also confirms this

interpretation, which shows that a part of the cold plume

flows around the perforated barrel to the opposite side of the

vessel and then, flowing upward, returns toward CL-1 due to

its inertia. Because the plume was predicted to be at a higher

velocity in the calculation, it may cause a higher backflow to

the downcomer, leaving less space for the stagnating hot

water in the upper part. ROCOM experiment 1.1 calculations

of others also highlighted that the upper hot region is smaller

than in the experiment [36]. The effect of in-vessel measure-

ment devicesmay play a role in the development of suspected

differences in velocity fields, not only preventing the flow of

the cold plume, but also holding the two circulation regions on

the sides of the downcomer (Fig. 12). Therefore, there is a

larger space for the upper hot layer, just as experienced in the

measurement. The velocity difference discussed above could

be justified only experimentally. However, these kinds of ex-

periments have not yet been performed.We assumed that the

measurement of a detailed velocity field would not only help

us to decide the validity of our assumption, but it will also be a

reference to determine parameters needed to model turbu-

lence. Thus, the deviations between the measurements and

CFD results will decrease. Quantitative analysis was per-

formed along reference lines presented in Fig. 7. Fig. 9 shows

the distribution of coolant temperature along the axial refer-

ence line denoted by “2.11 rad.” The analysis of the results in

Fig. 9 revealed that the hot water layer in the experiment is

about 10 cm deeper than in the calculation. In addition, the

transition between the hot and cold water regions is much

sharper in the simulation. The coolant temperature in the

downcomer’s lower part, obtained by calculation, exceeds

that obtained by measurements, indicating that the incoming

cold water plume entrains more hot water in the simulation.

In addition, the effect of in-vessel measurement devices may

. (A) Vectors colored by temperature at elevation

r of the downcomer 0.3275 m away from the centerline of

riments.

Page 8: Nuclear Engineering and Technology - CORE · was made to follow the best practice guidelines for the use of CFD in nuclear safety applications [34]. The generated mesh consisted of

Fig. 11 e Temperature distribution along downcomer perimeter. (A) At Z¼e0.7925 m and (B) at Z¼ 1.1955 m away from the

centerline of cold leg nozzles. CFD, computational fluid dynamics; Exp., Experiments.

Nu c l e a r E n g i n e e r i n g a n d T e c h n o l o g y 4 8 ( 2 0 1 6 ) 9 4 1e9 5 1948

play a role in deviation. These small devices are not taken into

account in the calculation.

Azimuthal temperature distributions at three downcomer

elevations are displayed in Figs. 10B and 11A and 11B. The

calculated plume is typically narrower than in the measure-

ment at each level. The calculated temperature profile on the

top elevation reference level, due to emerging local minima

and maxima along the two sides of the plume (Fig. 10B, below

and above 1.6 rad), qualitatively differs from experimental

data where temperature monotonously decreases and rea-

ches lowest temperature and then starts to increase. The

evolution of simulated temperature can be explained first by

the impact of coolant from CL-1 on the core support basket

and then on the accumulated hot water in stagnation. The

impact directs the cold flow back toward the external wall of

the vessel (Fig. 11A) forming local minima in the temperature

profile. This concept is also supported by the structure of

streamlines coming from CL-1 (Fig. 12).

Fig. 11A displays the azimuthal temperature profile in the

middle part of the downcomer. In this region, the

Fig. 12 e Streamlines coming from CL-1 colored according

to velocity amplitude (CFD model), (m/s).

CFD, computational fluid dynamics.

experimental and calculated profiles show good agreement

both qualitatively and quantitatively. The maximum relative

temperature is about 0.6. The relative temperature is defined

as Trel ¼ ðTmax � TÞ=Tmax. The hottest parts are formed near

the cold plume where the water flowing downward entrains

the hotter coolant. This region is called “mixing zone” (circu-

lation regions) in the description of Fig. 12. The calculated

temperatures exceed the experiment values by 5e6�C, exceptfor the center of the cold plume. The minimum temperatures

at the center of the cold plume are in good agreement. Fig. 11B

shows the temperature evolution along the lowest reference

line of the downcomer. Just like on the profiles of Fig. 11A,

experimental temperatures are smaller than those in calcu-

lations, but the tendencies of the profiles are similar. The fact

that measured temperatures are lower below the hot layer

domainmay lead to the conclusion that mixingwasweaker in

the regions further from the plume, which is a result of slower

recirculation. In addition, it can be noticed in the axial tem-

perature distribution of Fig. 9 that the axial temperature

gradient of the experiment is smaller than in the calculation.

