Next Article in Journal
Hydrothermal Carbonization of Digestate Produced in the Biogas Production Process
Next Article in Special Issue
Actions to Handle Increasing Distributed Generation and Rising Voltage during Low-Demand Periods: Tap Staggering Practices and Field Tests in the Italian Transmission Network
Previous Article in Journal
Electrical Transport Characteristics of Vertical GaN Schottky-Barrier Diode in Reverse Bias and Its Numerical Simulation
Previous Article in Special Issue
Unique Symbolic Factorization for Fast Contingency Analysis Using Full Newton–Raphson Method
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

DC Admittance Model of VSCs for Stability Studies in VSC-HVDC Systems

1
Department of Electrical Engineering (ETSEIB—UPC), Universitat Politècnica de Catalunya, Av. Diagonal 647, 08028 Barcelona, Spain
2
Department of Electrical Engineering (EPSEVG—UPC), Universitat Politècnica de Catalunya, Av. Victor Balaguer s/n, 08800 Vilanova i La Geltrú, Spain
*
Author to whom correspondence should be addressed.
Energies 2023, 16(14), 5457; https://doi.org/10.3390/en16145457
Submission received: 26 June 2023 / Revised: 13 July 2023 / Accepted: 16 July 2023 / Published: 18 July 2023
(This article belongs to the Special Issue Advanced Electric Power System 2023)

Abstract

:
High-voltage direct current (HVDC) systems linked to AC grids with converters are promising energy transmission systems. These systems present complex AC- and DC-side dynamic interactions. Impedance-based stability studies have recently been proposed to assess DC-side dynamics from DC-side characterization of voltage source converters (VSCs) considering AC-side dynamics. However, the existing approaches used for stability studies in VSC-HVDC systems do not completely model VSCs because they do not consider together the VSC delay, the grid voltage feedforward filter, and all the d- and q-reference current controls. Moreover, these approaches are analytically characterized from dq-real space vectors (less related to circuit theory than dq-complex space vectors), and some work with simple AC grids. The main contribution of this paper is a detailed and complete DC admittance model of VSCs from dq-complex space vectors, which considers the VSC delay, feedforward filter, and d- and q-reference current controls, and also a general AC grid. The proposed model can be used for DC-side stability studies in VSC-HVDC systems considering AC grid dynamics. The capabilities and drawbacks of impedance-based stability methods for DC-side stability assessment were analyzed, and the positive-net-damping criterion was validated as a robust approach. The model was validated by PSCAD/EMTDC simulations and applied to a stability study in a VSC-HVDC system.

1. Introduction

VSC-HVDC systems are emerging as a future transmission energy technology [1]. Near-synchronous [1,2,3] and harmonic [4,5] oscillatory instabilities due to low-damped resonances are a usual concern in these systems, and their accurate assessment is challenging for researchers. The state-space and frequency-domain methods are the most common approaches used to analyze these oscillatory phenomena [1,2,3,4,5,6,7,8,9,10]. The state-space method [3] is a global stability approach (i.e., stability is determined regardless of location [2]) but uses high-order dynamic models and requires specific information about all system elements, which is not always accessible. On the other hand, the frequency-domain methods assess system stability with less computing effort than the state-space method and are used either with simulations or system measurements. Broadly speaking, frequency-domain methods are applied using impedance-based stability criteria, such as the Nyquist criterion [4,5,6], impedance-based analysis [1], and the positive-net-damping stability criterion [4,5,6,7,8,9] for load–source equivalent systems, and the generalized Nyquist criterion (GNC) [2,6,10] and impedance matrix determinant analysis [3] for interconnected systems. Most of these criteria are local stability approaches (i.e., they are locally applied) sensitive to system partitions, which can lead to inaccurate stability predictions.
This weakness has been discussed in the literature to understand the proper application of the different criteria [2].
Stability studies in HVDC systems require DC-side equivalent modeling of converters. Early studies on VSC-HVDC systems considered weak and simple RL AC grids and a few VSC controls to obtain the DC-side model of VSCs [4,9]. This, in general, leads to inaccurate stability predictions because AC-side and VSC control dynamics are overlooked. Improved DC-side models of VSCs that consider AC-side dynamics have recently been proposed [11,12,13,14,15]. The first DC-side equivalent model for VSC rectifier and inverter stations of VSC-HVDC systems with active power and DC voltage controls, respectively, and grid alternating voltage control is derived from dq-real space vectors in [11]. The model disregards the VSC delay and phase-locked loop (PLL) control. Moreover, the study considers a simple RL impedance on the AC-side and wrongly concludes that AC-side instabilities are not detected on the DC side. A DC-side equivalent admittance model of a VSC-HVDC station in an offshore wind farm with DC voltage and PLL controls using dq-real space vectors is presented in [12]. Additionally, the modified sequence and phasor-domain models are derived from the previous model. The model disregards the VSC delay, grid voltage feedforward filter, and q-reference current controls. A general AC grid characterization to correctly analyze the influence of AC-side dynamics on DC-side instabilities (e.g., the mirror frequency effect [6,16]) is also considered in [12]. A DC impedance model of VSC rectifier and inverter stations of VSC-HVDC systems with active power and DC voltage controls, respectively, and PLL control is derived by using dq-real space vectors in [13,14,15]. The VSC delay and a general AC grid are also considered, but the grid voltage feedforward filter and the q-reference current controls are disregarded.
The above comments are summarized in Table 1. The first row (labeled “Phase-locked loop control”) shows the models in which the PLL control is included. The second row (labeled “VSC delay”) shows the models in which the VSC time delay transfer function is included. The third and fourth rows show the models which use voltage feedforward without and with a low-pass filter. The fifth and sixth rows show the models which include the outer control loops on channels d and q.
The influence of AC-side dynamics on stability and the stability assessment using the impedance-based stability criteria have been discussed. To better explain frequency coupling dynamics between the AC and DC sides (i.e., the mirror frequency effect), modified sequence [16] and dq-complex-domain [17] approaches have been proposed from the dq-real-domain model based on the Park (dq-real) and Ku (dq-complex) transformations [18]. Both dq-real- and dq-complex-domain approaches can be used for VSC modeling, but they have advantages and disadvantages for Laplace and frequency studies of grid-connected VSC stability:
  • dq-real-domain approach:
    Laplace stability studies: It allows system poles and participation factors to be easily obtained but is not appropriate for integrating components of large power systems. Moreover, frequency black-box models obtained from measurements cannot be used with this approach.
    Frequency stability studies: It is the most used in the literature [11,12,13,14,15] but reducing AC networks to their equivalent Thevenin circuit in dq-real variables is more laborious than in dq-complex variables. This is because AC-side impedances are 2 × 2 matrices with dq-real variables.
  • dq-complex-domain approach:
    Laplace stability studies: The characterization of system poles poses numerical problems because of rounding errors produced by the complex coefficients of dq-complex-domain transfer functions. Polynomials in the s-domain with only real coefficients are required in the DC impedance to correctly obtain the system poles.
    Frequency stability studies: It has recently been used in the literature [19,20]. In the dq-complex domain, the AC-side elements use the impedances ZTh(s + jω1) and ZTh (s − jω1) obtained by shifting the operational Thevenin impedance ZTh(s). The dq-complex domain is more appropriate for integrating VSC models in large power systems.
According to the above comments, the benefits of the dq-complex-domain approach over the dq-real-domain approach for grid impedance modeling and stability studies are greater.
This paper contributes to filling the above gaps in the characterization of the DC admittance model of VSCs. Moreover, it studies the inaccuracies of impedance-based stability criteria due to partitions in VSC-HVDC systems in two ways. The first contribution is a novel DC admittance model for VSCs based on dq-complex space vectors. The model extends existing models by considering the VSC delay, grid voltage feedforward filter, PLL, and main d-reference (active power and DC voltage controls) and q-reference (reactive power and grid alternating voltage controls) current outer loops, and also a general AC grid. The second contribution involves the study of VSC-HVDC system stability. The sensitivity of impedance-based stability criteria to VSC-HVDC system partitions is discussed, and the positive-net-damping stability criterion is proposed to overcome this issue. Finally, the influence of AC-side dynamics on the DC-side stability of VSC-HVDC systems was analyzed. The stability issues in a VSC-HVDC system were studied with the proposed DC admittance model to show the paper’s contribution. PSCAD/EMTDC simulations were performed to validate the study.

