Introduction
Simulations of nuclear-reactor circuit systems provide a versatile and intuitive tool to understand their operating mechanisms and performances, including their thermal-hydraulic properties [1-3], safety features [4, 5], and operational efficiency [6]. To analyze the detailed thermal-hydraulic behavior in the reactor-core circuit, the computational fluid dynamics (CFD) [7] method has been adopted in pressurized-water reactors (PWRs) [8] and boiling-water reactors [9], which solves the Navier–Stokes equations, large-eddy simulation equations, and Reynolds-averaged Navier–Stokes equations to simulate the flow and heat-transfer phenomena inside the circuit in detail. However, because the simulation of a nuclear-reactor circuit system involves large-scale multiphysics coupling calculations, real-time and ultra-real-time CFD simulations are challenging [10]. For the simulation of a full-circuit system, a coarse-mesh average distribution is often sufficient for engineering requirements. From this perspective, various simplified methods have been developed, including lumped parameter models [11], subchannel analysis [12], nodal methods [13], and the one-dimensional finite difference method (FDM) [14]. These studies have effectively supported the steady-state analysis [15], transient response [16, 17], and controller design [18] of Generation III reactor systems [19]. Simultaneously, they have provided significant assistance in the thermal-hydraulic analysis [20, 21], core design, and economic-benefit assessment [22] of Generation IV reactor systems [23].
Although these system programs can quickly achieve an accurate simulation of the entire circuit system, their computational efficiency still requires further improvement for real-time simulations, and especially for ultra-real-time simulations, which is particularly important for digital twins and data assimilation [24, 25]. In these applications, numerical simulations need to be synchronized with actual operating conditions or should be faster than actual operating conditions to simulate physical phenomena inside the circuit so that operators can carry out their corresponding operations. Therefore, for both commercial nuclear power plants and research reactors, establishing an accurate real-time or ultra-real-time simulation platform [23] is of paramount importance. In particular, when applied to power-plant condition monitoring and fault diagnosis, an efficient and accurate numerical method can provide data support for the data-assimilation process [24]. Therefore, reduced-order methods (ROMs) [26-29] may be an efficient approach.
The ROM constitutes an approach that can significantly reduce computational-resource requirements and processing time. This target is achieved by generating snapshots from preexisting data, extracting data characteristics from these snapshots, and subsequently employing these data characteristics to make accurate and efficient predictions of unknown conditions. Several ROMs have been proposed and applied to various engineering disciplines. The Routh approximation method was proposed in the field of order reduction for linear time-invariant systems [30]. Reduced-order modeling of large-scale dynamical systems has been achieved with the application of Krylov subspace methods [31]. Furthermore, linear-system reduction techniques such as the Pade approximation method have been employed, as demonstrated by [32]. Moreover, proper orthogonal decomposition (POD) has garnered widespread use in fluid-dynamic analyses [33] and the characterization of coherent features within fluid flows [34]. Sartori et al. [35] employed POD theory to investigate the single-channel model of a lead-cooled fast reactor. It also demonstrated efficacy in solving parameterized nonlinear partial differential equations [36], yielding commendable results in various applications.
This study aims to establish a POD ROM for modeling reactor primary-circuit systems to accelerate thermal-hydraulic calculations and maintain accuracy. The remainder of this paper is organized as follows. In Sect. 2, the theoretical foundations of the primary-circuit full-order model theory, POD reduced-order theory, and circuit reduced-order model theory are explained. In Sect. 3, we present the outcomes of computations conducted with both full- and reduced-order models under diverse operating conditions. These results were subjected to meticulous analysis and discussion, encompassing an examination of the relative errors between them. Finally, Sect. 4 presents the concluding remarks of this study.