This refers to the more important role of buoyancy phenom-

enon in the experiment compared with convection, and

therefore, a larger density cold water layer is formed than in

the calculation. In Fig. 11B the coldest point of the plume and

the profile are still below CL-1, so the temperature is radically

Fig. 13 e Position of lines chosen to evaluate results at core

inlet. The different colors represent the temperature

distribution: the highest temperature is red, whereas the

lowest temperature is dark blue.

Page 9: Nuclear Engineering and Technology - CORE · was made to follow the best practice guidelines for the use of CFD in nuclear safety applications [34]. The generated mesh consisted of

Fig. 14 e Temperature distributions at core inlet. (A) Measured. (B) Calculated.

Nu c l e a r E n g i n e e r i n g a n d T e c h n o l o g y 4 8 ( 2 0 1 6 ) 9 4 1e9 5 1 949

increasing in both azimuthal directions. In the experiment,

the plume is distributed between positions of 0.8 rad and 2.4

rad, and the correspondingminimumvaluemoves toward CL-

2, while the temperature gradient is significantly smaller than

in the simulation. A mixing of the cold plume also decreases

its velocity and, in such a way, the coolant downflow becomes

slower. In Fig. 13, the cross section of measurement sensors

and the reference lines denoted by capital letters and used for

quantitative comparison are shown at the core inlet.

Figs. 14A and 14B illustrate measured (experimental) and

calculated temperature distributions, respectively, at core

inlet. The largest temperature difference is 0.061 in the mea-

surement, whereas it is 0.118 in the calculation. However, the

lowest values of scales are almost the same. The temperature

difference is the difference between the maximum and

Fig. 15 e Temperature distribution at core inlet (Z¼e1.955 m).

presented in Fig. 13. CFD, computational fluid dynamics; Exp.,

minimum temperatures recorded in the experiments or by the

calculations. It is assumed to be the range of temperature

fluctuation at the measuring point. The biggest temperature

difference at the core inlet in the calculation is 10�C and in the

experiment it is 5�C, whereas the lowest temperature recor-

ded at the core by the calculation and in the measurement is

“equal.” This is why different colors represent identical tem-

peratures in the inner regions of the core. The experiment and

calculated temperature fields show similar behavior in Figs. 14

and 15. In both cases, domains of quasi-homogeneous tem-

perature are formed at the centerline of core inlet, which is

surrounded by a hotter ring at the boundaries. This hot

external ring is broken by the water coming from the cold

plume below CL-1. The lowest temperature is found at this

point in bothmeasurement and calculation. Nevertheless, the

Along (A) Line A, (B) Line B, (C) Line C, and (D) Line D, as

Experiments.

Page 10: Nuclear Engineering and Technology - CORE · was made to follow the best practice guidelines for the use of CFD in nuclear safety applications [34]. The generated mesh consisted of

Nu c l e a r E n g i n e e r i n g a n d T e c h n o l o g y 4 8 ( 2 0 1 6 ) 9 4 1e9 5 1950

coldest point in the experiment as already mentioned in the

discussion about downcomer phenomena is slightly shifted

toward CL-2 as shown in Fig. 11B. Quantitative comparisons

displayed in Figs. 15Ae15D support the results in Figs. 14A and

14B. They show that the extracted profiles of temperature

correspond well in the middle bulk region of the core. De-

viations experienced in the boundaries are caused by the

varied flow patterns in the downcomer and the corresponding

temperature fields.

