Influence of injection temperature and flow rate on mixing and stratification in small passively cooling enclosures

NUCLEAR ENERGY SCIENCE AND ENGINEERING

Influence of injection temperature and flow rate on mixing and stratification in small passively cooling enclosures

Wei-Qian Zhuo
Feng-Lei Niu
Jun-Chi Cai
Xiao-Wei Su
Ying-Qiu Hu
Yun-Gan Zhao
Yu Yu
Nuclear Science and TechniquesVol.27, No.4Article number 91Published in print 20 Aug 2016Available online 11 Jul 2016
5000

Thermal mixing and stratification phenomena may occur during the loss of a coolant accident (LOCA) or main steam line break accident (MSLB) in the containment of a Passive Containment Cooling System (PCCS), or in the suppression pools in BWR. However, the present study pays insufficient attention to the thermal stratification phenomena in the containment of small modular reactors (SMR). In this paper, an investigation on the mixing and thermal stratification phenomena caused by the plumes or buoyant jets in SMR containments was carried out. The experiments were both conducted under non-adiabatic and adiabatic conditions for a steel containment. In each condition, two key parameters, inlet temperature, and flow rate were tested by controlling variables to identify their influence on the thermal stratification phenomenon. The visualization experiments illustrated the jet mixing and stratification development. The experiment results were compared with the numerical computation and they reached a good agreement.

Mixing and thermal stratificationSMRVisualization experimentsNumerical computation

1 Introduction

The Passive Containment Cooling System (PCCS) is an important part of the advanced PWR passive safety systems. The containment plays a critical role in heat transfer during the design basis accidents[1]. Different from most of the PWR’s pre-stress concrete containments, the AP1000 has two independent containments, the inner cylindrical steel container and the outer reinforced concrete structure. The cylindrical steel container with elliptical upper and lower heads is the main heat transfer surface. During the loss of a coolant accident (LOCA) and main steam line break accident (MSLB), steam is cooled down on the inner wall of steel containment, and heat can be transferred to the outside of the containment. The heat is ultimately transferred to atmosphere through air convection, heat radiation, and water film evaporation. In the containment, one of the key issues relevant to passive cooling or natural heat convection is the thermal mixing and stratification phenomena which may occur during the LOCA or MSLB [2].

The suppression pools in BWR are another example where thermal stratification has been considered. Recent research indicated that the temperature of the suppression pool surface is very important to the overall containment pressure response since it determines the vapour partial pressure [3]. Therefore, thermal stratification tends to form in the pools after the initial rapid venting stage during the loss of coolant accident (LOCA) transients [4]. A one-tenth scaled-down ESBWR suppression pool at Purdue University conducted separate-effects tests indicating that significant thermal stratification is likely to exist in the pool [5]. Such a phenomenon has been carried out experimentally in the POOLEX experiments, which were performed at Lappeenranta University of Technology, Finland [6]. In addition, Haihua Zhao made the comparison numerically with a one dimensional modelling method to study phenomena relevant to thermal stratification and mixing[4].

The research mentioned above demonstrates that plumes or buoyant jets can normally be expected to stratify in large enclosures. Nevertheless, it is likely to occur in some smaller natural circulation enclosures. Peterson, in previous scaling analyses, developed the criteria for the prediction of ambient enclosure fluid stratification due to temperature [7], and for an injected buoyant jet case, the ambient fluid is stably stratified when

(Hsfdbjo)Ridbjo13(1+dbjo42αTHsf)23>1, (1) Ridbjo=(ρaρ0)gdbjoρaU02, (2)

where Hsf is the height of an enclosure, dbjo is the diameter of the jet source, αT = 0.05 is the Taylor’s jet entrainment constant,  ρa is the ambient fluid density, ρ0 is the source fluid density, g is the gravity constant, and U0 the jet source speed.