Modeling theory
Full-order model
This section introduces the modeling theory of a one-dimensional steady-state thermal-system circuit. An AP1000 reactor [37] is used as an example. This intricate circuit comprises essential components, namely, the reactor core, pressurizer, steam generator, and pump, all interconnected through a pipe-model representation. Notably, the pressurizer serves as a pressure-containment boundary within this circuit, maintaining a constant pressure of 15.5 MPa [38]. The main pump model is based on the nuclear main-pump head normalization curve provided by Zhu et al. [39], where the rated head of the AP1000 reactor coolant pump is 111.3 m, and the design mass flow rate is 17886 (m3/h). The spatial arrangement of these critical components and interconnecting pipes is shown in Fig.1 for reference.
-202411/1001-8042-35-11-003/alternativeImage/1001-8042-35-11-003-F001.jpg)
The proposed method is divided into two stages: offline and online. For the offline stage, the one-dimensional FDM for solving system thermal-hydraulic equations is first employed as a full-order numerical method to model the thermal-hydraulic behavior of components within the primary-circuit system. A series of typical samples of the circuit calculation results under different states from the full-order model were chosen as snapshots to generate a reduced-order model of the primary thermal system circuit (RO-PTSC). In the online stage, the proposed reduced-order model can be used to simulate the thermal-hydraulic properties with significant acceleration and high accuracy. The intricate process of constructing this reduced-order model is illustrated in Fig.2, providing a visual representation of our methodology.
-202411/1001-8042-35-11-003/alternativeImage/1001-8042-35-11-003-F002.jpg)
Core
The reactor core generates heat via nuclear fission, which is the primary energy source for nuclear power systems. The thermal-hydraulic process for the reactor core is characterized by a single-channel model [40], which describes a one-dimensional distribution along the axial direction. The governing equations of this model include the continuity, momentum, and energy equations, as follows:
The form resistance pressure-drop coefficient Ksp is expressed as follows [41]:
Pipe
Two types of distinct pipe models are introduced in this context. The first model is the discrete pressure-drop pipe model, which does not consider heat transfer. Conversely, the second model is the discrete heat-transfer pipe model, which explicitly accounts for heat-transfer effects [42]. The governing equations for the discrete pressure-drop pipe model include the continuity and momentum equations, which are given by the following equations:
In the discrete heat-transfer pipe model, an energy equation must be incorporated into the continuity and momentum equations. The energy equations can be expressed as follows:
Steam generator
The steam generator is the interface between the primary and second circuits, where the coolant of the primary circuit flows through the U-tube and transfers heat to the coolant of the second circuit. To simulate this equipment, the one-dimensional dynamic mathematical model proposed by Zhang et al. [44] is adopted, which introduces a distributed parameter model tailored for steam generators in PWRs. As shown in Fig.3, T1,f and T1,ad represent the fluid temperature of the primary side in the parallel- and counter-flow sections, respectively. Tw,in,f and Tw,out,f represent the inner and outer wall temperatures of the U-tube in the parallel-flow section, respectively. Tw,in,ad and Tw,out,ad represent the inner and outer wall temperatures of the U-tube in the counter-flow section, respectively. T2,r and T2,s represent the preheating and boiling section temperatures of the secondary side fluid, respectively. T1,in represents the inlet temperature on the primary side, T2,in is the inlet temperature on the secondary side, and T1,out represents the outlet temperature on the primary side.
-202411/1001-8042-35-11-003/alternativeImage/1001-8042-35-11-003-F003.jpg)
The steady-state heat dynamic balance equations of the parallel and counter flows of the primary side can be written as follows:
The heat-balance equations of the U-tube inner wall for parallel and counter flows can be written as:
The heat-balance equations of the outer wall of the U-tube for the parallel flow in the preheating and boiling sections can be written as follows:
In a natural-circulation steam generator, the heat-transfer process in the preheating section of the secondary side undergoes supercooled boiling. Therefore, the convection heat-transfer coefficient in the preheating section can be uniformly treated by following the approach used for the boiling section and the principles of boiling heat transfer. According to Rohsenow [46], the heat-transfer coefficient can be described as follows:
The heat-balance equations of the preheating and boiling sections of the secondary side are as follows:
Connection between different components
The primary circuit comprises the above components all connected. Mathematically, this connection is achieved by transferring physical quantities between different component equations, as shown in Fig. 4.
