Abstract
With the increase of the renewable energy generator capacity, the requirements of the power system for grid-connected converters are evolve, which leads to diverse control schemes and increased complexity of systematic stability analysis. Although various frequency-domain models are developed to identify oscillation causes, the discrepancies between them are rarely studied. This study aims to clarify these discrepancies and provide circuit insights for stability analysis by using different frequency-domain models. This study emphasizes the limitations of assuming that the transfer function of the self-stable converter does not have right half-plane (RHP) poles. To ensure that the self-stable converters are represented by a frequency-domain model without RHP poles, the applicability of this model of grid-following (GFL) and grid-forming (GFM) converters is discussed. This study recommends that the GFM converters with ideal sources should be represented in parallel with the - admittance model rather than the - impedance model. Two cases are conducted to illustrate the rationality of the - admittance model. Additionally, a hybrid frequency-domain modeling framework and stability criteria are proposed for the power system with several GFL and GFM converters. The stability criteria eliminates the need to check the RHP pole numbers in the non-passive subsystem when applying the Nyquist stability criterion, thereby reducing the complexity of stability analysis. Simulations are carried out to validate the correctness of the frequency-domain model and the stability criteria.
THE global pursuit of a sustainable and low-carbon future places significant emphasis on renewable energy generation, driving consequential modifications in the operation, structure, and dynamics of power systems. A notable manifestation of these changes is the transition from synchronous generator-dominated dynamics to converter-dominated dynamics in power systems [
Impedance analysis methods play a crucial role in identifying the underlying causes of oscillations in converter-dominated power systems, which effectively explore the intricate dynamics and interactions between converters and the grid, enabling the formulation of impedance models in various forms [
Originally concentrated on single-converter grid-connected systems, the frequency-domain models based on impedance or transfer function models extend their applicability to systems with several converters in GFL and GFM control. Therefore, the circuit representation of the converters is determined by the current and voltage characteristics of the terminal [
The Bode plots and generalized Nyquist criterion (GNC) are used in [
The GFL and GFM converters can be represented as equivalent power sources connected in series with - impedance or in parallel with - admittance [
To eliminate the need to check the RHP pole numbers in the non-passive subsystem, this study revisits the presence of RHP poles in self-stable converter as well as the adaptability of the frequency-domain model, and explores the hybrid frequency-domain modeling framework for the power system with GFL and GFM converters. The main contributions of this study can be summarized as follows.
1) This study contributes to clarifying the discrepancies between the V-I and - models in stability analysis, particularly rectifying the vague understanding regarding the pole numbers of the frequency-domain model of a self-stable equipment.
2) This study investigates the appropriate model selection, either the V-I or the - impedance/admittance model, for the stability analysis of converters in different control modes.
3) This study introduces a hybrid frequency-domain modeling framework and stability criteria for power systems with GFL and GFM converters. The frequency-domain model simplifies the stability analysis process by eliminating the need to check the RHP pole numbers in the non-passive subsystem.
The rest of this paper is organized as follows. Section II briefly introduces the system configuration and the frequency-domain models. Then, the circuit insights and discrepancy of frequency-domain models are discussed in Section III. Afterward, Section IV discusses the frequency-domain model applicability for converters with different control strategies. Section V presents the multi-converter parallel system. Finally, Section VI outlines the main conclusions.

Fig. 1 Studied system with GFL and GFM converters. (a) Topology of studied system. (b) Topology of grid-connected converter.
It is worth mentioning that all the converters mentioned in this study are grid-connected voltage-source converters (VSCs). The common control structures of GFL and GFM converters are shown in

Fig. 2 Common control structures of GFL and GFM converters. (a) GFL converter. (b) GFM converter.

Fig. 3 Control diagram of GFM converters. (a) Power-based synchronization loop controller. (b) Q-V droop controller.
Thus, although this study analyzes only two typical configurations, the presented tool can be adapted for other options of converter control implementation in different applications.
To analyze the interaction between the converters and the grid, the existing studies usually use and as the input and output signals of the frequency-domain model, respectively, deriving a V-I impedance or admittance model of the converter. The V-I impedance model of the converter can accordingly be expressed as [
(1) |
where the superscripts d and q denote the d-axis and q-axis parameters, respectively; Z is the impedance; and denotes the small disturbance component of the signal.