Small modular reactors (SMRs) are part of a new generation of nuclear power plant designs with an equivalent electric power of less than 700 MW(e), or even less than 300 MW(e). There is renewed interest in the development and application of SMRs due to the advantages of flexible on-site construction, heightened nuclear materials security, and increased containment efficiency. The size of the containment of the SMR is much smaller than that of PWR, but the specific value is varied. For instance, compared to the 39.62 m in diameter and 65.63 m in height of the containment in AP1000[8], the Westinghouse SMR containment design is reduced to 9.8 m in diameter and 27 m in height[9]. Interesting data from some typical plants are listed in Table 1, and it illustrates that the size of the containment or the maximum break size in the LOCA is related to its capacity.

Table 1.
Applicable dataof some SMRs
Name Capacity Developer Sizes of Containment
      Diameter (m) Height (m)
NuScale[10] 50 MWe NuScale Power + Fluor, USA 4.6 24.5
SVBR-100[11] 100 MWe AKME-engineering, Russia 4.5 7.86
Westinghouse SMR[9] 225 MWe Westinghouse, USA 9.8 27.0
VBER-300[12] 300 MWe OKBM, Russia 28.0 34.0
      Maximum break size of LOCA   (Diameter, cm)
CAREM[13] 27 MWe CNEA & INVAP, Argentina 3.81  
mPower[14] 180 MWe Babcock & Wilcox + Bechtel, USA 4.7  
Show more

The safety issues, including the effectiveness and integrity of containment, have drawn more and more attention [15]. However, the present study pays insufficient attention to the thermal stratification phenomena in the containments of SMRs. As a matter of fact, it is important in the application of PCCS technology to SMRs. Hence, it is significant to carry out research on those phenomena.

In this research, small steel containment experiments with hot air injection were conducted to study the mixing and stratification phenomena. To visualize the injection and stratification development, the simulation experiments using a small glass containment and dyed water were conducted. Two key parameters, i.e. injection temperature and flow rate, are taken into consideration to identify their influence on the mixing and thermal stratification. The experiments are conducted under non-adiabatic and adiabatic conditions.

2 Experimental

2.1 Steel containment

The experimental facility includes a steel containment, a hot air supply, and an injection system and a data acquisition system. Figure 1 presents the schematic of the facility and the details of the distribution of the measuring points on the containment.

Figure 1.
Schematic of the experimental facility
pic

The cylindrical containment is made of Type 304 stainless steel with the height of 3.34 m, a diameter of 2 m, and a wall thickness of 6 mm. The containment has an elliptical upper dome and its cylindrical lower part is embedded underground. There is a rectangle door made of a 0.3-cm-thick stainless steel plate and it can be removed when necessary. The jet tubing is horizontally inserted into the containment through this wall at its center. Air enters the containment from the jet and leaves through the exit on the bottom inside of the containment.

The air supply system provides steady hot, dry air for the experiment with an air blower, an air heater, a volumetric flow meter, a set of corrugated stainless steel jet tubing, an angle-adjust rack, and a nozzle. There are two platinum resistance temperature sensors on the air heater so the temperature can be controlled by the feedback signal. The jet tubing is wrapped by insulating materials to reduce the heat loss. The diameter of the nozzle is 25 mm and the height of the nozzle is 60 cm from the floor of the containment. The flow rate can be monitored by the volumetric flow meter and controlled by the valves.

Thermocouples are distributed in the test area inside the containment. They are connected to the Agilent multifunction switch/measure unit for data collection, which will be processed by the bundled software Bench Link Data Logger for further analysis. The distribution of thermocouple spots for the thermal stratification measurement is also shown in Figure 1. Details of the distribution are as follows:

(a) There are 7 layers (red lines) in the containment, and each elevation is given in Figure 1.

(b) The thermocouples are placed at the center of the cross section, at the midpoint of the radius, and 5 cm away from the inner wall, including the ones that adhere on the inner wall so there are 13 thermocouples on each layer in total.

(c) There is a thermocouple installed at the nozzle.

(d) The T-type thermocouples are used in the containment space and the K-type thermocouples are applied to obtain the wall temperature.