2. Model of VSCs in DQ-Complex Domain

Complex transfer functions are used in this section to model VSCs with dq-complex space vectors [6,16,17]. This approach is used in the next section to obtain the DC admittance of the VSCs.
The Laplace variable s will be omitted in the dq-complex transfer functions hereinafter to simplify notation, and the dq-complex space vectors are defined as Udq = (Ud + jUq)/√2 and Idq = (Id + jIq)/√2.

2.1. AC Grid Relations

The relation between voltage and current at the point of common coupling in Figure 1 is expressed as
Δ E d q Δ E d q * = Z T h + Z T h Z T h * Z T h + * Δ I d q Δ I d q * Δ E d q m = Z T h ± Δ I d q m ,
where Z T h + = ZTh(s + 1), ( Z T h + ) * = ZTh(s1) and Z T h = 0. The voltage balance across the converter RcLc filter [20] is expressed as
Δ V d q Δ V d q * = Z c + Z c Z c * Z c + * Δ I d q Δ I d q * + Δ E d q Δ E d q * Δ V d q m = Z c ± Δ I d q m + Δ E d q m ,
where Z c + = Rc + Lc(s + 1), ( Z c + ) * = Rc + Lc(s1), and Z c = 0.
Note that Z T h = Z c = 0 because the grid and the RcLc filter are three-phase symmetric systems whose impedance matrices in the dq-domain verify the relations Zdd(s) = Zqq(s) and Zdq(s) = −Zqd(s). Considering the above, the terms Z T h + and Z c + can also be expressed by using the coordinate transformation properties Z T h + = ZTh(s + 1) and Z c + = Zc(s + 1) = Rc + Lc(s + 1) [6].
The relation between VSC output voltage and AC current is derived from (1) and (2):
Δ V d q m = Z t ± Δ I d q m Z t ± = ( Z c ± + Z T h ± ) ,
where Z t + = Zt(s + 1) and Zt(s) = ZTh(s) + Rc + Lcs is the AC impedance.

2.2. DC-AC Converter Relations

The relation between the VSC output voltage and the DC voltage and the relation between the DC and AC currents are expressed [20] as
Δ V d q Δ V d q * = Δ m d q Δ m d q * V d c 0 + m d q 0 m d q 0 * Δ V d c Δ V d q m = Δ m d q m V d c 0 + m d q 0 m Δ V d c ,
Δ I d c = m d q 0 * m d q 0 Δ I d q Δ I d q * I d q 0 * I d q 0 Δ m d q Δ m d q * Δ I d c = ( m d q 0 m ) H Δ I d q m ( I d q 0 m ) H Δ m d q m ,
where mdq = (md + jmq)/√2 and Idq0 = (Id0 + jIq0)/√2, and the superscript H indicates the transpose and complex conjugate (Hermitian conjugate). Note that the three-phase set of modulation functions {ma, mb, mc} = M{cos(ω1t + ϕm), cos(ω1t + ϕm − 2π/3), cos(ω1t + ϕm + 2π/3)} is transformed into the dq-complex space vector mdq = (√3/2)Mϕm by applying the normalized Ku (dq-complex) transformation [6,18].