3.1. Conclusions

In this work, a numerical model of ROCOM has been devel-

oped. Important geometrical details of the facility have been

taken into account and their effects on flow were tested by

calculations. The position of inlet and outlet boundary con-

ditions and the distribution of detailed velocity/turbulence

parameters were determined by preliminary calculations. An

appropriate mesh resolution was obtained from numerical

tests with special regard to the suitability of nondimensional

wall distance (yþ). Because the steady-state solver, applied

according to the experiment description, did not give a

converging result, a transient simulationwas donewith initial

conditions obtained from the steady-state calculation.

Calculation results show good qualitative and quantitative

agreement with experiment results. The differences observed

in detailed studies could be related to the hydraulic resistance

of measurement grid. This may explain why the cold and

dense plume from CL-1 expands and slows down while

flowing toward the bottom of the vessel in the experiment.

The backflow of simulation to the downcomer is stronger than

in the experiment and it is resisted by the measurement grid.

This resistance slows the flow in the downcomer and results

in a more homogenous temperature distribution. At the same

time, a wider transition with smaller axial temperature

gradient is formed between the upper hot and lower cold re-

gions. In this way, the stagnating layer of hot water in the

downcomer top part extends deeper and a colder water layer

is formed in the vessel bottom than in the calculation. The

perforated barrel at vessel bottom homogenizes the flow, and

therefore, an internal domain of almost uniform temperature

is formed below the core basket. In both experiment and

calculation, the minimum and maximum temperatures occur

in the ring surrounding the internal homogeneous domain

where the water comes directly from CL-1 before the mixing

effect of the flow compensation barrel. Nevertheless, the

values of the lowest temperatures are equal in both cases.

This is essential from the point of view of safety analysis

because this parameter plays an important role when deter-

mining the reactor criticality. The calculation predicts pre-

cisely the experimental results at the core inlet central region

where the temperature is homogeneous. The temperature of

the fluid in contact with the vessel wall is crucial to determine

the vessel integrity, because temperature affects both the

stresses and the material toughness of the vessel material.

Therefore, the obtained results could be useful for further

investigation subjects, such as PTS and TF. The temperature of

the fluid in contact with the vessel wall is crucial to the

determination of vessel integrity because temperature affects

both the stresses and the material toughness of the vessel

material. According to all the aforementioned points, we can

say that the observations collected in this work can be useful

for developing different reactor vessel’s numerical models for

similar conditions.

Conflicts of interest

All authors have no conflicts of interest to declare.

r e f e r e n c e s

[1] U. Rohdea, T. H€ohnea, S. Kliema, B. Hemstr€omb,M. Scheuererc, T. Toppilad, A. Aszodie, I. Borose, I. Farkasf,P. Muhlbauerg, L. Vyskocilg, J. Klepach, J. Remish, T. Dury,Fluid mixing and flow distribution in a primary circuit of anuclear pressurized water reactordvalidation of CFD codes,Nucl. Eng. Des. 237 (2007) 1639e1655.

[2] H.D. Kweon, J.S. Kim, K.Y. Lee, Fatigue design of nuclear classpiping considering thermal stratification, Nucl. Eng. Des. 238(2008) 1265e1274.

[3] E.C. Eggertson, R. Kapulla, J. Fokken, H.M. Prasser, Turbulentmixingand itseffectson thermal fatigue innuclear reactors, Int.J. Math. Comput. Phys. Electr. Comput. Eng. 52 (2011) 206e213.

[4] K.J. Metzner, U. Wilke, Thermal Fatigue Evaluation of PipingSystem Tee-Connections (THERFAT) [Internet]. 2007 [cited2016 Mar 23]. Available from: http://cordis.europa.eu/documents/documentlibrary/97848051EN6.pdf.

[5] S.,M. Hosseini, K. Yuki, H. Hashizume, Classification ofturbulent jets in a T-junction area with a 90-deg bendupstream, Int. J. Heat Mass Transf. 51 (2007) 2444e2454.

[6] International Atomic Energy Agency (IAEA), PressurizedThermal Shock in Nuclear Power Plants: Good Practices forAssessment, Deterministic Evaluation for the Integrity ofReactor Pressure Vessel (Rep. No. IAEA-TECDOC-1627), IAEA,Vienna, Austria, 2010.