(e) There are five heat flow meters pasted on the inner wall (five black spots), and on the outer wall there are eight thermocouples (eight green spots), as shown in the sketch.

The parameters of the steel containment are listed in Table 2. It can be considered an even smaller modular reactor than the one listed in Table 1.

Table 2.
Parameters of the steel containment experiments
  Parameters   Values
Steel containment Height in total (m) Outside 3.34
    Inside 3.27
  Height of the dome (m) 0.57
  Height of the vertical body (m) 2.77
  Inner diameter (m) 1.98
  Thickness (m) 0.01
  Design pressure (MPa) 0.1
  Design temperature (℃) 20-100
Nozzle Diameter (mm) 25
Insulating layer Thickness (cm) 1
Working medium Air (m3/h) 0-30
  Temperature (℃) 0-100
Measuring points Points in the space 91
  Point at the nozzle 1
Show more

The experiments were carried out for non-adiabatic and adiabatic conditions. In non-adiabatic conditions, the steel containment was exposed to the air. In adiabatic conditions, the containment was covered with two layers of sponged rubber insulation sheets. In each condition, the experiments were performed at three different injection temperatures (Tin) with a constant flow rate (qv) of 15 m3/h, and two different flow rates (qv) with constant injection temperatures (Tin) of 100℃.Then, the comparison can be made by changing only one parameter to get a better understanding of the influence of each parameter. In the beginning of each test, the inlet temperature at the injection nozzle was controlled by setting the power of the air heater. Each individual experiment will last for more than 6 hours to reach a steady state, and then the data can be collected. Equipment in the containment, including thermal couples mounted on lines or on the wall, a tube and a rack, all occupied very little volume, so their impact on the data was limited. Besides, the data was collected in a steady state, so the influence of equipment in the containment was considered negligible for the stratification phenomena.

2.2 Visualized containment

In order to display the jet mixing and stratification development, the visualization experiments were carried out by using boron silicone glass containment. The experimental setup consists of two parts, the water supply control system and glass containment, as shown in Figure 2. The main function of the water supply control system is to provide and control the dyed water. The water tank is made of boron silicone glass and has 6 built-in heating rods. The dyed water must be heated to less than 80℃. A graduated cylinder is used to collect and buffer the dyed water, which is injected into the containment. The top and bottom of the graduated cylinder is sealed and its nominal capacity is 400~600 mL,with a minimum scale of 5 mL. There are three holes on the side of the graduated cylinder, which are distributed uniformly and can be connected to the flange. The middle hole is an inlet, the bottom hole is an outlet, and the top hole is a spare hole. The inner diameter of the graduated cylinder is 50 mm and the height is 300 mm. There is a joint at 50, 150, and 300 mm on the side of the graduated cylinder, respectively. Control valves are located both upstream and downstream of the graduated cylinder. A glass rotameter is used to control the volume flux.

Figure 2.
Visualized experimental setup.
pic

The transparent containment is made of boron silicone glass, which is full of saline water before conducting the experiments. There are holes on the glass containment body. The one at the top of the dome serves as the air vent to ensure that the injected water can fill with the containment. The other two are backup connectors located at the upper and middle vertical sections of the body. The glass containment body is supported by a steel pallet, and there is an inlet at the center and an outlet located nearby. A stainless steel nozzle is inserted in the inlet of the pallet to provide jet water. The parameters of the experimental setup are listed in Table 3.

Table 3.
Parameters of the experimental setup
Water tank   Glass containment  
Height 400 mm Total height 664 mm
Length 400 mm Height of the dome 114 mm
Width 400 mm Height of the cylindrical body 550 mm
Wall thickness 10 mm Outside diameter 400 mm
Diameter of the drain hole 30 mm Wall thickness 5 mm
Number of heating rods 6 Temperature 20℃
Power of heating rods 450 W×6 Volume flux of injected fluid 0~30 mL/s
Temperature 0—100℃ Temperature of injected fluid 10~80℃
Show more

3 Results

3.1 CFD model

CFD simulations were done to provide comparisons with the experimental results. Figure 3 shows the geometry of the CFD simulation. All parameters and boundary conditions in the models are consistent with the experiments. The ceiling dome and over ground cylindrical side wall are considered to be air cooled by natural convection under a constant room temperature. The underground wall and the floor are considered to be a constant temperature corresponding to the experimental conditions. The turbulence k-ε model and steady method were used in the calculations.