2.3. Simplified Model #1: Case without Current Control

In the case without current control, the modulation function is constant, Δ m d q * = 0 . Then, (4) and (5) result in
Δ V d q m = m d q 0 m Δ V d c Δ I d c = ( m d q 0 m ) H Δ I d q m .
By considering (3) and (6), the DC admittance is
Δ I d c = Y d c Δ V d c Y d c = ( m d q 0 m ) H ( Z t ± ) 1 m d q 0 m ,
where
Y d c = m d q 0 * m d q 0 Z t + 0 0 Z t + * 1 m d q 0 m d q 0 * = 3 M 2 4 1 Z t + + 1 Z t + * Z t + = Z t ( s + j ω 1 ) = Z T h ( s + j ω 1 ) + R c + L c ( s + j ω 1 ) Z t + * = Z t * ( s + j ω 1 ) = Z T h ( s j ω 1 ) + R c + L c ( s j ω 1 ) ,
and the DC impedance is the inverse of the DC admittance, Zdc = 1/Ydc.
The AC grid is represented by the circuit in Figure 2 [12], with the Thevenin transfer function written as
Z T h ( s ) = 1 1 R g + L g s + 1 Z f s ; Z f s = 1 C f s + 1 1 L f s + 1 R f .
Note that the frequency response of a conjugate transfer function such as (Z+t)*(s) in (8) can be calculated as [19].
Z * s = Z s * * Z * j ω = Z j ω * .

2.4. Simplified Model #2: Case with Current Control

Only the modulation function and current control of the VSC with delay are considered. Then, the current control loop [20] is expressed as
Δ V d q ,   r m   = F c c ± Δ I d q m Z ω ± Δ I d q m   + H f ± Δ E d q m ,
where F c c + = D(s)Fcc(s), Z ω + = D(sjLcω1, H f + = D(s)Hf(s), and F c c = Z ω = H f = 0 , with D(s) = esTd being the VSC time delay transfer function, Fscc(s) = ksp, cc + ksi, cc/s, and Hf(s) = αf /(s + αf), with ksp, cc, and ksi, cc being the proportional and integral gains of the current control, αf being the bandwidth of the voltage feedforward low-pass filter, and Td being the VSC time delay. The VSC reference voltage, ΔVmdq, r = Vdc0Δmmdq (11), is the dq-complex space vector determining the modulation function:
  Δ m d q m = 1 V d c 0 Z p i ± Δ I d q m   + H f ± Δ E d q m ; Z p i ± = F c c ± Z ω ± .
By considering (1) and (12), (5) becomes
Δ I d c = ( m d q 0 m ) H ( I d q 0 m ) H V d c 0 Z p i ±   H f ± Z T h ± Δ I d q m ,
and (4) can be expressed from (1), (2), and (12) as
Δ V d q m = Z c ± Δ I d q m + Δ E d q m = Z c ± + Z T h ± Δ I d q m = Z p i ±   H f ± Z T h ± Δ I d q m + m d q 0 m Δ V d c ,
which can be rewritten as
Δ I d q m = Z c ± Z T h ± Z p i ± + H f ± Z T h ± 1 m d q 0 m Δ V d c .
By replacing (15) in (13), the DC admittance function is
Δ I d c = Y d c Δ V d c Z c f ± = Z p i ± H f ± Z T h ± Y d c = ( m d q 0 m T ) * + ( I d q 0 m ) H V d c 0 Z c f ± ( Z t ± + Z c f ± ) 1 m d q 0 m .
The DC admittance in (16) can finally be expressed as
Y d c = 3 M 2 4 1 Z t + + Z c f + + 1 Z t + + Z c f + * + 1 2 V d c 0 ( I d 0 j I q 0 ) ( m d 0 + j m q 0 ) Z c f + Z t + + Z c f + + ( I d 0 + j I q 0 ) ( m d 0 j m q 0 ) ( Z c f + ) * ( Z t + ) * + Z c f + * ,
where
Z c f + = D ( s ) F c c ( s ) j L c ω 1 H f ( s ) Z T h ( s + j ω 1 ) ( Z c f + ) * = D ( s ) F c c ( s ) + j L c ω 1 H f ( s ) Z T h ( s j ω 1 ) ,
and the DC impedance is calculated with the inverse of the DC admittance, ZDC = 1/YDC. The DC impedance in the previous simplified models, (8) and (17), is numerically validated by PSCAD/EMTDC simulation.
Figure 3 shows these DC impedances (see data in Table 2). The case without current control, ZWC, calculated with (8), is plotted with a solid line whereas the frequency response obtained by PSCAD/EMTDC simulation is plotted with crosses. The DC impedance case with current control, ZCC, calculated with (17), is represented in the same way. The accurate results of the DC-side equivalent admittance obtained with the proposed model are worth noting.
In the case without current control, the coupling between the DC and AC sides is revealed by the two peak values of ZWC. The DC-impedance ZWC is affected by the mirror frequency effect. This is reflected in the expression of YDC (8) (see (1) in [12]), which depends on the impedances Zt(s + 1) and Zt(s1). As can be seen in the DC-impedance ZCC in Figure 3, the control damps this coupling effect between the DC and AC sides [13].