-202411/1001-8042-35-11-003/alternativeImage/1001-8042-35-11-003-F004.jpg)
In Fig. 4, the inlet and outlet of each component are represented by dashed-line and dashed-dotted-line boxes, respectively. Along the yellow arrows, the output variables of each component, including temperature, pressure, and velocity, are transferred to the next component as input values. For example, the output temperature, pressure, and velocity of the core are transferred to heat section 1 as inlet variables.
Numerical computation methods of the full-order model
The AP1000 primary-circuit model constructed in this study comprises a core, heat section, pressurizer, steam generator, pressure-drop pipe, cold section, and pump, as shown in Fig.1. Among these components, the physical field distributions in the core, heat section, steam generator, pressure-drop pipe, and cold section are computed using the FDM. The calculated area is discretized using a uniform mesh configuration along the coolant-flow direction. The physical properties of the coolant, including thermal conductivity, density, kinematic viscosity, and specific heat capacity, are defined as functions of temperature and pressure based on the IF97 thermophysical property table [47].
In the governing equations of the pipes and cores, the convective terms are discretized using the first-order upwind scheme [48], whereas the diffusion terms are discretized using a central differencing scheme [49]. The discretization formats of the governing equations for both the pipes and reactor core can be expressed as follows, where i represents the index of the internal grids and the subscripts ‘p’ and ‘c’ represent the pipe and core, respectively. Continuity equation:
For the governing equations of the steam generator, the discretization of Eq. (11), (12), (21), and (22) employs a first-order upwind scheme, which is written as follows:
Reduced-order model
Snapshots and POD bases
As previously mentioned, the construction of the reduced-order model involves offline and online stages, as shown in Fig. 5. In the offline stage, the full-order model is used to generate snapshots under various typical operating conditions. In the online stage, the reduced-order model is applied to predict the distributions of temperature, velocity, and pressure for unknown states. The first step in the offline stage is to calculate snapshots of different operation states, including the temperature, pressure, and velocity distributions. During this process, N operating states are selected within the operating range, and snapshots of these states are calculated using the full-order model (FOM) described in Sect. 2.1. The snapshot set generated in the offline stage based on the FOM is expressed as follows:
-202411/1001-8042-35-11-003/alternativeImage/1001-8042-35-11-003-F005.jpg)
The main modes of the system variation, namely the POD bases, can be extracted based on a thorough analysis and processing of these snapshots, thereby realizing the objective of reducing the system intricacy to a lower-dimensional representation. POD bases consist of a set of modes or functions employed to characterize the predominant characteristics of variation within a system. These modes are extracted from snapshots and are conventionally represented as feature vectors. According to POD theory [51], the POD basis can be obtained from the eigenvalues and eigenvectors of the matrix S, which can be expressed as follows:
As the governing equations and snapshots generated by the full-order model for each component are different, special POD bases are formed for different components. In addition, for the same type of pipe, although the governing equations remain the same, different positions in the circuit result in different diameters and lengths. Consequently, their snapshots and generated POD bases are different.
According to POD theory, the significance of eigenvectors is directly associated with the magnitude of their corresponding eigenvalues. To determine the POD order Np for reduced-order modeling, the following criteria can be used:
The operating conditions to be predicted, including the inlet temperature, inlet flow rate, and core power, can be introduced by imposing the corresponding data at node i. For example, the flow rate of inlet node i is set to the given inlet value.
