Open access peer-reviewed chapter - ONLINE FIRST

Two Modeling Approaches for sCO2-PCHEs Transients

Written By

Xin Sui and Peixin Dong

Submitted: 17 March 2024 Reviewed: 10 June 2024 Published: 09 July 2024

DOI: 10.5772/intechopen.1005824

Innovative Heat Exchanger Technologies, Developments and Applications IntechOpen
Innovative Heat Exchanger Technologies, Developments and Applicat... Edited by Peixin Dong

From the Edited Volume

Innovative Heat Exchanger Technologies, Developments and Applications [Working Title]

Peixin Dong and Xin Sui

Chapter metrics overview

6 Chapter Downloads

View Full Metrics

Abstract

Printed circuit heat exchangers (PCHEs) are ideal for sCO2 power cycles due to their compactness, effectiveness, and high-pressure capability. However, their unique architecture complicates the modeling of their dynamic behavior in power cycles, which experience rapid transients. Both the significant computational resources required and the high investment costs of experiments limit their widespread application. To model the component- and system-level transients of sCO2-PCHEs, this study presents two 1D modeling approaches for different purposes: one for component-level simulation based on local properties of sCO2, and the other for system-level simulation based on transfer functions. Given the significant discrepancies observed when using a fixed time constant in the latter approach, this paper introduces the concept of an optimal time constant to model the transient behavior of PCHEs as a first-order system with minimal prediction error. This optimal time constant varies with operating conditions, contrary to what the name might suggest. These results demonstrate the potential of PCHEs in advanced power cycles and provide valuable insights for accurate system-level sCO2 power cycle control studies.

Keywords

  • printed circuit heat exchangers (PCHEs)
  • transient behaviors
  • 1D numerical model
  • low-order approximation
  • optimal time constant

1. Introduction

Supercritical carbon dioxide (sCO2) power cycles are promising solutions for transforming traditional thermal powers to renewable energies [1, 2]. The sCO2 cycles find massive applications across vast working scenarios thanks to their high conversation efficiency but low carbon emissions. Among various types of heat exchangers, printed circuit heat exchangers (PCHEs) are favored as primary heaters and recuperators in sCO2 cycles because of their outstanding heat transfer performance and remarkable resistance to high temperature and pressure. Massive experimental and numerical studies have been conducted to investigate sCO2-PCHEs’ dynamic performance under varying scenarios, such as nuclear power [3, 4], concentrated solar power [1], wasted heat recovery [5, 6] and hypersonic aero-engines [7]. For these investigations, the most challenging task is to handle the difficulties brought by sCO2 fluid’s steep property variations.

However, the temperature-dependent properties violate the assumption of ‘constant property’ employed by most dynamic modeling approaches [8, 9], resulting in significant and lagging prediction errors, for which both three- and one-dimensional simulation techniques have paid special efforts to address the issue. The high-fidelity three-dimensional modeling approaches, such as Computational Fluid Dynamics (CFD) [10, 11] and Direct Numerical Simulation (DNS), capture the dramatic property variations at the price of magnificent computation cost [12]. The one-dimensional modeling methods prefer the discretized approach to capture local properties while requiring remarkable computational costs [13]. On the other hand, AI-assisted technologies have shown promise in enhancing power plant efficiency, but their application in predicting system-level off-design performance and optimizing control strategies remains underexplored [14, 15]. This limitation is primarily due to the scarcity of operational data needed to train AI models effectively. The data, which are derived from experimental tests, require significant investment costs, making them accessible only to a few well-equipped laboratories [16].

Utilizing inappropriate empirical correlations might induce unexpected convergence and stability issues to dynamic modeling [17, 18]. The dynamic simulation, aiming to provide insightful knowledge of sCO2-PCHEs’ transients for control studies, prefers the lower-order system to proxy PCHEs’ transients with acceptable errors. While experimental observations and analytical solutions have confirmed the feasibility of using the first-order system approximation for shell-and-tube and plate heat exchangers [19, 20], the same modeling technique has not been reported for sCO2-PCHEs, as we know. Therefore, this chapter will propose a component-level 1D model based on sCO2 local properties and a first-order approximation for sCO2-PCHEs. The feasibility of both models under different operating conditions are compared. A data-fitting approach will also be presented to compute the key parameter, time constant, of the proposed low-order approximation. As will be seen, a varying time constant, determined by both geometrical parameters and operating conditions, can significantly reduce the approximation’s errors. By the end of this study, the transients of sCO2 PCHEs can be approximated with minimal cost both experimentally and numerically, with good accuracy.