To analyze the power-frequency dynamic characteristics and their interaction of the converters, the output active power and reactive power Qe are usually used as input signals. The amplitude Vac and phase (frequency) of the voltage at converter switching bridge or PCC voltages are used as output signals [
(2) |
where , , , and are the elements of the equivalent impedance .
This section introduces the circuit insights and stability criteria based on different frequency-domain models. This section also studies the discrepancy of frequency-domain models of the converters and transmission lines.
The relationship between the output power and the PCC voltage/current, as well as the relationship between the dq-axis voltage and the amplitude and phase of the voltage at the converter switching bridge, is given as [
(3) |
(4) |
where the subscript 0 denotes the steady-state value; and I and V denote the current and voltage, respectively.
Substituting (3) and (4) into (2), the mathematical equivalence of the converter and the two subsystems using the - model can be expressed as [
(5) |
where the subscripts “vsc” and “g” denote the parameters associated with the converter and the grid, respectively. Similarly, if the converter employs GFL or GFM control, the subscript will change from “vsc” to “gfl” or “gfm”.
(6) |

Fig. 4 Equivalent circuit model of grid-connected converter system. (a) Norton equivalent model based on V-I model. (b) Thevenin equivalent model 1. (c) Thevenin equivalent model 2. (d) Simplified equivalent circuit based on - model.
where and are the reference voltage vector and closed-loop reference-to-output frequency-domain model of the converter, respectively; is the equivalent admittance; and is the identity matrix.
Following the Norton equivalent model and Ohm’s law,
(7) |
where and are the reference amplitude vectors of the equivalent current sources on grid and converter sides, respectively.
Subsequently, the Thevenin equivalent model 2 is obtained by adjusting the equivalent admittance of two subsystems with a coefficient matrix , as shown in
(8) |
Then, the - model of the two subsystems can be calculated from the derived impedance models and frequency sweep models. By multiplying the left and right sides of (8) by the inverse matrix and substituting it into the - model, the PCC voltage can be calculated as:
(9) |
where ; and .
Generally, the mathematical description of the converter typically comprises two components: the controller and the filter circuit. Neglecting the DC-side dynamics and switching losses of the converter, the output pulse width modulation signals of the control system can be approximately equal to the voltage at the switching bridge of the converter . That is to say,, the output signal of the control system is , which must be one of the input signals of the converter filter circuit. The state-space models of the two components can be expressed as:
(10) |
(11) |
where F, H, J, and K are the diagonal parameter matrices in the state-space representation of modules; x, a, and b are the state variables, input signals, and output signals, respectively; and the subscripts ctrl and filter represent the controller module and filter circuit module, respectively.
Moreover, the interconnection between the input signal and output signal of the composite system and the input/output signals of each module can be expressed by the algebraic equations as:
(12) |
(13) |
where uvsc and yvsc are the input and output signals for the converters of the composite system, respectively; and L1, L2, L3, and L4 are the parameter matrices that map the interconnection relationships among different components.
Then, according to the component connection-based modular state-space modeling method in [
(14) |
(15) |
where is the parameter matrix, and Fvsc, Hvsc, and Jvsc are the diagonal matrices with parameter matrices F, H, and J of submodule (control system and filter) as the diagonal elements; and , , and are the the parameter matrices of the overall state-space model of the converter.
Equations (
(16) |
Then, substituting (3) into (16) yields:
(17) |
Selecting the input signal in V-I model as the voltage , the output signal can be expressed in a linear form related to , namely . Then, the relationship between in the V-I model and in the - model is:
(18) |
It is worth noting that the matrices and may not be equal, as indicated by (18). This implies that the parameter matrix is not necessarily equal to the parameter matrix , because the matrices , , , and are the same. This suggests that the absence of an RHP pole in one frequency-domain model does not guarantee the absence of an RHP pole in others, as zero-pole cancellation can occur in the frequency-domain model.
The dynamic equations of the grid with inductor Lg and resistance in dq frame can be given as:
(19) |
where is the rated angular velocity of the grid.