3. Detailed Model of DC Admittance of VSCs

The PLL and the outer loop relations are included in the detailed model, where the current control loop is modified to account for the PLL effect. Now, the current control loop relations must consider the relationship between the variables in the converter dq-domain, which has the superscript “c”, and the variables in the grid dq-domain, where no superscript is used. The current control loop [20] is expressed as
Δ V d q ,   r c m   = F c c ± Δ I d q ,   r c m   Δ I d q c m Z ω ± Δ I d q c m   + H f ± Δ E d q c m .
The dq-complex space vector of the VSC reference voltage is written as ΔVcmdq, r = Vdc0Δmcmdq. The PLL relations [9,20] are expressed as
Δ I d q c m = Δ I d q m + G P L L i ± Δ E d q m G P L L i + = G P L L i = G P L L ( s ) ( I d 0 + j I q 0 ) / 2 ,
Δ E d q c m = G P L L v ± Δ E d q m G P L L v + = 1 G P L L ( s ) ( E d 0 + j E q 0 ) / 2 G P L L v = 1 G P L L v + = G P L L ( s ) ( E d 0 + j E q 0 ) / 2 ,
Δ m d q c m = Δ m d q m + G P L L m ± Δ E d q m G P L L m + = G P L L m = G P L L ( s ) ( m d 0 + j m q 0 ) / 2 ,
where Gpll(s) = Fpll(s)/(s + Vd0Fpll(s)), Fpll(s) = kp, pll + ki, pll/s, and kp, pll and ki, pll are the proportional and integral gains of the PLL control.
The reference currents generated by the outer control loops are related to the DC voltage and the input AC voltage and current of the VSC with dq-complex space vectors as the following general expression:
Δ I d q ,   r c m = F o m Δ V d c G o ± Δ I d q c m Y o ± Δ E d q c m ,
where the complex transfer matrices Fmo, G±o, and Y±o are the complex transfer matrices depending on the VSC operation mode (see Appendix A).
The outer control loop (23) can be introduced into the inner current control loop (19) as
V d c 0 Δ m d q c m = Δ V d q ,   r c m   = F c c ± F o m Δ V d c + Z p i ± + F c c ± G o ± Δ I d q c m + H f ± + F c c ± Y o ± Δ E d q c m ,
where (24) is expressed in the converter dq-complex domain. By considering the PLL relations (20), (21), and (22) in the grid dq-domain (where no superscript is used), (24) becomes
Δ m d q m   = 1 V d c 0 F c c ± F o m Δ V d c + ( Z p i ± + F c c ± G o ± ) Δ I d q m + H o ± Δ E d q m H o ± = ( Z p i ± + F c c ± G o ± ) G p l l i ± + ( H f ± + F c c ± Y o ± ) G p l l v ± V d c 0 G p l l m ± ,
and introducing (1) into (25) results in
Δ m d q m   = 1 V d c 0 F c c ± F o m Δ V d c + ( Z p i ± + F c c ± G o ± H o ± Z T h ± ) Δ I d q m .
The DC voltage in (4) can be related to the AC current from (26) and the AC grid relations (1) and (2) as follows:
Δ V d q m = Z c ± + Z T h ± Δ I d q m = ( m d q 0 m + F c c ± F o m ) Δ V d c + ( Z p i ± + F c c ± G o ± H o ± Z T h ± ) Δ I d q m .
By using (1), (25) is rewritten as
G V Δ V d c = Z V Δ I d q m G V = m d q 0 m + F c c ± F o m Z V = Z t ± + Z a ± Z a ± = Z p i ± + F c c ± G o ± H o ± Z T h ± .
The DC current in (5) can be related to the AC current using (25) and the AC grid relations (1) and (2) as follows:
Δ I d c = ( m d q 0 m T ) * Δ I d q m ( I d q 0 m ) H V d c 0 F c c ± F o m Δ V d c + Z a ± Δ I d q m ,
which can be rewritten as
Δ I d c = Y I Δ V d c G I Δ I d q m Y I = ( I d q 0 m ) H V d c 0 F c c ± F o m G I = ( m d q 0 m T ) * + ( I d q 0 m ) H V d c 0 Z a ± .
The DC-admittance Ydc = ΔIdcVdc is obtained from the DC voltage and current relations (28) and (30), respectively, resulting in
Y d c = Δ I d c Δ V d c = Y I + G I Z V 1 G V = ( I d q 0 m ) H V d c 0 F c c ± F 0 m + ( m d q 0 m ) H   + ( I d q 0 m ) H V d c 0 Z a ± ( Z t ±   + Z a ± ) 1 ( m d q 0 m   + F c c ± F 0 m ) .
PSCAD/EMTDC simulations were performed to study the DC impedance calculation accuracy of different control modes. The parameters of each case are listed in Table 2. Figure 4 shows the DC impedance in the case of the outer control mode PQ, ZPQ and PEac, ZPE. Figure 5 shows the DC impedance in the case of the outer control mode VdcQ, ZUQ and VdcEac, ZUE. The simulation data are shown with x crosses in Figure 4 and Figure 5. The continuous curves are calculated from the inverse of (31) and they fit well with the simulation results.

4. Impedance-Based Stability Criteria

Once the DC-side equivalent admittance of VSCs is obtained, the general HVDC interconnected system in Figure 6a can be considered for stability studies. The incremental symbol Δ of the small-signal variables is omitted for the sake of simplicity. The external components connected to the DC grid buses are represented by their equivalent circuits as follows:
  • Grid-connected VSCs are characterized by the current balance at their Norton equivalent circuit buses:
I i = I d c ,   i Y d c ,   i ( s ) V i ( i = 1 ,     a ) ,
where the expression of Ydc, i(s) is the DC admittance in (31).
  • DC loads are characterized by their admittance transfer function Yi(s) (i = 1, … b).
  • Terminals without any connected external component are represented as open-circuit buses characterized by zero values of the current source and admittance of the Norton equivalent circuit, Ii = Yi(s) = 0 (i = 1, … c).
Several impedance-based stability criteria are proposed in the literature [1,2,3,4,5,6,7,8,9,10]. Some of the differences between these criteria are discussed in [2]. This discussion is extended in the sections below considering the general DC system in Figure 6a where Ydc, i(s) (31) is used to assess DC interconnected system stability. The discussion is also supported by Section 5.

4.1. Norton Admittance Matrix-Based Stability Assessment

The relations between voltages and currents at the HVDC transmission grid terminals in Figure 6a are expressed as
i = Y G ( s ) v i = i E Y E ( s ) v v = Y G ( s ) + Y E ( s ) 1 i E ,
where YG(s) is the admittance matrix of the HVDC transmission grid; iE and YE(s) are the current source and the diagonal admittance matrix of the external components, respectively; and v and i are the voltage and the injected current at the grid terminals, respectively.
Stability can be analyzed in the frequency domain by rewriting (33) as
v = I + Z G ( s ) Y E ( s ) 1 Z G ( s ) i E ,
where I is the identity matrix, and ZG(s) = YG−1(s) is the grid impedance matrix. The stability of the closed-loop system in (34) can be assessed by the GNC if and only if L(s) does not have a right-half-plane (RHP) pole. This criterion (called SC1 in this paper) extends the traditional Nyquist criterion to the Nyquist curves of the eigenvalues of the loop transfer function L(s) = ZG(s)YE(s) [2,6,10,12,13]. Note that the RHP pole condition fails when system instability comes from the AC side of any VSC because the corresponding VSC subsystem in the diagonal of YE(s) is unstable. This instability was not detected in this case and, obviously, was also not detected when AC grid dynamics were disregarded in the DC-side equivalent admittance modeling of the VSCs.

