HVDC Power Injection

Embed Size (px)

Citation preview

  • 8/20/2019 HVDC Power Injection

    1/10

    1160 IEEE TRANSACTIONS ON POWER SYSTEMS, VOL. 28, NO. 2, MAY 2013

    Damping of Inter-Area Oscillations in MixedAC/DC Networks Using WAMS Based

    Supplementary Controller Robin Preece , Student Member, IEEE , Jovica V. Milanović , Fellow, IEEE , Abddulaziz M. Almutairi , Member, IEEE ,

    and Ognjen Marjanovic , Member, IEEE 

     Abstract— The paper presents a supplementary VSC-HVDC

    Power Oscillation Damping (POD) controller based on widearea measurement signals (WAMS). The controller is designedas Multi Input Single Output (MISO) using a Modal LinearQuadratic Gaussian (MLQG) methodology in order to targetcritical inter-area electromechanical modes. The approach hasbeen tested on a large (16 machine, 68 bus) test network incorpo-

    rating parallel HVDC/AC transmission and has shown improveddamping compared to a traditional Power System Stabilizer

    (PSS) based controller structure utilizing local signals. The designprocess has incorporated the effects of wide area signal transmis-

    sion delays. Variation in these signal delays and the complete lossof signals has been also investigated to establish the robustness of 

    the WAMS based controller and its sensitivity to loss of signals.Extension of the controller to incorporate reactive power modu-lation has been investigated, as has variation in available activepower modulation capacity. The proposed controller performance

    has been assessed through small and large disturbance analysis.

     Index Terms— Electromechanical modes, Linear Quadratic

    Gaussian (LQG) control, power oscillation damping, VSC-HVDC.

    I. I NTRODUCTION

    E LECTRICITY transmission networks of the future areexpected to incorporate large numbers of HVDC lines,leading to many instances of HVDC operation in parallel with

    AC lines. Worldwide interest in large renewable generation

    sources, often either offshore or long distances from traditional

    load centers is increasing. Moreover, in many cases planning

    consent for new overhead lines is becoming increasingly

    dif ficult to obtain, discouraging AC network reinforcements

    as too lengthy and costly. HVDC systems with higher power 

    transfer capacities per line, feasible lengthy subsea installations

    and no technical line length limitations provide an attractive

    alternative [1]. In the U.K., this is especially true with two

    Manuscript received January 09, 2012; revised February 28, 2012 and May15, 2012; accepted June 30, 2012. Date of publication August 15, 2012; date of current version April 18, 2013. This work was supported in part by the Engi-neering and Physical Sciences Research Council (EPSRC) and in part by Na-tional Grid Plc. Paper no. TPWRS-00030-2012.

    R. Preece, J. V. Milanović, and O. Marjoanovic are with the School of Elec-trical and Electronic Engineering, The University of Manchester, Manchester M60 1QD, U.K. (e-mail: [email protected]; [email protected]).

    A. M. Almutairi is with the College of Technological Studies, PAAET,Kuwait.

    Color versions of one or more of the  figures in this paper are available onlineat http://ieeexplore.ieee.org.

    Digital Object Identifier 10.1109/TPWRS.2012.2207745

     planned HVDC links to operate in parallel with the existing

    grid. These links will help to facilitate the increased power 

    transfer from the north to the south of the country when large

    renewable generation capacity is connected [2].

    A further benefit of HVDC that is largely unused in practical

    installations is that of power oscillation damping (POD). With

    fast acting power electronics within the converter stations it is

     possible to rapidly vary power   flow through the HVDC line.

    The potential for power injection control at non-generator buses

    (where HVDC systems are typically installed) has attracted re-

    cent interest and it has been shown that HVDC systems can be

    used with POD controllers to damp inter-area electromechan-

    ical oscillations [3]–[7].

    This paper   first presents a traditional Power System Stabi-

    lizer (PSS) based design for the supplementary HVDC POD

    controller with local signal input. Following this a more robust

    Wide Area Measurement Signals (WAMS) based controller 

    is designed using a targeted novel modal Linear Quadratic

    Gaussian (MLQG) methodology.

    LQG POD controllers have been applied to both centralized

    generator control [3], [4] and fast acting FACTS devices suchas Thyristor Controlled Series Capacitors (TCSCs) [5]–[7]. The

     paper shows that the novel MLQG controller design is effec-

    tive with an HVDC application. This design approach utilizes

    multiple global network signals subjected to transmission de-

    lays but controls only the HVDC line active power injection.