Substituting (19) and (4) into (3) to eliminate the variables and , the relationship among the transmission power and the voltages at two ends can be given as:
(20) |
(21) |
(22) |
where ; ; and the coefficients F2, F3, F4, and F5 are given in Appendix A.
From (19), (21), and (22), the frequency-domain models , , and do not exhibit RHP poles as the poles of and are . The RHP poles of can be obtained by solving . From (A6)-(A9) in Appendix A, it is observed that the - impedance model exhibits one RHP pole when . The steady-state operating point is set with p.u., p.u., p.u., and and are usually less than 1 p.u., which is easily satisfied in practice. Therefore, the - impedance model of the grid has one RHP pole, which highlights that different model forms of the grid may have discrepancies in their RHP poles. Even for a stable equipment, one model may have RHP poles, whereas the other models may not.
To conclude, both the V-I model and the - model are utilized in the stability analysis for grid-connected converter systems. While the utilization of these models for assessing the stability of the entire closed loop yields consistent results, there may be discrepancies in how accurately different frequency-domain models describe the physical properties of a system. It is crucial to recognize that not all models can accurately reflect the self-stability nature of the apparatus. Improper choice of the model results in an additional step in the stability analysis, requiring the examination of RHP poles of the open-loop gain matrices.
This section focuses on the selection of the appropriate frequency-domain model of the converter in stability analysis based on its design model. Two cases are given to illustrate the rationality of the frequency-domain model.
By using the current decoupling control and neglecting the dynamics of PLL, the d-axis closed-loop frequency-domain model of GFL converters with a proportional-integral (PI) inner controller and a filter circuit is given as [
(23) |
where is the d-axis component of the desired current; and is the filter inductance.
Similarly, the q-axis frequency-domain model can be obtained. Obviously, the design model of GFL converter involves a frequency-domain model with current as the input signal and voltage as the output signal. Furthermore, during the design process of the converter, both frequency-domain models and are ensured to be free of RHP poles and exhibit a certain stability margin. Considering that the design model can be considered as a V-I admittance model that neglects the dynamics of PLL, it is anticipated that the V-I admittance model of the GFL converter does not possess any RHP poles. However, it is worth mentioning that there is no absolute guarantee that the V-I admittance model of the converter is completely free of RHP poles, especially under non-unity power factor conditions. Fortunately, the converter parameters are optimized through a trial-and-error process to ensure the self-stability of the converter. In contrast to the - model, which is not involved during the design process of the converter, it is more reasonable that the self-stability of the converter implies the absence of RHP poles in the V-I admittance model of the GFL converters.
Thus, a more appropriate approach for assessing the stability is to model the GFL converter as an ideal current source in parallel with the V-I admittance model. The stability analysis cases for GFL converters in [
The converter under the GFM control is given as:
(24) |
where and are the inertias of active power loop and reactive power loop, respectively; and are the active droop coefficient and the reactive droop coefficient, respectively; is the PCC voltage phasor; and , , and are the active power reference, the reactive power reference, and the rated root mean square value of the PCC voltage, respectively.
Assuming that the grid is inductive, the inductive component of the converter filter and grid inductor is significantly larger than the resistance component . Thus, we can neglect the resistance component. Disregarding the transmission line dynamics and the power coupling term [
(25) |
where is the power angle between the converter voltage and the grid voltage.
Therefore, the closed-loop transfer functions of the active power loop and reactive power loop can be easily expressed as:
(26) |
In contrast to the GFL converter, the parameter tuning model (26) of the GFM converters does choose output power and the magnitude/phase of the PCC voltage as input/output signals [

Fig. 5 Bode diagram of and .
Module | Parameter | Symbol | Value |
---|---|---|---|
Filter circuit | Rated power | 2 MW, 2 MW | |
Rated frequency | 50 Hz | ||
Transformer ratio | 66 kV/690 V | ||
Equivalent inductance and resistance of grid | 0.2 p.u., 0.02 p.u. | ||
Filter inductance, capacitance, and resistance | 26.3 mH, 40 μF,0.5 | ||
Controller | Vector current controller | 0.33, 0.6283 | |
Direct-voltage controller | 1.47, 132 | ||
PLL | 40, 800 | ||
Inertia of active power loop and active droop coefficient | 0.02, 0.02 | ||
Inertia of reactive power loop and reactive droop coefficient | 0.02, 0.20 |
For an active power loop, the crossover frequency is 14.6 Hz, the phase margin is 28.5°, and the magnitude of the active power loop gain at 100 Hz is dB (which corresponds to 0.0239).