Advertisement

2. Modeling methodologies

2.1 1D numerical model for sCO2-PCHEs

A sCO2 Recompression Brayton Cycle (Figure 1a) usually chooses PCHEs as its precooler, high-temperature and low-temperature recuperator. The sCO2 fluid experiences dramatic thermal-hydraulic variations in these components, as shown in the changes of Prandtl numbers in Figure 1b and c.

Figure 1.

sCO2’s property variations per components of recompression Brayton cycles.

To investigate the dynamics of sCO2 power cycles, 1-D numerical modeling [2] usually discretizes a straight- or zigzag-channel PCHE into N1 heat transfer units (Figure 2), where each unit consists of one hot volume, one cold volume, and one separating metal wall volume. For the jth unit j=1N1, the governing conservation equations for mass, energy, and momentum for either the cold or hot side are,

Figure 2.

Schematic of discretized PCHE model.

Mjt=Alρ:jp:jṗ:j+ρ:jT:jṪ:j+ρ:j+1T:j+1Ṫ:j+1+ẇ:jẇ:j+1E1
Alρjcv,jṪj=w:j+1h:j+1w:jh:j+Qx,jE2
lAwjt=p:jp:j+1Δpfric,jE3

where the mass Mj, averaged density ρj, averaged pressure pj, averaged temperature Tj, averaged specific heat capacity cv,j, and mass flow rate wj are volume-average quantities of volume j, A is the cross-sectional area (identical for node and volume), l=L/N1 is the volume length, ρ:j,p:j,T:j,h:j,w:j+1 are averages of density, pressure, temperature, specific enthalpy, and mass flow rate over the cross-section area of node :j, and Δpfric,j is the local pressure drop through the volume caused by the friction.

For the energy conservation equation (Eq. (2)), Qx,j denotes the heat transferred between a flow volume j and its adjacent wall volume j, where the subscript x is either h for hot side or c for cold side, such as,

Qx,j=Tw,jTx,jωlhcx,jE4
hcx,j=Nux,jkx,jDhydE5

where Tw,j,Tx,j are temperatures of wall volume j and flow volume j, respectively, ω is the wetted perimeter, and hcx,j is the local heat transfer coefficient for convective heat transfer, which is evaluated via the local Nusselt number Nux,j. Thus, for this heat transfer unit j, the heat transferred from hot volume Qh,j to cold volume Qc,Nj is balanced at the in-between wall volume,

lρ¯wAwṪw,j=Qh,j+Qc,NjE6

Lastly, a critical task for the above modeling is to select a proper Nu number correlation to determine the Nusselt numbers Nu and then the local heat transfer coefficients hcx,j (Eq. (5)). A comparison of multiple correlation candidates, as well as more discussions for the above models, are presented in the literature [2].

2.2 A first-order approximation for sCO2-PCHEs

As reported in several published studies, the exit temperature of heat exchangers can be expressed as a first-order approximation [19, 20]. In specific, if one inlet is experiencing a step change in mass flow rate or temperature, the exit temperature is a function of time constant τ and time delay tr, such as,

ttr:Tt=T0t>tr:TtTT0T=expttrτE7

Using this approximation, the dynamic behavior of the heat exchanger can be represented as a first-order system described by a transfer function (TF) [17, 21], such as,

Gs1st=YsFs=Kτs+1exptrsE8

where Ys,Fs are the Laplace transformation of system’s response yt and forcing input ft, respectively, τ, time delay tr and gain K are to-be-determined parameters. For sCO2-PCHEs, the time delay tr usually equals 0 regarding PCHEs’ short flow path, and the gain K normally equals 1 for the symmetry of two flows, i.e., a step change in one flow would lead to a change of the same scale in the other flow. Therefore, this chapter will demonstrate how to calculate the time constant τ with great attention.