When using the LSM, the POD coefficient cm can be determined by minimizing the residual equation, which converts Eq. (41) to the following equations:
Primary circuit POD reduced-order model
To construct an entire reduced-order model for the primary circuit of the reactor, reduced-order models of the primary components within the circuit must first be established. For the AP1000 primary circuit considered in this study, the components primarily encompassed the core, pipes, and steam generator. Other components with relatively minor effects on computational efficiency, such as regulators and pumps, do not require the development of reduced-order models. The spatial arrangement of AP1000 primary-circuit components along the coolant traversing distance is shown in Fig. 6.
-202411/1001-8042-35-11-003/alternativeImage/1001-8042-35-11-003-F006.jpg)
As previously mentioned, the establishment of the reduced-order model is based on a collection of snapshots. In this section, the variables used to construct the reduced-order model are systematically selected based on the governing equations associated with the various components within the circuit. In the following sections, the selection of snapshots and establishment of the ROM for different components are introduced. To simplify the subsequent construction of the residual function and compute the coefficient matrix of the POD bases, the functions differentiated within the governing equations are chosen as the variables.
Core
Table 1 lists the selected variables contained in the snapshots, POD bases, and boundary conditions for the reactor core, which are determined from the single-channel thermal-hydraulic equations (1)-(3) and their discretization. The core variables can be reconstructed using the POD bases according to
Core Variables | POD Basis | Inlet boundary conditions | Outlet boundary conditions |
---|---|---|---|
- | |||
- | |||
p | dp/dx=0 | ||
ρ | - | ||
- | |||
pu | - | ||
For each internal node, by applying Eq.(43) into Eqs.(24), (26), and (28) and calculating the summation of the residue over all internal nodes, the accumulated residual functions of the core are generated as Eqs. (45), (49), and (46), where superscripts ‘con’, ‘mom’, and ‘ene’ represent continuity, momentum, and energy, respectively, and NXc represents the number of discretized grid points in the reactor core.
Similarly, to satisfy the inlet and outlet boundaries of the eight selected variables, the following two residual functions for the inlet and outlet are defined in Eqs. (47) and (48), respectively. Then, the total residual function of the core can be written as
Steam generator
The variable selections based on the steam generator (SG) governing equations are shown in Table 2. Because the temperature at the last grid point of the parallel-flow section is transferred to the last grid point of the counter-flow section, the inlet boundary condition for the coolant flow on the primary side is set as T1,f,1=T1,0. The outlet boundary condition of the coolant flow on the primary side is considered to be a fully developed boundary, indicating a temperature gradient of zero at the outlet of the counter-flow section.
Core Variables | POD Basis | Inlet Boundary Conditions | Outlet Boundary Conditions |
---|---|---|---|
T1,f | - | ||
T1,ad | - | dT/dx=0 | |
Tw,in,f | - | - | |
Tw,out,f | - | - | |
Tw,in,ad | - | - | |
Tw,out,ad | - | - | |
T2,r | dT/dx=0 | ||
- | - | ||
- | - | ||
hs | - | - | |
- | - |
Similar to Eq.(43), the SG variables can be reconstructed using the POD bases according to
Let the number of discrete meshes in the parallel-flow and counter-flow sections of the SG U-tube be defined as
Pipes
The selected variables and boundary conditions for the pipes are presented in Table 3. The thermal-hydraulic variables of the pipe can be reconstructed using the following equation:
Core variables | POD basis | Inlet boundary conditions | Outlet boundary conditions |
---|---|---|---|
- | |||
- | |||
p | dp/dx=0 | ||
ρ | - | ||
- | |||
Based on the boundary conditions listed in Table 3, the residual functions at the inlet boundary of the pressure-drop pipes can be expressed using Eq. (61).
The governing equations for the outlet boundary grids of the pressure-drop pipes can be discretized as follows:
Implementation of ROMs
After establishing a reduced-order model for each component, the implementation of the proposed ROMs for predicting different working conditions is listed as follows:
Offline stage:
Step 1: The key state parameters for different equipment and circuit systems such as inlet flow rate, inlet temperature, and core power are determined.