4.2. SISO-Based Stability Assessment

The SISO system in Figure 6b can be derived at the VSC terminals, where Ieq, i and Yeq, i(s) represent the current source and the admittance of the DC grid equivalent circuit observed from any VSC. The relation between the voltage Vi and the current sources is expressed as
V i = Z t ( s ) ( I d c ,   i + I e q ,   i ) ; Z t ( s ) = 1 Y e q ,   i ( s ) + Y d c ,   i ( s ) .
Two approaches are used to assess the stability of the SISO closed-loop system in (35):
  • Impedance-ratio stability criterion
Stability can be analyzed in the frequency domain by rewriting (35) as
V i = Z e q ,   i ( s ) 1 + Z e q ,   i ( s ) Y d c ,   i ( s ) ( I d c ,   i + I e q ,   i ) = Z T , i ( s ) I T ,   i .
If and only if the impedance ratio Zeq, i(sYdc, i(s) does not have a right-half-plane (RHP) pole, the stability of the closed-loop system in (36) can be assessed by the Nyquist stability criterion (denoted as SC in this paper), which evaluates the Nyquist curve of the impedance ratio [4,5,6,9,11,15]. This criterion can only be applied if instability comes from the interaction between the VSC and the DC grid equivalent circuit because the RHP pole condition fails when system instability comes from the AC side of the VSC or the DC grid equivalent circuit observed from the VSC.
  • Positive-net-damping stability criterion
The positive-net-damping stability criterion (denoted as SC3 in this paper) evaluates system transfer impedance damping in (36) at resonance frequencies, i.e., the value of Re{ZT, i()} at the frequencies of the peak values of ZT, i() [4,5,6,7,8,9]. This criterion assesses the closed-loop transfer function of the system, which avoids inaccuracies derived from local instabilities in the stability conclusion.

5. Application

The stability in the VSC-HVDC system of Figure 7a was studied. Two converter stations, VSC1 and VSC2, were connected through a 1000 MW 440 kV HVDC transmission grid. VSC1 operated in mode VdcQ, controlling the DC voltage of the HVDC transmission grid, whereas VSC2 operated in mode PQ, controlling the transmitted active power. Both converters worked at unity power factor, typical of normal operating conditions [8,12,13,14,15]. According to Figure 7b, VSC1 is characterized by the DC-side equivalent admittance in (31), while VSC2 is considered as constant power consumption P20 characterized by the fictive resistance R20 = −(Vdc2, 0)2/P20 [4,9]. This converter could also be characterized with the DC-side equivalent admittance in (31), but a simpler model was used in the stability study to focus the attention on VSC1. VSC-HVDC system data are in Table 3. Three cases were studied in the application:
  • Case #1 (stable reference case): this corresponds to the data in Table 3, with the power consumed by VSC2 being half the nominal power (i.e., P20 = 500 MW).
  • Cases #2 and #3: the influence of the short-circuit ratio SCR and the VSC2-consumed power P20, respectively, on system stability were analyzed.
The system stability was verified by PSCAD/EMTDC time-domain simulations, and the predictions obtained by the stability criteria SC1, SC2, and SC3 in Section 4 were investigated. The results are shown in Figure 8 and Figure 9.
Case #1: The PSCAD/EMTDC simulations before SCR and P20 changes in Figure 8 show the system stability. The results obtained from the stability criteria are as follows (see Figure 9, left):
  • SC1: the GNC verifies system stability because the curves of the eigenvalues of L(s) do not encircle the −1 point.
  • SC2: the Nyquist criterion verifies system stability because the Nyquist curve of Zeq,iYdc, i does not encircle the −1 point.
  • SC3: The positive-net-damping stability criterion verifies system instability because the damping Re{ZT()} at resonance frequencies fr ≈ 537.5, 712.3 and 810.3 Hz is positive. The first resonance is due to the HVDC grid capacitors and inductors whereas the second and third resonances are at the DC mirror resonance frequencies of the AC-side resonance between the AC grid inductance Lg and the VSC1 filter capacitor Cf at fr = 760 Hz. It is observed that the VSC control damps (but does not eliminate) the AC-side resonance at the DC side. This means that AC-side dynamics are only damped at the DC side by the VSC control and should therefore be considered in the DC-side equivalent admittance modeling of the VSCs [12,13].
Case #2: The short-circuit ratio SCR increases from 15 pu to 30 pu at 0.1 s. The PSCAD/EMTDC simulation in Figure 8 shows that the system becomes unstable after the SCR value steps up. It was numerically verified that this is because the AC system resonance frequency is shifted to a negative damping frequency range and the AC-side voltage oscillations at f ≈ 980 Hz are amplified. The results obtained by the stability criteria are as follows (see Figure 9, middle):
  • SC1: the GNC leads to an inaccurate stability assessment (i.e., the curves of the eigenvalues of L(s) do not encircle the −1 point) because the instability comes from the AC system resonance, causing the DC-admittance Ydc, 1(s) in the diagonal of YE (s) to be unstable.
  • SC2: the Nyquist criterion also leads to an inaccurate stability assessment because Ydc, 1(s) is unstable [13].
  • SC3: The positive-net-damping stability criterion verifies system stability (i.e., Re{ZT()} is negative at mirror resonance frequencies fr ≈ 935.5 and 1045 Hz). Closed-loop transfer function assessment by the positive-net-damping stability criterion avoids the SC1 and SC2 stability inaccuracy derived from the instability of Ydc, 1(s). If AC-side dynamics are disregarded in the DC-side equivalent admittance model, AC-side resonances are overlooked, which leads to inaccurate stability assessments [12,13].
Case #3: The VSC2-consumed active power increases from P20 = 500 MW to the nominal value P20 = 1000 MW at 0.3 s. The PSCAD/EMTDC simulation in Figure 8 shows that the system becomes unstable after the P20 value ramps up. It was numerically verified that this is because the damping becomes negative at the DC resonance frequency, and the DC-side voltage oscillations at f ≈ 535 Hz are amplified. The results obtained by the stability criteria are as follows (see Figure 9, right):
  • SC1: the GNC verifies system instability because the curve of one eigenvalue intersects the unit circle at approximately 527 Hz, enclosing the −1 point in a clockwise direction.
  • SC2: the Nyquist criterion leads to an inaccurate stability assessment because of instability in Yeq, 1(s).
  • SC3: The positive-net-damping stability criterion verifies system instability (i.e., Re{ZT()} is negative at fr ≈ 537 Hz). The closed-loop transfer function assessment in the positive-net-damping stability criterion avoids the SC2 stability inaccuracy derived from the instability of Yeq, 1(s).