Figure 3.
Geometry for CFD simulations.
pic
3.2 Comparison

When equations (1) and (2) are applied in this experiment, Hsf is the inside height of the containment and dbjo is the diameter of the nozzle. The ambient fluid density,  ρa, is calculated using the average temperature in the containment. The source fluid density,  ρ0, is calculated by the jet temperature at the nozzle. The jet source speed, U0, can be calculated by the flow rate.

Table 4 lists some key parameters of the tests. The Richardson number and the criteria of eq. (1) are calculated to provide the stratification prediction for the corresponding conditions. All the conditions carried out in the experiments were stratified.

Table 4.
Experimental conditions and stratification prediction
Conditions Parameters Tin (℃) Flow Rate (m3/h) Average Temperature in Containment(℃) EXP CFD Stratification Prediction
        EXP CFD Ri(×10-4) Criteria Ri(×10-4) Criteria  
Non-adiabatic Hsf=3.27m 50 15 27.96 27.54 2.32 8.94 2.36 8.23 Stratified
  dbjo=25mm 80 15 31.98 31.28 4.62 10.30 4.69 10.34 Stratified
  αT=0.05 100 15 34.12 33.30 6.02 11.24 6.09 11.29 Stratified
    100 10 28.35 27.49 14.72 15.15 14.90 15.21 Stratified
    100 18 30.76 29.90 4.39 10.12 4.45 10.16 Stratified
Adiabatic   50 15 28.01 29.10 2.28 7.26 2.20 8.03 Stratified
    80 15 32.58 33.04 4.56 10.25 4.52 10.22 Stratified
    100 15 34.57 35.42 5.98 11.22 5.90 11.17 Stratified
    100 10 33.21 33.59 13.72 14.80 13.65 14.77 Stratified
    100 17.4 36.14 37.62 4.34 10.08 4.24 10.00 Stratified
Show more

The comparison between the experimental results and the calculation results are given in Figure 4 and Figure 5. The X-axis in the graph is the height in the containment, and the Y-axis represents the average temperature of the cross section in the corresponding height. In CFD simulation, the average temperatures of each layer are taken from the cross sections at 20 cm intervals from the bottom to the top. The curves of the simulation temperatures can be extended to the vertex while the experimental data are only available to a height of 315 cm due to the experimental conditions.

Figure 4.
Temperature distribution in the containment, Non-adiabatic atdifferent initial conditions.
pic
Figure 5.
The temperature distribution in the containment, Adiabatic atdifferent initial conditions.
pic

Figure 4 illustrates the temperature distribution comparisons in the containment both experimentally and numerically under non-adiabatic conditions. In the test conditions shown in (a), the flow rate is 15 m3/h, and inlet temperatures of 50℃, 80℃, and 100℃ are applied, respectively. It can be seen that the temperature distribution demonstrates stronger stratification with a higher injection temperature. This can be demonstrated by the CFD simulations, which reach a good agreement with the experiments. From the floor to some height after the nozzle, the temperatures climb up dramatically along the axial height with a large temperature gradient and, hence, cause severe thermal stratification. Then, the temperature increases slowly and becomes smoother from a height of 100 cm to 300 cm and maintains a relatively high temperature, which implies that most energy is stored in the upper part of the cylindrical body. In the upper part of the containment, the temperature increases again due to the shape of the dome, in which heat can be accumulated.