[7] C. St�ephan, High cycle thermal fatigue damage prediction inmixing zones of nuclear power plants: engineering issuesillustratedon theFATHERcase, Procedia. Eng.66 (2013)240e249.

[8] K.E. Stahlkopf, Pressure vessel integrity under pressurizedthermal shock conditions, Nucl. Eng. Des. 80 (1984) 171e180.

[9] S.M. Hosseini, K. Yuki, H. Hashizume, The three-dimensionalstudy of flow mixing phenomenon in a T-junction area with90-degree bend upstream, Proceedings of the InternationalConference “Nuclear Energy for New Europe 2005,” Bled(Slovenia), September 5e8, 2005. 037.1e037.10.

[10] C. Faidy, Thermal fatigue in mixing areas: status andjustification of French assessment method, Proceedings ofthe Third International Conference on Fatigue of ReactorComponents, EPRI/US NRC/OECD NEA, Seville, Spain, 2004.

[11] T. Muramatsu, Numerical analysis of non-stationary thermalresponse characteristics for a fluid-structure interactionsystem, J. Press. Vessel Technol. 121 (1999) 276e282.

[12] H. Ezddin, G. Val�er, F. Istv�ann�e, �E. Gy€orgy, B. G�abor,Investigation of mixing coolant in a model of reactor vesseldown-comer and in cold leg inlets using PIV, LIF, and CFDtechniques, Therm. Sci. (2015) 121e134.

[13] M. Igarashi, Study on fluid mixing phenomena for evaluationof thermal striping a mixing tee, 10th International TopicalMeeting on Nuclear Reactor Thermal Hydraulic (NURETH-10), Seoul, Korea, 2003, pp. 1e12.

[14] M. Tanaka, Experimental investigation of flow field structurein mixing tee, Proceedings of Sixth International Conferenceon Nuclear Thermal Hydraulics, Operation and Safety,(NUTHOS-6), Nara, Japan, 2004.

Page 11: Nuclear Engineering and Technology - CORE · was made to follow the best practice guidelines for the use of CFD in nuclear safety applications [34]. The generated mesh consisted of

Nu c l e a r E n g i n e e r i n g a n d T e c h n o l o g y 4 8 ( 2 0 1 6 ) 9 4 1e9 5 1 951

[15] S.M. Hosseini, Experimental investigation of thermal-hydraulic characteristics at a mixing tee, International HeatTransfer Conference (FCV-17), Sydney, Australia, 2006.

[16] N. Kimura, M. Nishimura, H. Kamide, Study on convectivemixing for thermal striping phenomena: experimentalanalyses on mixing process in parallel triple-jet andcomparisons between numerical methods, JSME Int. J. 45(2002) 592e599.

[17] H. Ogawa, M. Igarashi, N. Kimura, H. Kamide, Experimentalstudy on fluid mixing phenomena in T-pipe junction withupstream elbow, 11th International Topical Meeting onNuclear Reactor Thermal-Hydraulics (NURETH-11), PopesPalace Conference Center, Avignon, France, 2005.

[18] N. Aminia, Y.A. Hassan, Measurements of jet flowsimpinging into a channel containing rod bundle usingdynamic PIV, Int. J. Heat Mass Transf. 52 (2009) 5479e5495.

[19] Fluent, Inc, FlowLab 1.2, Turbulent Flow and Heat Transfer ina Mixing Elbow, 2007. Available from: https://www.yumpu.com/en/document/view/3338571/turbulent-flow-and-heat-transfer-in- a-mixing-elbow/6.

[20] V. Gottlasz, E. Hutli, D. Tar, G. �Ezs€ol, G. Baranyai, Measuringof velocity and temperature field in a model of reactor vessel,downcomer and cold-legs inlets, 10th InternationalSymposium on Particle Image VelocimetrydPIV13, Delft, theNetherlands, July 2e4, 2013, pp. 1e9.

[21] H. Ezddin, G. Valer, T. D�aniel, E. Gyorgy, B. Gabor,Experimental approach to investigate the dynamics ofmixing coolant flow in complex geometry using PIV and PLIFtechniques, Therm. Sci. 19 (2015) 989e1004.