6. Comparison of DC Admittance Models

The influence of the different simplifications that are usually carried out in DC admittance models is shown in Table 4. The results obtained with the full model in Section 3 for the three cases in Section 4 were compared with those obtained considering five different simplifications. It can be observed that the simplification of the d-outer control loop and the PLL do not affect the results of the full model, while this is not true for the simplifications related to the inner control loop (i.e., the VSC delay and the bandwidth of the grid voltage feedforward simplifications). Nevertheless, it must be noted that the above conclusions correspond to the harmonic instabilities of the studied cases, and they cannot be extrapolated to other types of instabilities. In general, harmonic instabilities are mainly affected by the inner control loop parameters, while synchronous and subsynchronous instabilities are affected by the outer control loops and the PLL.

7. Conclusions

This paper presents a complete and detailed DC admittance model for VSC rectifier and inverter stations of VSC-HVDC systems with active power, DC voltage, grid alternating voltage, and reactive power controls. The model also considers the VSC delay, grid voltage feedforward filter, PLL, and several outer control loops. It is derived from dq-complex space vectors commonly used in circuit theory. A comparison between the different models in the literature and a study of the impact on the instability of the different model simplifications were carried out to illustrate the contribution of the paper. The proposed model was applied to discuss VSC-HVDC system stability and the inaccuracies of impedance-based stability criteria due to system partitions. It is concluded that the stability assessment of VSC-HVDC systems from VSC and DC grid equivalent circuits could lead to inaccurate conclusions by the Nyquist criterion (e.g., when instability comes from AC-side dynamics) whereas the positive-net-damping criterion overcomes this issue because it is a global stability criterion which evaluates system transfer impedance damping at resonance frequencies.

Author Contributions

Conceptualization and methodology, J.P. and L.S.; software development and validation, L.M. and J.P.; writing—original draft preparation, J.P. and L.S.; visualization, L.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the MCIN (Proyectos de Generación de Conocimiento) under grant number PID2021-123633OB-C33 (J-02895).

Data Availability Statement

Data sharing not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A. Outer Loop Controls

The VSC control is characterized by the current controller (CC) and the outer control loops, which make it possible to characterize the main VSC operation modes (i.e., mode PQ, mode VdcQ, mode VdcEac, and mode PEac) [8].
VdcQ control: the direct voltage control (DVC) and reactive power control (QPC) of Figure 1 are expressed with dq-complex space vectors [20] using the notation of (23) as
F o ,   U Q m = 1 2 F d c ( s ) F d c ( s ) ; G o ,   U Q ± = F q ( s ) E d 0 2 1 1 1 1 ; Y o ,   U Q ± = F q ( s ) 2 I d q 0 * I d q 0 I d q 0 * I d q 0 ,
where Fdc(s) = kp, dc + ki, dc/s, with kp, dc and ki, dc being the proportional and integral gains of the DVC, and Fp(s) = kp, q + ki, q/s, with kp, q, and ki, q being the proportional and integral gains of the QPC.
PQ control: the alternating power control (APC) and reactive power control of Figure 1 are expressed with dq-complex space vectors [20] using the notation of (23) as
G o ,   P Q ± = F p ( s ) E d 0 1 0 0 1 ; Y o ,   P Q ± = 2 F p ( s ) 0 I d q 0 I d q 0 * 0 ,
where Fp(s) = kp, p + ki, p/s, with kp, p = kp, q and ki, p = ki, q being the proportional and integral gains of the APC.
VdcEac control: the direct voltage control (DVC) and alternating voltage control (AVC) control law of Figure 1 are expressed with dq-complex space vectors using the notation of (23) as
F o ,   U E m = 1 2 F d c ( s ) F d c ( s ) ; Y o ,   U E ± = F v ( s ) 2 j j j j ,
where Fv(s) = kp, v + ki, v/s, with kp, v and ki, v being the proportional and integral gains of the AVC.
PEac control: the alternating power control (APC) and alternating voltage control (AVC) control law of Figure 1 are expressed with dq-complex space vectors using the notation of (23) as
G o ,   P E ± = F p ( s ) E d 0 2 1 1 1 1 ; Y o ,   P E ± = F p ( s ) 2 I d q 0 * I d q 0 I d q 0 * I d q 0 + F v ( s ) 2 j j j j .