    The MLQG approach enables targeted damping only of the crit-

    ical inter-area system modes, leaving local modes unaffected.

    Through this design, vastly improved post-disturbance system

    stability is observed.

    The simplicity and transparency of this design approach

    distinguishes this controller from other existing multivariable

    synthesis techniques. Controller tuning is completed simply

    and effectively by applying nonzero weights only to thosetargeted modes requiring supplementary damping. This is far 

    easier than the case of standard LQG, requiring participation

    analysis and complex weightings, and approaches, where

    weighting functions must accurately model the uncertainties to

    ensure the practical robustness of the controller.

    II. MODAL LINEAR  QUADRATIC GAUSSIAN CONTROL

    The LQG control design is a cornerstone of modern optimal

    control theory and its advantages led to widespread research

    into it use in power system damping [3]–[7]. However, the de-

    sign approach is rarely straightforward, especially within large

     power systems where many generators participate in the critical

    0885-8950/$31.00 © 2012 IEEE

  • 8/20/2019 HVDC Power Injection

    2/10

    PREECE et al.: DAMPING OF INTER-AREA OSCILLATIONS IN MIXED AC/DC NETWORKS 1161

    modes which require additional damping. In these situations,

    the controller tuning process can become prohibitively complex.

    Specifically, the issues arise surrounding the correct selection of 

    system states upon which to target controller action.

    Participation factor analysis is required to identify the electro-

    mechanical states involved in targeted system modes. Weight-

    ings can then be assigned to these states. However, if these states

    are involved in other targeted modes (or modes that do not re-quire altering), the damping of these modes will also be affected,

    sometimes adversely. This results in a complex and time con-

    suming tuning process in which it is often not possible to obtain

    exact target damping factors. These complexities and problems

    can be overcome through the novel use of a modal representa-

    tion of the control design problem discussed below.

    Consider the following linearized state-space plant (power 

    system) model [8] described by (1) and (2):

    (1)

    (2)

    where w is process noise and v is measurement (sensor)

    noise. They are usually assumed to be uncorrelated zero-mean

    Gaussian stochastic processes with constant power spectral

    density matrices and respectively. The LQR control

     problem, in the modal formulation [9], is to devise a feedback 

    control law, which minimizes the cost function (3):

    (3)

    where and are appropriately chosen weighting matricessuch that and , and is

    a real matrix which provides mapping between system modal

    variables and state variables as in (4):

    (4)

    where modal variables are directly associated with system

    modes ( where ). The real transformation ma-

    trix is obtained using Real Schur Decomposition [10] and

    relates to the matrix of right eigenvectors as .

    The weighting matrices and are commonlyconstructed

    as diagonal. Values of the diagonal elements of are set in order 

    to penalize the corresponding controller’s outputs from high ac-tions. Values of the diagonal elements of are set in order 

    to penalize the corresponding modal variables when deviating

    from their steady-state values. Each diagonal element in

    is directly associated with a modal variable and hence with

    the corresponding mode . A higher value of modal weight

    corresponds to a higher effort by the controller to stabilize the

    corresponding mode. In order to focus on adding damping to the

    modes of interest only, these modes will be given some weights

    in while the other modes’ weights are set to zero. In this

    way, control effort of the designed LQR is directed towards the

    modes of interest only, by shifting them to the left in the com-

     plex plane, while keeping locations of other modes unaltered.

    The LQR controller gain is computed by solving the asso-

    ciated algebraic Riccati equation (ARE), based on cost function

    (3), and the  final LQG feedback control law can be written as

    (5):

    (5)

    where is an estimate of the states obtained using Kalman

    filter, described by (6):

    (6)

    where is a constant estimation error feedback matrix. The

    optimal choice of is that which minimizes

    . It is calculated by solving the ARE associated with the cost

    function (7):

    (7)

    In this paper, the weighting matrices and are tuned

    using the loop transfer recovery (LTR) procedure at plant input

    [8]. The Kalman  filter is synthesized such that the loop transfer 

    function , where is the plant transfer func-

    tion, approaches the LQR loop transfer function

    . The tuning parameters of Kalman   filter are

    calculated as in (8) and (9):

    (8)

    (9)

    where and refer to the nominal model, is any positive

    definite matrix and is a constant.Full recovery of robustness is achieved as . Care is re-

    quired as full recovery would lead to excessively high gains and

    therefore deteriorates the nominal performance of the true noise

     problem. For non-minimum phase systems, which is a common

    case in power systems, only partial recovery can be achieved

    [8].