2.3 Time constant calculation

The time-domain response of the first-order system (Eq. (8)), when taking a step-down input ft=Aut, can be described as,

Tt=T0et/τ+KAuttr1ettr/τE9

where T0 is the initial value of Tt for ttr. Therefore, we can locate the best guess of time constant through the data-fitting approach. The retrieved time constant τ as well as gain K and delayed time tr should fit the system with the experimental or simulation results with minimum errors [22, 23], such as,

argminτ,K,trR1ni=1nTiyi2E10

where Ti=Tt=ti is the time-domain response of the system (Eq. (9)) at time ti, yi is the PCHE’s outlet temperature at the same moment, i=1,,n indexes n moments of a dynamic experiment or simulation process. In the above optimization, mean square errors (MSEs) are employed to evaluate prediction errors, for which maximum absolute error σmax is an alternative metric, such as,

σmax=maxTiyi,i=1,,nE11
Advertisement

3. Validation case

3.1 Set-ups

This section sets up a parameterized, general-purpose Modelica program to simulate sCO2 PCHEs’ dynamic behavior under various experimental scenarios, with which we can calculate time constant τ accordingly. The program, as published in our open source project (SCOPE) [2], adopts two source mass flow components and two sink pressure components. By setting the temperature, mass flow rate, and pressure of the sources to the experimental scenarios, the program can simulate the PCHE’s dynamic behavior under various geometrical parameters and operating conditions. The graphical user interface of the simulation program is illustrated in Figure 3.

Figure 3.

Modelica test program GUI.

3.2 First-order system validation

We select the configuration (Table 1) of a HTR of a 10 WMe sCO2 Recompressed Brayton Cycle [24] as a demonstration. When taking a step-down change in hot inlet mass flow rate, the hot outlet temperature exponentially drops as defined in Eq. (9). The time constant τ identified by Eq. (10) is 251.92s, which yields a good match between SCOPE’s simulation output and the first-order system, with an MSE of 0.6761°C2 and a σmax=1.4188°C (Figure 4a, Table 2).

NameSymbolValue
Hot/Cold inlet temperatureTh,in/Tc,in (°C)578.15 / 181.95
Hot/Cold outlet temperatureTh,out/Tc,out (°C)192.0 / 532.1
Hot/Cold inlet pressurePh,in/Pc,in (bar)89.62 / 237.48
Hot/Cold mass flow rateṁh/ṁc (kg/s)100.17 / 100.17
Weight / HeightW/H (mm)360 / 3710
Channel lengthLfp (mm)800
Channel diameterDch (mm)1.5
Metal wall thicknessTwall (mm)0.51
Channel numberNch ([1])700,000

Table 1.

Geometry and operation parameters for time constant evaluation tests.

Figure 4.

Comparison between local property-based simulation and LTI system prediction.

Name caseStandardOff-Des. case AOff-Des. case B
ṁhot,in100.17 kg/s95 kg/s105 kg/s
τop251.92197.66320.08
MSE0.67610.32091.2938
σmax1.41892.16621.8018
MSEstd206.7400200.5752
σmax,std17.288917.3021

Table 2.

Detailed discrepancies between standard case and two off-design cases.

However, the same value of τ will produce significant errors when applying for other starting mass flow rates (Figure 4b and c). For a below-design-point operation where ṁhot,in starts from 95kg/s (Figure 4b), the hot outlet temperature predicted by the first-order system, with τ=251.92, diverges from the SCOPE’s result by an MSE of 206.74°C2 and an σmax,std of 17.2889C at t=1000s. Similarly, for an above-design-point operation where ṁhot,in starts from 105kg/s, the discrepancies are MSE=320.08°C2 and σmax,std=17.3021°C at t=1000s. The remarkable discrepancies reflect the drawback of the assumption of constant heat transfer coefficients held by the low-order system approximation. Under the assumption, the time constant is a system instinct property determined by heat exchangers’ geometrical parameters and should remain constant during a system’s transient. However, for sCO2-PCHEs where the working fluid experiences steep non-linear variations, the time constant’s value varies as the initial operating condition changes. As a consequence, employing the τ identified at the on-design condition will produce significant prediction discrepancies if the PCHE works at alternative operating conditions (Figure 4b and c). A variable time constant, on the contrary, can achieve minimal prediction errors under various operating conditions.

To prove the above point, we identify the value of time constants per operating condition, which produces good agreements between the first-order system and SCOPE’s outputs. In specific, a time constant τ valued at 197.66 produces a prediction with MSE=0.3209°C2 and σmax=2.1662°C for the below-design-point operation (Figure 4b) and a time constant τ valued at 320.08 generates an MSE=1.2938°C2 and σmax=1.8018°C for the above-design-point case (Figure 4c). Therefore, it is reasonable to introduce a notation named optimal time constant, τop, as a refined key parameter to generate minimal prediction errors for a sCO2 PCHE applications under varying configurations toward dynamic and control strategic studies.

Advertisement

4. Conclusions

For sCO2-PCHEs, a variable time constant determined by the system’s operating conditions is favorable to approximating the component’s dynamic behavior with a first-order system. Moreover, PCHEs are critical components for sCO2 power blocks considering their substantial impacts on the heat transformation and flexibility of the entire power block. A better quantitative understanding of its flexibility through analyzing time constant can effectively serve the investigation of the control strategy of sCO2 power cycles. Understanding the τop‘s distribution under varying operating conditions provides informative knowledge for designing and optimizing control strategies. For example, it is reasonable to maintain the working fluids at relatively low initial mass flow rates and alter the mass flow rate with the highest possible magnitude (within the permitted limit) to achieve the lowest time constant and highest flexibility. Therefore, for power block control studies, it is essential to take the trade-off between τ and η into account and utilize the advantages to achieve the control objective with the highest speed but lowest cost.

References

  1. 1. Lock A, Bone V. Off-design operation of the dry-cooled supercritical CO2 power cycle. Energy Conversion and Management. 2022;251:114903. DOI: 10.1016/j.enconman.2021.114903
  2. 2. Sui X, Russell H, Gurgenci H, Lee S, Guan Z, Dong P. A one-dimensional modelling methodology of printed circuit heat exchangers for steady, off-design and transient feasibility in a supercritical CO2 power block. International Journal of Heat and Mass Transfer. 2022;193:122938. DOI: 10.1016/j.ijheatmasstransfer.2022.122938
  3. 3. Park JH, Kwon JG, Kim MH, Cha JE, Jo H. Experimental investigation of buoyancy effects on local heat transfer of supercritical pressure CO2 in horizontal semicircular tube. International Journal of Heat and Mass Transfer. 2021;164:120496. DOI: 10.1016/j.ijheatmasstransfer.2020.120496
  4. 4. Park JH, Kwon JG, Kim TH, Kim MH, Cha J-E, Jo H. Experimental study of a straight channel printed circuit heat exchanger on supercritical CO2 near the critical point with water cooling. International Journal of Heat and Mass Transfer. 2020;150:119364. DOI: 10.1016/j.ijheatmasstransfer.2020.119364
  5. 5. Zhang H, Shi L, Xuan W, Chen T, Li Y, Tian H, et al. Analysis of printed circuit heat exchanger (PCHE) potential in exhaust waste heat recovery. Applied Thermal Engineering. 2022;204:117863. DOI: 10.1016/j.applthermaleng.2021.117863
  6. 6. Alfani D, Binotti M, Macchi E, Silva P, Astolfi M. sCO2 power plants for waste heat recovery: Design optimization and part-load operation strategies. Applied Thermal Engineering. 2021;195:117013. DOI: 10.1016/j.applthermaleng.2021.117013
  7. 7. Li H, Liu H, Zou Z. Experimental study and performance analysis of high-performance micro-channel heat exchanger for hypersonic precooled aero-engine. Applied Thermal Engineering. 2021;182:116108. DOI: 10.1016/j.applthermaleng.2020.116108
  8. 8. Liu G, Huang Y, Wang J, Liu R. A review on the thermal-hydraulic performance and optimization of printed circuit heat exchangers for supercritical CO2 in advanced nuclear power systems. Renewable and Sustainable Energy Reviews. 2020;133:110290. DOI: 10.1016/j.rser.2020.110290
  9. 9. Bone V, McNaughton R, Kearney M, Jahn I. Methodology to develop off-design models of heat exchangers with non-ideal fluids. Applied Thermal Engineering. 2018;145:716-734. DOI: 10.1016/j.applthermaleng.2018.09.082
  10. 10. Saeed M, Berrouk AS, Salman Siddiqui M, Ali Awais A. Numerical investigation of thermal and hydraulic characteristics of sCO2-water printed circuit heat exchangers with zigzag channels. Energy Conversion and Management. 2020;224:113375. DOI: 10.1016/j.enconman.2020.113375