References

  1. Sun, J.; Li, M.; Zhang, Z.; Xu, T.; He, J.; Wang, H.; Li, G. Renewable energy transmission by HVDC across the continent: System challenges and opportunities. CSEE J. Power Energy Syst. 2017, 3, 353–364. [Google Scholar] [CrossRef]
  2. Zhang, C.; Molinas, M.; Rygg, A.; Cai, X. Impedance-Based Analysis of Interconnected Power Electronics Systems: Impedance Network Modeling and Comparative Studies of Stability Criteria. IEEE J. Emerg. Sel. Top. Power Electron. 2020, 8, 2520–2533. [Google Scholar] [CrossRef] [Green Version]
  3. Liu, H.; Xie, X.; Liu, W. An Oscillatory Stability Criterion Based on the Unified dq-Frame Impedance Network Model for Power Systems with High-Penetration Renewables. IEEE Trans. Power Syst. 2018, 33, 3472–3485. [Google Scholar] [CrossRef]
  4. Stamatiou, G.; Bongiorno, M. Stability Analysis of Two-Terminal VSC-HVDC Systems Using Net-Damping Criterion. IEEE Trans. Power Deliv. 2016, 31, 1748–1756. [Google Scholar] [CrossRef]
  5. Sainz, L.; Cheah-Mane, M.; Monjo, L.; Liang, J.; Gomis-Bellmunt, O. Positive-net-damping Stability Criterion in Grid-connected VSC Systems. IEEE J. Emerg. Sel. Top. Power Electron. 2017, 4, 1499–1512. [Google Scholar] [CrossRef]
  6. Harnefors, L. Modeling of Three-Phase Dynamic Systems Using Complex Transfer Functions and Transfer Matrices. IEEE Trans. Ind. Electron. 2007, 54, 2239–2248. [Google Scholar] [CrossRef]
  7. Harnefors, L. Proof and Application of the Positive-net-damping Stability Criterion. IEEE Trans. Power Syst. 2011, 26, 481–482. [Google Scholar] [CrossRef]
  8. Harnefors, L.; Bongiorno, M.; Lundberg, S. Input-admittance Calculation and Shaping for Controlled Voltage-source converters. IEEE Trans. Ind. Electron. 2007, 54, 3323–3334. [Google Scholar] [CrossRef]
  9. Wen, B.; Boroyevich, D.; Burgos, R.; Mattavelli, P.; Shen, Z. Analysis of D-Q Small-Signal Impedance of Grid-Tied Inverters. IEEE Trans. Power Electron. 2016, 31, 675–687. [Google Scholar] [CrossRef]
  10. Pinares, G.; Bongiorno, M. Analysis and Mitigation of Instabilities Originated from DC-side Resonances in VSC-HVDC Systems. IEEE Trans. Ind. Appl. 2016, 52, 2807–2815. [Google Scholar] [CrossRef]
  11. Xu, L.; Fan, L.; Miao, Z. DC Impedance-model Based Resonance Analysis of a VSC-HVDC System. IEEE Trans. Power Deliv. 2015, 30, 1221–1230. [Google Scholar] [CrossRef]
  12. Shah, S.; Parsa, L. Impedance Modeling of Three-phase Voltage Source Converters in dq, Sequence and Phasor Domains. IEEE Trans. Energy Convers. 2017, 32, 1139–1150. [Google Scholar] [CrossRef]
  13. Amin, M.; Molinas, M. Small-signal Stability Assessment of Power Electronics Based Power Systems: A Discussion of Impedance- and Eigenvalue-based Methods. IEEE Trans. Ind. Appl. 2017, 53, 5014–5030. [Google Scholar] [CrossRef]
  14. Amin, M.; Molinas, M.; Lyu, J.; Cai, X. Impact of Power Flow Direction on the Stability of VSC-HVDC Seen from the Impedance Nyquist Plot. IEEE Trans. Power Electron. 2017, 32, 8204–8217. [Google Scholar] [CrossRef]
  15. Liu, J.; Tao, X.; Yu, M.; Xia, Y.; Wei, W. Impedance Modeling and Analysis of Three-Phase Voltage-Source Converters Viewing from DC Side. IEEE J. Emerg. Sel. Top. Power Electron. 2020, 8, 3906–3916. [Google Scholar] [CrossRef]
  16. Rygg, A.; Molinas, M.; Zhang, C.; Cai, X. A Modified Sequence-domain Impedance Definition and its Equivalence to the dq-domain Impedance Definition for the Stability Analysis of AC Power Electronic Systems. IEEE J. Emerg. Sel. Top. Power Electron. 2016, 4, 1383–1396. [Google Scholar] [CrossRef] [Green Version]
  17. Wang, X.; Harnefors, L.; Blaabjerg, F. Unified Impedance Model of Grid-connected Voltage-source Converters. IEEE Trans. Power Electron. 2018, 33, 1775–1785. [Google Scholar] [CrossRef] [Green Version]
  18. Paap, G.C. Symmetrical Components in the Time Domain and their Application to Power Network Calculations. IEEE Trans. Power Syst. 2000, 15, 522–528. [Google Scholar] [CrossRef] [Green Version]
  19. Harnefors, L.; Wang, X.; Chou, S.; Bongiorno, M.; Hinkkanen, M.; Routimo, M. Asymmetric Complex-Vector Models With Application to VSC–Grid Interaction. IEEE J. Emerg. Sel. Top. Power Electron. 2020, 8, 1911–1921. [Google Scholar] [CrossRef]
  20. Pedra, J.; Sainz, L.; Monjo, L. Three-port small-signal admittance-based model of VSC for studies of multi-terminal HVDC hybrid AC/DC transmission grids. IEEE Trans. Power Syst. 2021, 36, 732–743. [Google Scholar] [CrossRef]
Figure 1. Grid-connected VSC system for DC impedance characterization.
Figure 1. Grid-connected VSC system for DC impedance characterization.
Energies 16 05457 g001
Figure 2. AC grid connected at the VSC terminals [12].
Figure 2. AC grid connected at the VSC terminals [12].
Energies 16 05457 g002