In addition, the magnitude of the reactive power loop gain at 100 Hz is dB (which corresponds to 0.0749). Therefore, the amplitude margin and the phase margin of the active power loop and the reactive power loop satisfy the required specifications, and the GFM converter is deemed self-stable [
Set the series compensation level (SCL) as 36% and 60%, respectively. In

Fig. 6 Nyquist curves using V-I model and - model , pole distribution, and eigenvalue distribution. (a) Nyquist curves of when . (b) Nyquist curves of when . (c) Eigenvalue distribution of closed-loop system. (d) Pole distribution of and of GFM converter. (e) Nyquist curves of when . (f) Nyquist curves of when .
However, by examining the eigenvalue distribution of the closed-loop system shown in
To validate the stability analysis conclusion, the nonlinear model of a single-converter grid-connected system is built on the MATLAB/Simulink platform.

Fig. 7 Time-domain analysis results of single-converter grid-connected system. (a) Three-phase output current and active power waveforms of converter when . (b) Three-phase output current and active power waveforms of converter when . (c) FFT results from output current when .
In summary, the utilization of different models during the design process of GFL and GFM converters leads to different interpretations of self-stability. For the GFL converter, self-stability implies the absence of RHP poles in the V-I admittance model, whereas the RHP pole number in the - model remains uncertain. Conversely, the self-stability in the GFM converter denotes the absence of an RHP pole in the - admittance model, while the pole in the V-I model may not possess this attribute. To reduce the complexity of the stability analysis, it is recommended that GFL converters are modeled as an ideal current source in parallel with the V-I admittance model, while GFM converters are modeled as an ideal power source in parallel with the - admittance model.
This section presents a hybrid frequency-domain modeling framework and stability criteria for multi-converter parallel systems with the frequency-domain models of GFL and GFM converters built in the V-I and - models, respectively. The stability criteria are validated through a simulation case involving a studied system with both GFL and GFM converters.
The assessment of the RPH pole number of the high-order transfer function matrix can be challenging, given the complexity involved in both calculating the poles of this matrix and identifying the parameters of the measurement model. The potential existence of RHP poles in the frequency-domain model of converter and the impedance aggregation process contribute to the emergence of RHP poles in the subsystem model. From the analysis in Section IV, it is evident that the V-I model of the GFM converter may exhibit RHP poles, resulting in an inaccurate conclusion. To address this, it is recommended to set the GFM converters as an ideal power source in parallel with the - admittance model. Subsequently, the V-I admittance of the GFL converter, the - admittance of the GFM converter, and the V-I impedance of the equivalent grid are embedded as diagonal elements in the transfer function matrices of the non-passive subsystem. In this modeling approach, each component in the non-passive system operates independently as an individual subsystem without interacting with other components. This ensures that the poles of the system models are the union of the poles of each submodule, thereby preventing the emergence of RHP poles during the impedance aggregation process.
Assuming each power plant shares the same control structure and parameters, each power plant is simplified and represented as a single equivalent source converter (ESC) using the capacity-weighted average method [
(27) |
where subscripts cs and vs denote the collective vectors or matrices of the equivalent GFL converter and the equivalent GFM converter, respectively; is the dq component of the voltage at the point ; and is the transfer function matrix, which is a diagonal matrix with equipment admittances or impedances as diagonal elements, .
According to (20), the small-signal representation of transmission line from the
(28) |
where and are the output vectors and disturbance vectors of the
Substituting (19) into (3) to eliminate the variable , the transmission line current from the
(29) |
With respect to the topology shown in
(30) |
where is the parameter matrix, which is obtained by summing for k ranging from to , namely ; and other parameters are given as:
(31) |
Substituting (30) to (27), the closed-loop transfer function of the overall system can be written as:
(32) |
where is the transfer function matrix representing the small-signal dynamics of the interlinking line.
With the assumption that the converters remain self-stability without considering the grid dynamics, the transfer function matrix is assumed to be free of RHP poles. However, the impedance matrix may contain RHP poles, as discussed in Section III-C. Therefore, the pole numbers of the system open-loop gain equal the pole numbers of . If the encirclement number of the system open-loop gain around the point satisfies , the system loses stability according to the GNC.
In
Model | RPH pole in system open-loop gain |
---|---|
Impedance aggregation model [ | Existence of RPH poles in system open-loop gain can stem from either converter itself or impedance aggregation process. RPH pole number in both subsystems must be checked before evaluating stability of power system. |
V-I model [ | Emergence of RHP poles in system open-loop gain may arise from converter itself. Before evaluating stability of power system with GFL and GFM converters, it is essential to check RHP pole number in the model of non-passive subsystem. Such checks are not required for power systems exclusively employing GFL converters. |
Hybrid frequency-domain model | There has no requirement to check presence of RPH poles in model of non-passive subsystem. For power system with GFL and GFM converters, it is crucial to check RPH pole number in model of passive subsystem. |
To verify the effectiveness and merit of the frequency-domain model, the stability analysis methods based on impedance aggregation model [
Referring to [
(33) |
(34) |
(35) |
Therefore, , , and hybrid models and of the transmission line can be written as:
(36) |
(37) |
(38) |
(39) |

Fig. 8 Pole distribution of frequency-domain models of converters and two subsystems. (a) Frequency-domain models of converters. (b) Two subsystems.

Fig. 9 Critical Nyquist curves of , , and . (a) . (b) . (c) .
As each of the models , , and has one RHP pole, three models reach the same conclusion that the system loses stability with two RHP poles.
The analysis conclusion aligns with the eigenvalue distribution, as well as the waveforms of the output current and active power at the point depicted in

Fig. 10 Time-domain analysis results of multi-converter parallel system. (a) Eigenvalue distribution of closed-loop system. (b) Three-phase output current and active power waveforms at point . (c) FFT results from output current at point .
To assess the stability of power-electronics-based power systems, the frequency-domain analysis methods have emerged as a crucial analytical tool widely employed [
Another commonly used stability analysis method involves employing the Nyquist criterion and the phase and gain margins from the Bode diagrams [
Regrettably, unlike in single-input single-output (SISO) systems, obtaining system stability margin information through Bode diagrams is not feasible when analyzing multi-input multi-output (MIMO) systems using this method. Actually, the current research does not clearly define the stability margin of MIMO systems unless the MIMO system is decoupled into multiple SISO systems [
This study highlights that there is no definitive guarantee that the poles of both the V-I model and the - model of a stable equipment reside exclusively in the left-half plane. To guarantee a frequency-domain model without an RHP pole, the applicability of frequency-domain models of self-stable converters is discussed. Following the design model, this study suggests that GFL converters are more suitably represented by an ideal current source in parallel with V-I admittance model, while GFM converters are more suitably represented by an ideal power source in parallel with - admittance model. According to the principle of equivalence, this paper proposes a hybrid frequency-domain modeling framework and stability criteria for grid-connected converter system with the frequency-domain models of GFL and GFM converters built in the V-I and - framework, respectively, which eliminates the requirements for checking RHP poles in the non-passive subsystem model when applying the Nyquist criterion. By avoiding this step, the complexity of modeling and stability analysis for complex systems is reduced.
References
J. Shair, H. Li, J. Hu et al., “Power system stability issues, classifications and research prospects in the context of high-penetration of renewables and power electronics,” Renewable and Sustainable Energy Reviews, vol. 145, p. 111111, Jul. 2021. [Baidu Scholar]
Y. Cheng, L. Fan, J. Rose et al., “Real-world subsynchronous oscillation events in power grids with high penetrations of inverter-based resources,” IEEE Transactions on Power Systems, vol. 38, no. 1, pp. 316-330, Jan. 2023. [Baidu Scholar]
B. Wen, D. Boroyevich, R. Burgos et al., “Analysis of D-Q small-signal impedance of grid-tied inverters,” IEEE Transactions on Power Electronics, vol. 31, no. 1, pp. 675-687, Jan. 2016. [Baidu Scholar]
R. H. Lasseter, Z. Chen, and D. Pattabiraman, “Grid-forming inverters: a critical asset for the power grid,” IEEE Journal of Emerging and Selected Topics in Power Electronics, vol. 8, no. 2, pp. 925-935, Jun. 2020. [Baidu Scholar]
F. Zhao, X. Wang, and T. Zhu, “Power dynamic decoupling control of grid-forming converter in stiff grid,” IEEE Transactions on Power Electronics, vol. 37, no. 8, pp. 9073-9088, Aug. 2022. [Baidu Scholar]
S. Shah, W. Yan, V. Gevorgian et al., “Power-domain impedance theory for the analysis and mitigation of interarea oscillations,” in Proceedings of 2020 IEEE Power & Energy Society General Meeting, Montreal, Canada, Aug. 2020, pp. 1-5. [Baidu Scholar]
Z. Yang, C. Mei, S. Cheng et al., “Comparison of impedance model and amplitude-phase model for power-electronics-based power system,” IEEE Journal of Emerging and Selected Topics in Power Electronics, vol. 8, no. 3, pp. 2546-2558, Sept. 2020. [Baidu Scholar]
C. Li, J. Liang, L. Cipcigan et al., “DQ impedance stability analysis for the power-controlled grid-connected inverter,” IEEE Transactions on Energy Conversion, vol. 35, no. 4, pp. 1762-1771, Dec. 2020. [Baidu Scholar]
Y. Li, Y. Gu, Y. Zhu et al., “Impedance circuit model of grid-forming inverter: visualizing control algorithms as circuit elements,” IEEE Transactions on Power Electronics, vol. 36, no. 3, pp. 3377-3395, Mar. 2021. [Baidu Scholar]
X. Wang, F. Blaabjerg, and W. Wu, “Modeling and analysis of harmonic stability in an AC power-electronics-based power system,” IEEE Transactions on Power Electronics, vol. 29, no. 12, pp. 6421-6432, Dec. 2014. [Baidu Scholar]
W. Cao, Y. Ma, and F. Wang, “Sequence-impedance-based harmonic stability analysis and controller parameter design of three-phase inverter-based multibus AC power systems,” IEEE Transactions on Power Electronics, vol. 32, no. 10, pp. 7674-7693, Oct. 2017. [Baidu Scholar]
Z. Zou, J. Tang, X. Wang et al., “Modeling and control of a two-bus system with grid-forming and grid-following converters,” IEEE Journal of Emerging and Selected Topics in Power Electronics, vol. 10, no. 6, pp. 7133-7149, Dec. 2022. [Baidu Scholar]
G. Li, Y. Chen, A. Luo et al., “An enhancing grid stiffness control strategy of STATCOM/BESS for damping sub-synchronous resonance in wind farm connected to weak grid,” IEEE Transactions on Industrial Informatics, vol. 16, no. 9, pp. 5835-5845, Sept. 2020. [Baidu Scholar]
N. Liu, H. Wang, L. Sun et al., “Modular impedance modeling and stability analysis of hybrid AC/DC power systems with grid-forming and grid-following converters,” IEEE Access, vol. 12, pp. 4063-4077, Jan. 2024. [Baidu Scholar]
J. Samanes, A. Urtasun, E. L. Barrios et al., “Control design and stability analysis of power converters: the MIMO generalized Bode criterion,” IEEE Journal of Emerging and Selected Topics in Power Electronics, vol. 8, no. 2, pp. 1880-1893, Jun. 2020. [Baidu Scholar]
W. Cao, Y. Ma, L. Yang et al., “D-Q impedance based stability analysis and parameter design of three-phase inverter-based AC power systems,” IEEE Transactions on Industrial Electronics, vol. 64, no. 7, pp. 6017-6028, Jul. 2017. [Baidu Scholar]
Y. Li, Z. Shuai, X. Liu et al., “Stability analysis and location optimization method for multiconverter power systems based on nodal admittance matrix,” IEEE Journal of Emerging and Selected Topics in Power Electronics, vol. 9, no. 1, pp. 529-538, Feb. 2021. [Baidu Scholar]
J. Jia, X. Yan, B. Qin et al., “Modeling and analysis of the torque-frequency dynamics for multi-VSC parallel system based on the equivalent admittance,” IEEE Transactions on Power Delivery, vol. 37, no. 5, pp. 3597-3607, Oct. 2022. [Baidu Scholar]
Z. Yang, M. Zhan, D. Liu et al., “Small-signal synchronous stability of a new-generation power system with 100% renewable energy,” IEEE Transactions on Power Systems, vol. 38, no. 5, pp. 4269-4280, Sept. 2023. [Baidu Scholar]
Y. Xu, M. Zhang, L. Fan et al., “Small-signal stability analysis of type-4 wind in series-compensated networks,” IEEE Transactions on Energy Conversion, vol. 35, no. 1, pp. 529-538, Mar. 2020. [Baidu Scholar]
L. Harnefors, M. Hinkkanen, U. Riaz et al., “Robust analytic design of power-synchronization control,” IEEE Transactions on Industrial Electronics, vol. 66, no. 8, pp. 5810-5819, Aug. 2019. [Baidu Scholar]
F. Liu, J. Liu, H. Zhang et al., “Stability issues of Z+Z type cascade system in hybrid energy storage system (HESS),” IEEE Transactions on Power Electronics, vol. 29, no. 11, pp. 5846-5859, Nov. 2014. [Baidu Scholar]
D. Yang and X. Wang, “Unified modular state-space modeling of grid-connected voltage-source converters,” IEEE Transactions on Power Electronics, vol. 35, no. 9, pp. 9700-9715, Sept. 2020. [Baidu Scholar]
L. Harnefors, M. Bongiorno, and S. Lundberg, “Input-admittance calculation and shaping for controlled voltage-source converters,” IEEE Transactions on Industrial Electronics, vol. 54, no. 6, pp. 3323-3334, Dec. 2007. [Baidu Scholar]
H. Wu, X. Ruan, D. Yang et al., “Small-signal modeling and parameters design for virtual synchronous generators,” IEEE Transactions on Industrial Electronics, vol. 63, no. 7, pp. 4292-4303, Jul. 2016. [Baidu Scholar]
K. M. Alawasa and Y. A. R I. Mohamed, “Impedance and damping characteristics of grid-connected VSCs with power synchronization control strategy,” IEEE Transactions on Power Systems, vol. 30, no. 2, pp. 952-961, Mar. 2015. [Baidu Scholar]
S. Liao, M. Huang, X. Zha et al., “Emulation of multi-inverter integrated weak grid via interaction-preserved aggregation,” IEEE Journal of Emerging and Selected Topics in Power Electronics, vol. 9, no. 4, pp. 4153-4164, Aug. 2021. [Baidu Scholar]
J. Sun, “Frequency-domain stability criteria for converter-based power systems,” IEEE Open Journal of Power Electronics, vol. 3, pp. 222-254, Mar. 2022. [Baidu Scholar]
J. Shair, X. Xie, W. Liu et al., “Modeling and stability analysis methods for investigating subsynchronous control interaction in large-scale wind power systems,” Renewable and Sustainable Energy Reviews, vol. 135, p. 110420, Jan. 2021. [Baidu Scholar]
Y. Zhu, Y. Gu, Y. Li et al., “Impedance-based root-cause analysis: comparative study of impedance models and calculation of eigenvalue sensitivity,” IEEE Transactions on Power Systems, vol. 38, no. 2, pp. 1642-1654, Mar. 2023. [Baidu Scholar]
W. Zhou, Y. Wang, R. E. Torres-Olguin et al., “Effect of reactive power characteristic of offshore wind power plant on low-frequency stability,” IEEE Transactions on Energy Conversion, vol. 35, no. 2, pp. 837-853, Jun. 2020. [Baidu Scholar]
H. Zhang, X. Wang, M. Mehrabankhomartash et al., “Harmonic stability assessment of multiterminal DC (MTDC) systems based on the hybrid AC/DC admittance model and determinant-based GNC,” IEEE Transactions on Power Electronics, vol. 37, no. 2, pp. 1653-1665, Feb. 2022. [Baidu Scholar]