Step 2: The ranges of different state parameters are determined based on the actual operating conditions and a series of state points are selected within these ranges for a full-order simulation to obtain snapshots.
Step 3: The POD process is immposed according to Section 2.2.1 to obtain the POD bases and a reduced-order model is built for the primary circuit.
Online stage:
Step 1: The state parameters are set for the state to be predicted.
Step 2: The POD coefficients cm are calculated for each component using Eqs. (44), (57), (64), (69), and (42).
Step 3: The temperature, pressure, and velocity distributions of the unknown state are reconstructed using Eqs. (43), (50), and (58).
Results and Discussion
This section presents a validation of the full-order model of the primary-circuit system against the designated parameters. After validation, the results obtained from the reduced-order model corresponding to the primary circuit of the reactor are analyzed and compared. The reduced-order model is validated in two parts. First, the results of the reduced-order model are compared with those of the FOM corresponding to the operating conditions included in the snapshots. The second part uses the reduced-order model to calculate the unknown operating conditions not included in the snapshots and compares them with the full-order results. In this study, 11 specific operating conditions covering a 70%–100% rating power are chosen as snapshots. Under different operating conditions, the core power and boundary conditions on the secondary side of the steam generator are considered as the input parameters. These selected snapshots serve as the foundation for generating POD bases and establishing the corresponding POD reduced-order models. Following Eq.(39), the values of Np for each component of the primary circuit are listed in Table 4.
Component | Np |
---|---|
Core | 2 |
Heat pipe section 1 | 3 |
Heat pipe section 2 | 3 |
Steam generator | 2 |
Pressure drop pipe | 3 |
Cold pipe section | 3 |
Full-order model verification
First, the simulation results of the primary-circuit FOM developed in this study are compared with the design parameters of the AP1000 reactor [38]. Tables 5 and 6 compare the full-power operating conditions with the full-order calculation results for the AP1000 reactor core and SG sections, respectively. These comparisons indicate that the proposed FOM can be used to accurately simulate the thermal-hydraulic behavior of the system circuit. For verification, snapshots are obtained using the proposed FOM.
Design value | Simulation value | Errors (%) | |
---|---|---|---|
Inlet temperature (K) | 552.59 | 552.554 | 0.006515 |
Outlet temperature (K) | 597.81 | 597.306 | 0.084308 |
Average flow rate (m/s) | 4.816 | 4.81658 | 0.012043 |
Core pressure drop (MPa) | 0.43 | 0.43512 | 1.190698 |
Design value | Simulation value | Errors (%) | |
---|---|---|---|
Primary side inlet temperature (K) | 594.25 | 594.212 | 0.006395 |
Primary side outlet temperature (K) | 553.75 | 553.664 | 0.01553 |
Reduced-order model verification
Second, the full-power operating conditions of the snapshot set are used to verify the correctness of the reduced-order model. Figure 7 provides the temperature distribution of the reduced-order model and FOM, which is compared with those of the design values. This comparison shows that both the FOM and reduced-order model solutions are in good agreement with the design values. The reduced-order model accurately predicts the physical distribution of a circuit system.
-202411/1001-8042-35-11-003/alternativeImage/1001-8042-35-11-003-F007.jpg)
Figure 7 also further provides the relative error between the reduced-order model and FOM solutions. The maximum relative error between the results of the reduced-order model and the FOM is 0.233% for the primary circuit of the entire reactor. Compared with the design parameters, the relative deviation of the core inlet temperature calculated by the reduced-order model is 0.147%, that of the core exit temperature is 0.018%, that of the SG inlet temperature is 0.145%, and that of the SG outlet temperature is 0.006%, indicating that the proposed reduced-order model can accurately predict the thermal-hydraulic behavior of the system circuit.
Under full-power operating conditions, the computational time for the FOM and reduced-order model are 63.9 s and 0.0404 s, respectively. The reduced-order model shows an acceleration ratio of 1,582 times, indicating that the proposed reduced-order model can effectively improve the computing speed. The FOM computing time of our self-developed code is similar to those of other widely used programs with the same accuracy. This further indicates that the proposed reduced-order model can be used to effectively accelerate reactor-circuit calculations. In digital-twin applications, the efficiency of current reduced-order models can strongly support real-time or even ultra-real-time assimilation simulations.
Finally, to verify the extensibility and accuracy of the reduced-order model, 86.5% and 74.5% power operating conditions, which are not included in the snapshots, are calculated using the reduced-order model. The temperature, pressure, and velocity fields obtained from the reduced-order model are compared with the FOM results. Under operating conditions of 74.5% and 86.5% power, the acceleration ratios of the reduced-order model are 1768 and 1493 times, respectively.
Figure 8 illustrates the temperature distribution along the flow direction calculated by both the full-order and reduced-order models under operating conditions of 86.5% and 74.5% power. In addition, the relative errors at various grid points are presented. The reduced-order results agreed well with the full-order results. For the 86.5% power operating condition, the maximum relative error is 0.117% and the average error is 0.015%; for the 74.5% power operating condition, the maximum relative error is 0.063% and the average error is 0.019%.
-202411/1001-8042-35-11-003/alternativeImage/1001-8042-35-11-003-F008.jpg)
The pressure-field distributions for both the 86.5% and 74.5% full-power operating conditions and the relative errors between the reduced- and full-order results are illustrated in Fig. 9. Under the 86.5% operating condition, the maximum relative error is 0.012%, and the average error is 0.006%; under the 74.5% operating condition, the maximum relative error is 0.01% and the average error is 0.007%.
-202411/1001-8042-35-11-003/alternativeImage/1001-8042-35-11-003-F009.jpg)
Figure 10 presents the velocity-field distributions for both the 86.5% and 74.5% power operating conditions. The reduced-order results are in good agreement with the full-order results. Under the 86.5% operating condition, the maximum relative error is 0.024% and the average error is 0.008%; under the 74.5% operating condition, the maximum relative error is 0.026% and the average error is 0.02%. These solutions indicate that the proposed reduced-order model can accurately predict the physical processes of unknown states.
-202411/1001-8042-35-11-003/alternativeImage/1001-8042-35-11-003-F010.jpg)
Conclusions
In this study, an RO-PTSC based on POD and LSM was developed. A full-order circuit model is used to generate snapshots under various operating conditions. The POD bases were subsequently constructed from the snapshots. Then, for all main circuit components, the residual functions were constructed based on their governing equations, and the POD coefficient matrices were calculated using the LSM. The variables of interest in the governing equations could be reconstructed using these coefficient matrices and the POD basis. Subsequently, the temperature, pressure, and velocity fields were obtained by employing the combinatorial relationships between the reconstructed variables and invoking the IF97 physical property calculation function. To verify the proposed RO-PTSC method, the AP1000 primary circuit was simulated using a reduced-order model and compared with the full-order results. For 86.5% and 74.5% power operating conditions, the temperature, pressure, and velocity fields along the flow direction from the reduced-order model agreed well with the full-order results and attained an acceleration ratio exceeding 1000.
Thermal hydraulic phenomena related to small break LOCAs in AP1000
. Prog. Nucl. Energ. 53, 407-419 (2011). https://doi.org/10.1016/j.pnucene.2011.02.007.Numerical analysis of thermal-hydraulic behavior of supercritical water in vertical upward/downward flow channels
. Nucl. Sci. Tech. 19, 178-186 (2008). https://doi.org/10.1016/j.pnucene.2011.02.007.Steady thermal hydraulic analysis for a molten salt reactor
. Nucl. Sci. Tech. 19, 187-192 (2008). https://doi.org/10.1016/S1001-8042(08)60048-2.Prediction of flow and temperature distributions in a high flux research reactor using the porous media approach
. Sci Technol. Nucl. Ins. 2017,Development of a safety analysis code for molten salt reactors
. Nucl. Eng. Des. 239, 2778-2785 (2009). https://doi.org/10.1016/j.nucengdes.2009.08.020.Developing structural, high-heat flux and plasma facing materials for a near-term DEMO fusion power plant: The EU assessment
. J. Nucl. Mater. 455, 277-291 (2014). https://doi.org/10.1016/j.jnucmat.2014.06.014.Models, algorithms and validation for opensource DEM and CFD–DEM
. Prog. Comput. Fluid Dy. 12, 140-152 (2012). https://doi.org/10.1504/pcfd.2012.047457.CFD simulation of a four-loop PWR at asymmetric operation conditions
. Nucl. Eng. Des. 300, 591-600 (2016). https://doi.org/10.1016/j.nucengdes.2016.02.018.Modelling and analysis of the MSFR transient behaviour
. Ann. Nucl. Energy 64, 485-498 (2014). https://doi.org/10.1016/j.anucene.2013.08.003.Assessment of reactor flow field prediction based on deep learning and model reduction
. Ann. Nucl. Energy. 179,A lumped-parameter dynamical model of a nuclear heating reactor cogeneration plant
. Energy 145, 638-656(2018). https://doi.org/10.1016/j.energy.2017.12.153.FIDAS: Detailed subchannel analysis code based on the three-fluid and three-field model
. Nucl. Eng. Des. 120, 147-161 (1990). https://doi.org/10.1016/0029-5493(90)90369-9.Review on neutronic/thermal-hydraulic coupling simulation methods for nuclear reactor analysis
. Ann. Nucl. Energy 137,Stability analysis of BWR nuclear-coupled thermal-hydraulics using a simple model
. Nucl. Eng. Des. 177, 155-177 (1997). https://doi.org/10.1016/S0029-5493(97)00192-1.The physical closure laws in the CATHARE code
. Nucl. Eng. Des. 124, 229-245 (1990). https://doi.org/10.1016/0029-5493(90)90294-8.An object-oriented approach to simulation of IRIS dynamic response
. Prog. Nucl. Energy 53, 48-58 (2011). https://doi.org/10.1016/j.pnucene.2010.09.004.A simple dynamic model of the primary circuit in VVER plants for controller design purposes
. Nucl. Eng. Des. 237, 1071-1087 (2007). https://doi.org/10.1016/j.nucengdes.2006.12.002.Demonstration of the Advanced Dynamic System Modeling Tool TRANSFORM in a Molten Salt Reactor Application via a Model of the Molten Salt Demonstration Reactor
. Nucl. Technol. 206, 478-504 (2020). https://doi.org/10.1080/00295450.2019.1627124.Thermal hydraulic analysis of a pebble-bed modular high temperature gas-cooled reactor with ATTICA3D and THERMIX codes
. Nucl. Eng. Des. 246, 286-297 (2012). https://doi.org/10.1016/j.nucengdes.2012.02.014.HYLIFE-II: A molten-salt inertial fusion energy power plant design
. Fusion Technol. 25, 5-25 (1994). https://doi.org/10.13182/FST94-A30234.Real-time prediction of nuclear power plant parameter trends following operator actions
. Expert Syst. Appl. 186,Data assimilation concepts and methods March 1999, Meteorological Training Course Lecture Series
. ECMWF 718, 59 (2002).An efficient digital twin based on machine learning SVD autoencoder and generalised latent assimilation for nuclear reactor physics
. Ann. Nucl. Energy. 179,Reduced-order modeling: new approaches for computational physics
. Prog. Aerosp. Sci. 40, 51-117 (2004). https://doi.org/10.1016/j.paerosci.2003.12.001.Physics-Constrained neural network for solving discontinuous interface K-eigenvalue problem with application to reactor physics
. Nucl. Sci. Tech. 34, 161 (2023). https://doi.org/10.1007/s41365-023-01313-0.Ensemble Bayesian method for parameter distribution inference: application to reactor physics
. Nucl. Sci. Tech. 34, 199 (2023). https://doi.org/10.1007/s41365-023-01356-3.conducted an uncertainty analysis of a data-enabled physics-informed neural network to solve the neutron diffusion eigenvalue problem
. Nucl. Sci. Eng. (2023) https://doi.org/10.1080/00295639.2023.2236840.Routh approximations for reducing order of linear, time-invariant systems
. IEEE Trans. Automat. Contr. 20, 329-337 (1975). https://doi.org/10.1109/TAC.1975.1100953.Krylov subspace techniques for reduced-order modeling of large-scale dynamical systems
. Appl. Numer. Math. 43, 9-44 (2002). https://doi.org/10.1016/s0168-9274(02)00116-2.Linear system reduction using Pade approximation to allow retention of dominant modes
. Int. J. Control 21, 257-272 (1975). https://doi.org/10.1080/00207177508921985.Finite volume POD-Galerkin stabilised reduced order methods for the parametrised incompressible Navier-Stokes equations
. Comput. Fluids 173, 273-284 (2018). https://doi.org/10.1016/j.compfluid.2018.01.035.Dynamic mode decomposition of numerical and experimental data
. J. Fluid Mech. 656, 5-28 (2010). https://doi.org/10.1017/S0022112010001217.A multi-physics reduced order model for the analysis of Lead Fast Reactor single channel
. Ann. Nucl. Energy 87, 198-208 (2016). https://doi.org/10.1016/j.anucene.2015.09.002.Nonlinear Model Reduction Via Discrete Empirical Interpolation
. SIAM J. Sci. Comput. 32, 2737-2764 (2010). https://doi.org/10.1137/090766498.Westinghouse AP1000 advanced passive plant
. Nucl. Eng. Des. 236, 1547-1557 (2006). https://doi.org/10.1016/j.nucengdes.2006.03.049.The AP1000TM reactor: passive safety and modular design
. Energy Proced. 7, 293-302 (2011). https://doi.org/10.1016/j.egypro.2011.06.038.The research on AP1000 nuclear main pumps’ complete characteristics and the normalization method
. Ann. Nucl. Energy 99, 1-8 (2017). https://doi.org/10.1016/j.anucene.2016.08.014.A Generalized computational model for transient two phase thermal hydraulics in a single channel
.Proposal for pressure drop prediction for a fuel bundle with grid spacers using Rehme pressure drop correlations
. Nucl. Eng. Des. 240, 1830-1842 (2010). https://doi.org/10.1016/j.nucengdes.2010.03.039.Dynamic heat transfer performance study of steam generator based on distributed parameter method
. Ann. Nucl. Energy 63, 658-664 (2014). https://doi.org/10.1016/j.anucene.2013.09.005.Experimental investigation of convective heat transfer in a narrow rectangular channel for upward and downward flows
. Nucl. Eng. Technol. 46, 195-206 (2014). https://doi.org/10.5516/NET.02.2013.057.A method of correlating heat-transfer data for surface boiling of liquids
. Trans. Am. Soc. Civ. Eng. 74, 969-975 (1952). https://doi.org/10.1115/1.4015984.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), IAPWS R7-97
(2007).Beyond first‐order upwinding: the ultra‐sharp alternative for non‐oscillatory steady‐state simulation of convection
. Int. J. Numer. Meth. Eng. 30, 729-766 (1990). https://doi.org/10.1002/nme.1620300412.On central-difference and upwind schemes
. J. Comput. Phys. 101, 292-306 (1992).Proper orthogonal decomposition and its applications - Part II: Model reduction for MEMS dynamical analysis
. J. Sound Vibr. 256, 515-532 (2002). https://doi.org/10.1006/jsvi.2002.5007.Reduced-order with least square-finite difference method for neutron transport equation
. Ann. Nucl. Energy 191,The authors declare that they have no competing interests.