Figure 3. DC-impedance in the case without control, ZWC, and with current control, ZCC. (Lines: inverse of (8) and (17). Cross: PSCAD/EMTDC simulations.)
Figure 3. DC-impedance in the case without control, ZWC, and with current control, ZCC. (Lines: inverse of (8) and (17). Cross: PSCAD/EMTDC simulations.)
Energies 16 05457 g003
Figure 4. DC-impedance of P-Q and P-E control modes. (Lines: inverse of (31). Cross: PSCAD/EMTDC simulations.)
Figure 4. DC-impedance of P-Q and P-E control modes. (Lines: inverse of (31). Cross: PSCAD/EMTDC simulations.)
Energies 16 05457 g004
Figure 5. DC-impedance of Vdc-Q and Vdc-E control modes. (Lines: inverse of (31). Cross: PSCAD/EMTDC simulations.)
Figure 5. DC-impedance of Vdc-Q and Vdc-E control modes. (Lines: inverse of (31). Cross: PSCAD/EMTDC simulations.)
Energies 16 05457 g005
Figure 6. Stability assessment in HVDC transmission grids: (a) HVDC transmission grid. (b) SISO equivalent system partitioned at VSC terminals.
Figure 6. Stability assessment in HVDC transmission grids: (a) HVDC transmission grid. (b) SISO equivalent system partitioned at VSC terminals.
Energies 16 05457 g006
Figure 7. HVDC application: (a) HVDC transmission system. (b) HVDC transmission grid with VSC equivalent admittances. (c) Equivalent HVDC circuit.
Figure 7. HVDC application: (a) HVDC transmission system. (b) HVDC transmission grid with VSC equivalent admittances. (c) Equivalent HVDC circuit.
Energies 16 05457 g007
Figure 8. Stability study: Case #2 (top) and Case #3 (bottom).
Figure 8. Stability study: Case #2 (top) and Case #3 (bottom).
Energies 16 05457 g008
Figure 9. Impedance-based stability study.
Figure 9. Impedance-based stability study.
Energies 16 05457 g009
Table 1. Comparison of models.
Table 1. Comparison of models.
References
Model Feature [11][12][13][14][15]Proposed
Model
Phase-locked loop controlEnergies 16 05457 i001Energies 16 05457 i002Energies 16 05457 i002Energies 16 05457 i002Energies 16 05457 i001Energies 16 05457 i002
VSC delayEnergies 16 05457 i001Energies 16 05457 i001Energies 16 05457 i002Energies 16 05457 i002Energies 16 05457 i001Energies 16 05457 i002
Grid voltage feedforwardEnergies 16 05457 i001Energies 16 05457 i001Energies 16 05457 i002Energies 16 05457 i002Energies 16 05457 i002Energies 16 05457 i002
Grid voltage feedforward filterEnergies 16 05457 i001Energies 16 05457 i001Energies 16 05457 i001Energies 16 05457 i001Energies 16 05457 i001Energies 16 05457 i002
D-outer control loopEnergies 16 05457 i002Energies 16 05457 i002Energies 16 05457 i002Energies 16 05457 i002Energies 16 05457 i002Energies 16 05457 i002
Q-outer control loopEnergies 16 05457 i001Energies 16 05457 i001Energies 16 05457 i001Energies 16 05457 i001Energies 16 05457 i002Energies 16 05457 i002
Table 2. AC Grid and VSC Parameters.
Table 2. AC Grid and VSC Parameters.
Vga12.25/√2 kVPvsc30 MWRf, Lf2.3 Ω, 0.18 mH
Vdc030 kVRg, Lg0.52 Ω, 7.8 mH Rc, Lc0, 4 mH
Cf35 µF fsw, Td10 kHz, 0.3 ms Id0, Iq0−1996.3 A, 0 A
αf106.18 s−1md00.5168 pumq00.0836 pu
kp,cc2 pukp,dc0.8662 pukp,pll0.2757 pu
ki,cc0.096 puki,dc0.0334 puki,pll0.2203 pu
kp,ac0.9 pukp,P0.13 pukp,Q0.13 pu
ki,ac0.008 puki,P0.0083 puki,Q0.0083 pu
Table 3. VSC-HVDC System Data (UB, AC = 220 kV, UB, DC = 440 kV, SB = 1000 MVA).
Table 3. VSC-HVDC System Data (UB, AC = 220 kV, UB, DC = 440 kV, SB = 1000 MVA).
AC main gridU0 (f1), Xg/Rg
SCR = Scc/(Pvsc1, N)
220 kV (50 Hz), 20 pu
15 pu
VSC AC filterCf6 mF
DC cableRdc, Ldc10.8 mΩ/km, 0.149 mH/km
Cdc, Length0.145 mF/km, 25 km
VSC DC filtersCcvi (i = 1 to 2)30 mF
VSC1Pvsc1, N, Vdc1, 01000 MW, 440 kV
VSC1 controlfs, Td2 kHz, 0.3 ms
Rc, Lc0.025 pu, 0.25 pu
kp,cc, ki,cc1.21 pu, 0.121 pu
kp,pll, ki,pll0.48 pu, 0.031pu
kp,dc, ki,dc0.15 pu, 0.058 pu
kp,p, ki,p0.15 pu, 0.058 pu
Table 4. Comparison of simplifications.
Table 4. Comparison of simplifications.
Cases
Model Feature Case #1Case #2Case #3
Full modelStableInstableInstable
No phase-locked loop controlStableInstableInstable
No VSC delayStableStableStable
No grid voltage feedforward (αf = 0)StableStableInstable
No grid voltage feedforward filter (αf = ∞)StableInstableInstable
No d-outer control loopStableInstableInstable
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pedra, J.; Sainz, L.; Monjo, L. DC Admittance Model of VSCs for Stability Studies in VSC-HVDC Systems. Energies 2023, 16, 5457. https://doi.org/10.3390/en16145457

AMA Style

Pedra J, Sainz L, Monjo L. DC Admittance Model of VSCs for Stability Studies in VSC-HVDC Systems. Energies. 2023; 16(14):5457. https://doi.org/10.3390/en16145457

Chicago/Turabian Style

Pedra, Joaquín, Luis Sainz, and Lluís Monjo. 2023. "DC Admittance Model of VSCs for Stability Studies in VSC-HVDC Systems" Energies 16, no. 14: 5457. https://doi.org/10.3390/en16145457

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop