Skip to main content
Log in

Computation of the controllable swing mode spectrum of FACTS and HVDC in large power systems

  • Published:
Sādhanā Aims and scope Submit manuscript

Abstract

This paper presents computation of swing modes of a large power system that could be significantly affected by power swing damping controllers in FACTS or HVDC devices at a given location. Modal controllability is a suitable measure to isolate these modes for analysis. Computation of the controllable swing mode spectrum is useful, especially in situations where the controller structure and feedback signals are not frozen (e.g., at the planning stage). This paper proposes two important steps that allow us to map the problem of finding highly controllable swing modes to the problem of finding the swing modes that have high transfer function residues (for which efficient algorithms are available). The steps are: (a) normalization of the eigenvectors corresponding to different modes and (b) identification of specific feedback signals for each type of FACTS/HVDC device such that the modal observability and modal controllability are tightly coupled. Once the mapping is done, a computationally efficient method like the Subspace Accelerated Dominant Pole Algorithm [16] (SADPA) can be adapted to find the highly controllable swing modes. The effectiveness of this approach is demonstrated by case studies of FACTS and HVDC devices in a 16-machine system and the Indian power grid.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Figure 1
Figure 2
Figure 3
Figure 4
Figure 5
Figure 6
Figure 7
Figure 8

Similar content being viewed by others

Notes

  1. For complex poles, the complex–conjugate pair is considered as one mode.

  2. G(s) is calculated using Eq. (16) for a frequency range of 1.26–19 rad/s in steps of 0.01 rad/s.

References

  1. Padiyar K R 2007 FACTS controllers in power transmission and distribution. New Delhi: New Age International(P) Ltd

    Google Scholar 

  2. Padiyar K R 2012 HVDC power transmission systems, 2nd ed. New Delhi: New Age International(P) Ltd

    Google Scholar 

  3. Martins N, Lima L T G and Pinto H J C P 1996 Computing dominant poles of power system transfer functions. IEEE Trans. Power Syst. 11(1): 162–167

    Article  Google Scholar 

  4. Uchida N and Nagao T 1988 A new eigen-analysis method of steady-state stability studies for large power systems: Smatrix method. IEEE Trans. Power Syst. 3(2): 2085–2092

    Article  Google Scholar 

  5. Semlyen A and Wang L 1988 Sequential computation of the complete eigen system for the study zone in small signal stability analysis of large power systems. IEEE Trans. Power Syst. 3(2): 715–725

    Article  Google Scholar 

  6. Wang L and Semlyen A 1989 Application of sparse eigenvalue techniques to the small signal stability analysis of large power systems. In: Proceedings of the Power Industry Computer Application Conference, pp. 358–365

  7. Angelidis G and Semlyen A 1995 Efficient calculation of critical eigenvalue clusters in the small signal stability analysis of large power systems. IEEE Trans. Power Syst. 10(1): 427–432

    Article  Google Scholar 

  8. Yang D and Ajjarapu V 2007 Critical eigenvalues tracing for power system analysis via continuation of invariant subspaces and projected Arnoldi method. IEEE Trans. Power Syst. 22(1): 324–332

    Article  Google Scholar 

  9. Li Y, Geng G and Jiang Q 2016 An efficient parallel Krylov–Schur method for eigen-analysis of large-scale power system. IEEE Trans. Power Syst. 31(2): 920–930

    Article  Google Scholar 

  10. Rommes J and Martins N 2008 Computing large-scale system eigenvalues most sensitive to parameter changes, with applications to power system small-signal stability. IEEE Trans. Power Syst. 23(2): 434–442

    Article  Google Scholar 

  11. Chung C Y and Dai B 2013 A combined TSA–SPA algorithm for computing most sensitive eigenvalues in large-scale power systems. IEEE Trans. Power Syst. 28(1): 149–157

    Article  Google Scholar 

  12. Chung C Y and Dai B 2015 A generalized approach for computing most sensitive eigenvalues with respect to system parameter changes in large-scale power systems. IEEE Trans. Power Syst. pp. 1–11, https://doi.org/10.1109/TPWRS.2015.2445792

    Article  Google Scholar 

  13. Byerly R T, Bennon R J and Sherman D E 1982 Eigenvalue analysis of synchronizing power flow oscillations in large electric-power systems. IEEE Trans. Power Appl. Syst. PAS-101: 235–243

    Article  Google Scholar 

  14. Wong D Y, Rogers G J, Porretta B and Kundur P 1988 Eigenvalue analysis of very large power systems. IEEE Trans. Power Syst. 3(2): 472–480

    Article  Google Scholar 

  15. Martins N 1997 The dominant pole spectrum eigensolver. IEEE Trans. Power Syst. 12(1): 245–254

    Article  Google Scholar 

  16. Rommes J and Martins N 2006 Efficient computation of transfer function dominant poles using subspace acceleration. IEEE Trans. Power Syst. 21(3): 1471–1483

    Article  Google Scholar 

  17. Mhaskar U P and Kulkarni A M 2006 Power oscillation damping using FACTS devices: modal controllability, observability in local signals, and location of transfer function zeros. IEEE Trans. Power Syst. 21(1): 285–294

    Article  Google Scholar 

  18. Rogers G 2000 Power system oscillations. Norwell, MA: Kluwer

    Book  Google Scholar 

  19. Rommes J 2007 Methods for eigenvalue problems with applications in model order reduction. PhD Thesis, Utrecht University, Utrecht, http://dspace.library.uu.nl/handle/1874/21787 [Accessed 28 October 2016]

  20. Shubhanga K N and Anatholla Y 2000 Manual for a multi-machine small-signal stability programme (Version 1.0). Department of Electrical Engineering, NITK Surathkal, Karnataka, India, p 54

  21. EIGS. http://www.mathworks.com/help/matlab/ref/eigs.html [Accessed 28 October 2016]

  22. Latorre H F, Ghandhari M and Soder L 2008 Active and reactive power control of VSC-HVDC. Electr. Power Syst. Res. 78: 1756–1763

    Article  Google Scholar 

  23. Power System Operation Corporation Ltd. POC data, https://posoco.in/transmission-pricing/poc-data [Accessed 28 October 2016]

  24. Central Electricity Authority manual on transmission line planning criteria, http://cea.nic.in [Accessed 28 October 2016]

  25. Kundur P 1994 Power systems stability and control. New York: McGraw-Hill

    Google Scholar 

Download references

Acknowledgements

The authors gratefully acknowledge the contributions of Mr Abishek R S and Mr A Sinkar in preparation of the small signal model of the Indian Power System. Both of them have received their Master’s degrees from the Department of Electrical Engineering, IIT Bombay.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Vedanta Pradhan.

Appendices

Appendix I. Derivation of constraint in Eq. (12)

Following a perturbation, the energy in the disturbance can be expressed as sum of potential and kinetic energy as follows:

$$\begin{aligned} E= & {} \frac{1}{2}(\Delta \delta ^TA_r\Delta \delta \;+\; \Delta \omega ^TM\Delta \omega ) \nonumber \\= & {} \frac{1}{2}\begin{bmatrix} \Delta \delta \\ \Delta \omega \end{bmatrix}^T\begin{bmatrix} A_r&[0]\\ [0]&M \end{bmatrix}\begin{bmatrix} \Delta \delta \\ \Delta \omega \end{bmatrix}. \end{aligned}$$
(A1)

It is constant for the unforced system (i.e., \(u = 0\)). If the \(i{\hbox {th}}\) mode only is excited, then from Eq. (5) we have

$$\begin{aligned} \begin{bmatrix} \Delta \delta \\ \Delta \omega \end{bmatrix}= \begin{bmatrix} x_{\delta _i}\\x_{\omega _i} \end{bmatrix}z_{m_i} + \begin{bmatrix} x_{\delta _i}^*\\x_{\omega _i}^* \end{bmatrix}z_{m_i}^*. \end{aligned}$$
(A2)

Using Eq. (A2) in Eq. (A1), it can be shown that

$$\begin{aligned} E= & {} \frac{1}{2}[z_{m_i}^2(x_{\delta _i}^TA_rx_{\delta _i}+x_{\omega _i}^TMx_{\omega _i}) + (z_{m_i}^*)^2(x_{\delta _i}^{H}A_rx_{\delta _i}^*+x_{\omega _i}^{H}Mx_{\omega _i}^*) \nonumber \\&\quad +\, 2|z_{m_i}|^2(x_{\delta _i}^{H}A_rx_{\delta _i}+x_{\omega _i}^{H}Mx_{\omega _i})]. \end{aligned}$$
(A3)