[22] T. H€ohne, S. Kliem, U. Rohde, F.-P. Weiss, Boron dilutiontransients during natural circulation flow inPWRdexperiments and CFD simulations, Nucl. Eng. Des. 238(2008) 1987e1995.

[23] T. H€ohne, S. Kliem, Modeling of a buoyancy-driven flowexperiment in pressurized water reactors using CFD-methods, Nucl. Eng. Technol. 39 (2007) 327e336.

[24] T. H€ohne, Numerical simulation of coolant mixing in apressurized water reactor with different CFD methods basedon complex meshes, Int. J. Nucl. Energy Sci. Technol. 3 (2013)399e412.

[25] S. Kliem, T. H€ohne, U. Rohde, F.-P. Weiss, Experiments onslug mixing under natural circulation conditions at theROCOM test facility using high-resolution measurementtechniques and numerical modeling, Nucl. Eng. Des. 240(2010) 2271e2280.

[26] V. Petrov, A. Manera, Effect of pump-induced cold-leg swirlson the flow field in the RPV of the EPRTM: CFD investigations

and comparison with experimental results, Nucl. Eng. Des.241 (2011) 1478e1485.

[27] G. Pochet, X. Zhang, D. L�eonard, CFD simulations ofbuoyancy driven flow mixing experiments performed at theROCOM facility, Presented at the 14th International TopicalMeeting on Nuclear Reactor Thermal Hydraulics (NURETH-14), Toronto, Canada, September 25e30, 2011.

[28] V. Petrov, A. Manera, Validation of Star-CCMþ for buoyancydriven mixing in a PWR reactor pressure vessel, Presented atthe 14th International Topical Meeting on Nuclear ReactorThermal Hydraulics (NURETH-14), Toronto, Canada,September 25e30, 2011.

[29] S. Kliem, H.-M. Prasser, T. Suhnel, F.-P. Weiss, A. Hansen,Experimental determination of the boron concentrationdistribution in the primary circuit of a PWR after a postulatedcold leg small break loss-of-coolant:accident with cold legsafety injection, Nucl. Eng. Des. 238 (2008) 1788e1801.

[30] H.-M. Prasser, G. Grunwald, Th. H€ohne, S. Kliem, U. Rohde,F.-P. Weiß, Coolant mixing in a PWR-de-boration transients,steam line breaks and emergency core coolinginjectiondexperiments and analyses, Nucl. Technol. 143(2003) 37e56.

[31] M.J. da Silva, S. Thiele, T. H€ohne, R. Vaibar, U. Hampel,Experimental studies and CFD calculations for buoyancydriven mixing phenomena, Nucl. Eng. Des. 240 (2010)2185e2193.

[32] A. Bucalossi, F. Moretti, D. Melideo, A. Del Nevo, F. D’Auria,T. H€ohne, E. Lisenkov, D. Gallori, Experimental investigationof in-vessel mixing phenomena in a VVER-1000 scaled testfacility during unsteady asymmetric transients, Nucl. Eng.Des. 241 (2011) 3068e3075.

[33] S.T. Jayaraju, P. Sathiah, E.M.J. Komen, E. Baglietto, LargeEddy simulation for an inherent boron dilution transient,Nucl. Eng. Des. 262 (2013) 484e498.

[34] J.H. Mahaffy, B. Chung, F. Dubois, F. Ducros, E. Graffard,M. Heitsch, M. Henriksson, E. Komen, F. Moretti, T. Morii,P. Muhlbauer, U. Rohde, M. Scheuerer, B.L. Smith, C. Song,T. Watanabe, G. Zigh, Best Practice Guidelines for the Use ofCFD in Nuclear Reactor Safety, Nuclear Energy Agency, Paris,France, 2007 (Rep. No. NEA/CSNI/R5).

[35] H.-M. Prasser, A. B€ottger, J. Zschau, A new electrode-meshtomograph for gaseliquid flows, Flow Meas. Instrum. 9 (1998)111e119.

[36] S. Kliem, Comparison of the results of the pre-testcalculations with the experimental data for ROCOM tests 1.1and 1.2, 1st PKL2 Analytical Workshop, Pisa, Italy, 2010.