Figure 4(b) displays the temperature distribution with an inlet temperature of 100℃, and flow rates of 10 m3/h and 18 m3/h, respectively. It shows the comparison between different injection flow rates, where the stratification is more pronounced with a higher flow rate. The reasons will be discussed thoroughly through the dimensionless analysis of Figure 6.

Figure 6.
Comparison of dimensionless temperature distribution with dimensionless height for different conditions at low and high modes of flow rate respectively in experiment.
pic

Figure 5 indicates that the trend of the results under adiabatic conditions is in good agreement with the non-adiabatic conditions, i.e. the temperatures surge up in the area near the nozzle and then slightly ascend from the heights of 100 cm to 300 cm. They continue rising to reach the peak at the vertex of the dome. The simulations, however, show a higher temperature than the experiments as a whole. There are two reasons that may account for this. The main reason may lie in insufficient isolation from the sponged rubber sheet wrapping on the containment. The geometrical shape changes significantly at the connection part of ceiling wall and the cylindrical wall so it impairs the heat isolation of the insulating layer. This can explain why there is a slight mismatch between the CFD simulations and experiment that reach a maximum in the upper portion. In addition, there is an outlet at the bottom inside containment and air would exhaust through it, which may take away some energy. However, the temperature there was relatively low, so there was not supposed to be too much heat loss this way.

In our experiments, the increasing flow rate of the vertical injection causes more severe stratification in both conditions. Theoretically, by increasing flow rate, the thermal stratification will be weakened and the stratification will break down due to the enhancement of circulation flow, according to equations (1) and (2). Such phenomena cannot be observed in our experiments, but it can be demonstrated from Table 4 that Ri and the value of the criteria decrease with the increasing flow rate. This is different from the increasing inlet temperature. This can be explained by the insufficient power provided by the air blower. The flow rate is not large enough to reach the critical value to mix the stratification, but a higher flow rate does inject more energy and, as a result, increasing the inlet temperature causes similar thermal stratification to occur.

This point can be interpreted more clearly through the data in Figure 6, which is taken from the experiments. The results are in terms of the dimensionless temperature, (T-T1)/(T2-T1), with the dimensionless height, z/H, where T1 and T2 being, respectively, the temperatures at the bottom and the top of the containment, and H being the total height of the containment. Besides the slightly higher temperature under adiabatic conditions than under non-adiabatic conditions, the stratifications are all formed near the nozzle, and above the nozzle is the high temperature zone. The experiments carried out in the visualized containment will show such a phenomenon in section 3.3. We can deduce from Figure 4 to Figure 6 that the stratifications are formed because:

(a) The increasing flow rate and inlet temperature share the same principle, the increasing the hot air flow rates are actually increasing the temperature in the upper zone within our experimental conditions;

(b) In the containment, the cylindrical wall underground keeps the lower part a relatively low temperature, especially in adiabatic conditions, since most heat was conducted there, hence, accounting for the stable stratification;

(c) Due to (a) and (b), the stable stratified layers are built up and becomes strong enough to overcome mixing forces in these cases. In other words, the flow rates applied in the experiments cannot mix the temperature in such conditions.

In comparing the same conditions under non-adiabatic and adiabatic conditions, respectively, it can be concluded that the thermal stratification is more pronounced under adiabatic conditions rather than the condition that is exposed to the air. In other words, the wall exposed to the air influences the heat conduction in the containment, while the insulating layer decrease the heat loss shown in the comparison between Figure 4 to Figure 5. The cylindrical wall underground keeps the lower part of containment a relatively low temperature in all cases, hence, accounting for the thermal stratification. Therefore, the thermal stratification is the consequence of heat dissipation through the wall and the heat conduction underground when a LOCA scenario occurs.

3.3 Illustrations by the visualized containment

The experiments carried out by the visualization of the containment provided images of the buoyant jet development, and a transient simulation model using ANSYS CFX was done based on the experimental conditions. In the simulation, the dyed water was simplified to clear water, and the transient process lasted for 101 s. The simulation pictures were outputed every second. The meshing figure and cross-sectional views are shown in Figure 7.