From Eqs. (3) and (4), it can be shown that

$$\begin{aligned}&x_{\delta _i}^TA_rx_{\delta _i}+x_{\omega _i}^TMx_{\omega _i} = 0 {\text {,}} \end{aligned}$$
(A4)
$$\begin{aligned}&x_{\delta _i}^{H}tA_rx_{\delta _i}^*+x_{\omega _i}^{H}Mx_{\omega _i}^* = 0, \end{aligned}$$
(A5)

and therefore

$$\begin{aligned} E = |z_{m_i}|^2\;(x_{\delta _i}^{H}A_rx_{\delta _i}+x_{\omega _i}^{H}Mx_{\omega _i}) = E_i \end{aligned}$$
(A6)

where \(E_i\) denotes the energy when only the \(i{\hbox {th}}\) mode is excited. It can be shown that when several modes are excited, the disturbance energy is given by the sum of all modal energies, i.e, \(E = \sum _{i = 1}^{(n_g-1)} E_i\). Note that \(E_i\) is also a constant and does not depend on scaling of the eigenvectors.

The strength of the modal signal \(y = v_i^Tz = z_{m_i}\) is given by its amplitude, which is equal to \(|z_{m_i}|\). When only one mode is excited at a time, with the same disturbance energy \(E_i\), strength of the modal signal being equal across modes implies \((x_{\delta _i}^{H}A_rx_{\delta _i}+x_{\omega _i}^{H}Mx_{\omega _i})\) is the same across all modes (as seen from Eq. (A6)). Without loss of generality, we assign its value to be 1 and obtain the following condition:

$$\begin{aligned} x_{\delta _i}^{H}A_rx_{\delta _i}+x_{\omega _i}^{H}Mx_{\omega _i} = 1. \end{aligned}$$
(A7)

Using the initial assumption of \(x_{\delta _i}\) being real, we obtain the result of Eq.  (12).

Appendix II. Example of dominant pole at infinity

Let us consider the state-space matrices \(E = \begin{bmatrix} I&0\\0&0 \end{bmatrix},\; A = \begin{bmatrix} \alpha&0\\ 0&\alpha \end{bmatrix} {\text {where}}\; \alpha = \begin{bmatrix} 0.1&0.2\\ -\,0.2&0.3 \end{bmatrix}\) and \(b = c = [0.1\;0.1\;1.0\;1.0]^T\). Generalized eigenvalue analysis of this system yields a mode \(0.2 \pm 0.17321i\) with residue of 0.01 and two poles at infinity with residue of \(\infty \).

Trajectories of the eigenvalue estimate over iterations, when SADPA is applied on this system with \(|\rho _i|\) (or \(|\rho _i||\lambda _i|\)) (case (a)), and \(|\rho _i|/|\lambda _i|\) (case (b)) as selection criteria separately are shown in the following table. It is clearly seen that the selection criterion in case (b) avoids convergence to the dominant pole at \(\infty \).

Iteration

Eigen-estimate

Case (a)

Case (b)

0

0 +j 0.1700

0 +j 0.1700

1

25.2768 –j 5.1847

25.2768 –j 5.1847

2

(2.72 +j 3.50)\(\times 10^5\)

0.2944 +j 0.1997

3

\(\infty \)

0.2000 –j 0.1732

Abcd:

state-space matrices

z :

state variables

\(\delta , \omega \) :

rotor angle (rad), speed (rad/s)

\(n_g\) :

number of generators

\(\Omega _i\) :

frequency of the \(i{\hbox {th}}\) swing mode (rad/s)

\(x_{\delta _i}, x_{\omega _i}\) :

components of right eigenvector corresponding to rotor angle and speed, respectively

\(v_{\delta _i}, v_{\omega _i}\) :

components of left eigenvector corresponding to rotor angle and speed, respectively

Re(.):

real part of a complex number

\((.)^*\) :

element-wise conjugate of a complex vector

\((.)^T\) :

transpose

\((.)^H\) :

conjugate-transpose of a vector

G(s):

SISO transfer function

SSPF:

sum of slip participation factors

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Pradhan, V., Kulkarni, A.M. & Khaparde, S.A. Computation of the controllable swing mode spectrum of FACTS and HVDC in large power systems. Sādhanā 43, 176 (2018). https://doi.org/10.1007/s12046-018-0960-5

Download citation

  • Received:

  • Revised:

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s12046-018-0960-5

Keyword

Navigation