  11. 11. Saeed M, Ali Awais A, Berrouk AS. CFD aided design and analysis of a precooler with zigzag channels for supercritical CO2 power cycle. Energy Conversion and Management. 2021;236:114029. DOI: 10.1016/j.enconman.2021.114029
  12. 12. Zhou L, Garg D, Qiu Y, Kim S-M, Mudawar I, Kharangate CR. Machine learning algorithms to predict flow condensation heat transfer coefficient in mini/micro-channel utilizing universal data. International Journal of Heat and Mass Transfer. 2020;162:120351. DOI: 10.1016/j.ijheatmasstransfer.2020.120351
  13. 13. Kwon JS, Bae SJ, Heo JY, Lee JI. Development of accelerated PCHE off-design performance model for optimizing power system operation strategies in s-CO2 Brayton cycle. Applied Thermal Engineering. 2019;159:113845. DOI: 10.1016/j.applthermaleng.2019.113845
  14. 14. Shahzad MW, Nguyen VH, Xu BB, Tariq R, Imran M, Ashraf WM, et al. Machine learning assisted prediction of solar to liquid fuel production: A case study. Process Safety and Environmental Protection. 2024;184:1119-1130. DOI: 10.1016/j.psep.2024.02.060
  15. 15. Muhammad Ashraf W, Moeen Uddin G, Afroze Ahmad H, Ahmad Jamil M, Tariq R, Wakil Shahzad M, et al. Artificial intelligence enabled efficient power generation and emissions reduction underpinning net-zero goal from the coal-based power plants. Energy Conversion and Management. 2022;268:116025. DOI: 10.1016/j.enconman.2022.116025
  16. 16. Yang J, Yang Z, Duan Y. A review on integrated design and off-design operation of solar power tower system with s-CO2 Brayton cycle. Energy. 2022;246:123348
  17. 17. Chen J, Cui G, Xiao Y. An analytical solution to the dynamic behavior of heat exchanger networks. International Journal of Heat and Mass Transfer. 2018;126:466-478. DOI: 10.1016/j.ijheatmasstransfer.2018.05.041
  18. 18. Yang R, Tran CT. An analytical heat exchanger model to study dynamic behavior in the case of simultaneous inlet variations. International Journal of Thermal Sciences. 2022;174:107452. DOI: 10.1016/j.ijthermalsci.2021.107452
  19. 19. Lachi M, El Wakil N, Padet J. The time constant of double pipe and one pass shell-and-tube heat exchangers in the case of varying fluid flow rates. International Journal of Heat and Mass Transfer. 1997;40(9):2067-2079. DOI: 10.1016/S0017-9310(96)00274-8. Available from: https://www.sciencedirect.com/science/article/pii/S0017931096002748
  20. 20. Das SK, Spang B, Roetzel W. Dynamic behavior of plate heat exchangers—Experiments and modeling. Journal of Heat Transfer. 1995;117(4):859-864. DOI: 10.1115/1.2836302
  21. 21. Cui GM, Wang JY, Jiang T, Hu XB. Transfer behavior of heat exchanger. Advanced Materials Research. 2011;233–235:1281-1287. DOI: 10.4028/www.scientific.net/AMR.233-235.1281
  22. 22. Alfani D, Astolfi M, Binotti M, Campanari S, Casella F, Silva P. Multi objective optimization of flexible supercritical CO2 coal-fired power plants. In: Volume 3: Coal, Biomass, Hydrogen, and Alternative Fuels; Cycle Innovations; Electric Power; Industrial and Cogeneration; Organic Rankine Cycle Power Systems. Phoenix, Arizona, USA: American Society of Mechanical Engineers; 2019. p. V003T06A027. DOI: 10.1115/GT2019-91789
  23. 23. Ma T, Li M-J, Xu J-L, Ni J-W, Tao W-Q, Wang L. Study of dynamic response characteristics of s-CO2 cycle in coal-fired power plants based on real-time micro-grid load and a novel synergistic control method with variable working conditions. Energy Conversion and Management. 2022;254:115264. DOI: 10.1016/j.enconman.2022.115264
  24. 24. Jiang Y, Liese E, Zitney SE, Bhattacharyya D. Design and dynamic modeling of printed circuit heat exchangers for supercritical carbon dioxide Brayton power cycles. Applied Energy. 2018;231:1019-1032. DOI: 10.1016/j.apenergy.2018.09.193

Written By

Xin Sui and Peixin Dong

Submitted: 17 March 2024 Reviewed: 10 June 2024 Published: 09 July 2024