Figure 7.
CFD modeling and simulation for visualized containment.
pic

The initial temperature of the containment is 15℃. The volumetric flow rate of the injected water is 50 L/h and the temperature is 60℃. The density of the saline water, which fills with the containment, is 1010 kg/m3. Therefore, the dyed water in the experiment corresponds to the hot water in the simulation.

The comparisons between the experiment and simulation are illustrated from Figure 8 to Figure 13 and show good agreement. The development of the buoyant jets in the containment can be sufficiently described by three stages. At the initial stage, the hot water came from the buoyant jet with a lower density raised by the momentum at the inlet nozzle, then raised to the vertex rapidly (Figure 8 to Figure 9). Then, in the second stage, called developing stage, where the buoyant jet spreads along the inner wall from the top and forms a mushroom shape due to the features of the dome. It gradually spreads out in the space of the containment, and a circular vortex is formed between the jet water and the wall (Figure 10 to Figure 11). Meanwhile, an interface appeared due to the density difference of the jet water and the ambient water in this stage. As time went on, the interface headed downward gradually until it entered the stable stratified stage (Figure 12 to Figure 13), and the stratified interface was finally maintained slightly above the nozzle (Figure 13).This result, in turn, meets the stratification location of the analysis in Figure 6. The experimental results and the simulation show a good agreement in the first stage (jetting stage) and the third stage (stratified stage). In the second stage, as the developing stage, the dyed water diffused a little bit faster than the simulation, but qualitatively, the visualized experiments combined with the numerical simulation illustrated the development of buoyant jets in the containment.

Figure 8.
The 1st second of the experiment and the 1st second of the simulation.
pic
Figure 9.
The 10thsecond of the experiment and the 10thsecond of the simulation.
pic
Figure 10.
The 30th second of the experiment and the 30th second of the simulation.
pic
Figure 11.
The 50th second of the experiment and the 50thsecond of the simulation.
pic
Figure 12.
The 70th second of the experiment and the 70thsecond of the simulation.
pic
Figure 13.
The 100th second of the experiment and the 100th second of the simulation.
pic

4 Conclusion

In this paper, small steel containment experiments with hot air injection were conducted to study the mixing and stratification phenomena. The experimental data taken from the stainless steel containment and numerical results were compared to show the temperature distribution. Two key parameters, injection temperature and flow rate, are taken into consideration, and the experiments are conducted under non-adiabatic and adiabatic conditions. The conclusions can be drawn as follows:

1) Higher inlet temperature can cause a more severe thermal stratification within our experimental conditions. It can be demonstrated by three cases, shown in Figure 4 (a) and Figure 5 (a), where the temperature gradient becomes larger as the injection temperature is increased. It also demonstrates that the initial phase of the LOCA is the most dangerous period due to the serious stratification phenomenon caused by the high injection temperature of the break. The thermal stratification will reduce the heat transfer efficiency. But with the cooling of the containment, the temperature will drop down and the stratification will be weakened.

2) Under the critical value, higher inlet flow rate can also enhance the thermal stratification. With the increasing flow rate, the heat is accumulated and the temperature rises in the upper part. A large flow rate experiment will be carried out in the near future.

3) The thermal stratification is more pronounced under adiabatic conditions rather than non-adiabatic.

To visualize the injection and stratification development in the containment, experiments using a small glass containment and dyed water were carried out. With a comparison to the numerical simulation, it provided a better understanding of the transient process. After the buoyant jet came out from nozzle, it rose to the vertex rapidly and formed a mushroom shape due to the features of the dome. Then a circular vortex formed between the jet water and the wall and, meanwhile, an interface appeared. The interface headed downward gradually and was finally maintained slightly above the nozzle, hence, forming a stable stratification.

Temperature distributions from the steel containment showed that the temperature gradient of the height under the nozzle was larger than that above the nozzle. The visualized experiments also showed that the stratification occurred near the nozzle. Both of the results, therefore, imply that the height of the nozzle is another key parameter which can influence thermal stratification. We will focus on that parameter in the future works.

REFERENCES
[1] International Atomic Energy Agency, Natural Circulation Phenomena and Modelling for Advanced Water Cooled Reactors (IAEA, Vienna, 2012), p. 105
[2] H.H. Zhao, P.F. Peterson,

An Overview of Modeling Methods for Thermal Mixing and Stratification

. The 8th International Topical Meeting on Nuclear Thermal-Hydraulics, Operation and Safety (NUTHOS-8), Shanghai, China, October 10-14, 2010
Baidu ScholarGoogle Scholar
[3] R.E. Gamble, T.T. Nguyen, B.S. Shiralkar et al.,

Pressure suppression pool mixing in passive advanced BWR plants

. Nucl. Eng. Des. 204, 321-336 (2001). doi: 10.1016/S0029-5493(00)00363-0
Baidu ScholarGoogle Scholar
[4] H. Zhao, L. Zou, Zhang H.

Simulation of thermal stratification in BWR suppression pools with one dimensional modeling method

. Ann. Nucl. Energy 63, 533-540 (2014). doi: 10.1016/j.anucene.2013.08.031
Baidu ScholarGoogle Scholar
[5] T.L. Norman, S.T. Revankar,

Jet-plume condensation of steam- air mixtures insubcooled water, Part 1: experiments

. Nucl. Eng. Des. 240, 524-532 (2010). doi:10.1016/j.nucengdes.2009.09.019
Baidu ScholarGoogle Scholar
[6] J. Laine, M. Puustinen, Thermal Stratification Experiments with the Condensation Pool Test Rig. Finland, Lappeenranta University of Technology, NKS-117, 2006
[7] P.F. Peterson,

Scaling and analysis of mixing in large stratified volumes

. Int. J. Heat Mass Trans. 37, 97-106 (1994). doi:10.1016/0017-9310(94)90013-2
Baidu ScholarGoogle Scholar
[8] C. Lin, Z. Yu, Y. Ou, Passive and Advanced Pressurized Water Reactor Nuclear Power Technology (Atomic Energy Press, Beijing, 2010), p. 482
[9] International Atom Energy Agency.

Status of Small and Medium Sized Reactor Designs

. https://www.iaea.org/NuclearPower/Downloadable/SMR/files/smr-status-sep-2012.pdf
Baidu ScholarGoogle Scholar
[10] International Atom Energy Agency.

NuScale Power Modular and Scalable Reactor

. https://aris.iaea.org/PDF/NuScale.pdf
Baidu ScholarGoogle Scholar
[11] Tatiana Antysheva.

SVBR-100: New Generation Power Plants for Small and Medium Size Power Applications

. https://www.iaea.org/NuclearPower/Downloadable/Meetings/2011/2011-07-04-07-08-WS-NPTD/2_RUSSIA_SVBR_AKME-eng_Antysheva.pdf
Baidu ScholarGoogle Scholar
[12] International Atom Energy Agency.

Status report 66—VBER-300

. https://aris.iaea.org/sites/.%5CPDF%5CVBER-300.pdf
Baidu ScholarGoogle Scholar
[13] R. Mazzi.

Carem: An Innovative-integrated PWR

. 18th International Conference on Structural Mechanics in Reactor Technology, Beijing, China, August 7-12, 2005
Baidu ScholarGoogle Scholar
[14] Matt Ales.

Reactor Design Overview

. http://www.uxc.com/smr/Library/Design%20Specific/mPower/Presentations/2012%20-%20Reactor%20Design%20Overview.pdf
Baidu ScholarGoogle Scholar
[15] B.R. Upadhyaya, C. Mehta, et al.

Approaches to Process Monitoring in Small Modular Reactors

. Proceedings of the ASME 2014 Small Modular Reactors Symposium, Washington, DC, USA, 2014
Baidu ScholarGoogle Scholar