    The closed-loop dynamics of the LQG controller can be de-

    scribed by (10):

    (10)

    and the transfer function of the LQG controller, from to , isthen given by (11):

    (11)

    The novel modal LQG (MLQG) controller provides many

     benefits of the traditional LQG approach:

    • Participation factor analysis is no longer required as LQR 

    weightings directly correspond to system modes.

    • It is simple to address the damping of several modes by

     providing non-zero weightings to all modes requiring ad-

    ditional damping.

    • Exact damping factors can be achieved if desired through

    fine tuning of the weightings.

  • 8/20/2019 HVDC Power Injection

    3/10

    1162 IEEE TRANSACTIONS ON POWER SYSTEMS, VOL. 28, NO. 2, MAY 2013

    Fig. 1. Injection model for an HVDC line in parallel with an AC line.

    • All untargeted modes are left completely unaffected by

    the controller, allowing local controllers to maintain good

    damping.

    A full comparison between the traditional and modal LQG

    controller designs is presented in [9].

    III. HVDC MODELING

    HVDC system modeling can range in complexity from full

    detail models which include valve switching to simple models

    which are only acceptable when the HVDC system is remote

    and has no significant impact on the stability analysis [11]. Thisresearch is concerned with electromechanical oscillations and

    so the operation of the power electronics within the HVDC con-

    verter stations is neglected. As such, the converter stations are

    represented simply as sinks and sources of active and reactive

     power [12].

     A. Injection Modeling 

    The method employed to model the HVDC system is injec-

    tion modeling [13]. The model complexity is determined by the

    level of detail required for the studies being performed. Concep-

    tually, the HVDC converter stations are replaced by equivalent

    generator buses connected to the AC system across a reactance

    , representing the converter transformer. Hence, the equiva-

    lent injection model for an HVDC line in parallel with an AC

    transmission line is given by Fig. 1.

    The voltage and angle at the equivalent generator buses are

    varied to produce the desired power   flow into and out of the

    HVDC system as dictated by (12) and (13). This method can be

    used to produce both voltage source conversion based HVDC

    (VSC-HVDC) and line commutated conversion based HVDC

    (LCC-HVDC) dynamic system models [14]:

    (12)

    (13)

     B. VSC-HVDC Injection Model Development 

    The more recently operationally viable VSC-HVDC pro-

    vides some benefits over its more established counterpart

    LCC-HVDC, and is seeing an increase in popularity. Its

    smaller footprint, decreased harmonic injection and reactive

     power injection capabilities were previously compromised by

    higher losses and lower operating capacities [15]. However 

    modular multi-level converters (MMC) have reduced these

    losses to levels comparable with LCC systems and many

    VSC-HVDC projects are currently under construction.

    The internal current control loop of the VSC is modeled as

    ideal for the system stability studies, meaning the active and

    TABLE IACTIVE POWER  IMPORTED I NTO  NETS  FROM  SURROUNDING  AREAS

    WITH  NO HVDC LINK  I NSTALLED

    reactive power control reference setpoints are met instantly.

    Converter station controllers are or integral regulators

    with clamped anti-windup. These controllers and the DC line

    dynamics (modeled as a -section) are utilized to determine

    the expected  flow into and out of the VSC-HVDC system with

    full description given in [14] and [16]. A common control

    scheme is used with one converter maintaining the DC voltage

    and the other regulating active power   flow. Reactive power control is independent at each converter station. Additional

    signals, and , are available to vary the active and

    reactive power reference setpoints and modulate power   flow

    for stabilizing purposes.

    IV. TEST  SYSTEM

    A 16-machine, 68-bus network is chosen as the test system

    for the studies, shown in Fig. 2. This represents a reduced order 

    equivalent model of the New England Test System (NETS)

    and the New York Power System (NYPS). Five separate areas

    are present: NETS consisting of G1–G9, NYPS consisting of G10–G13, and three further infeeds from neighboring areas

    are represented separately by G14, G15 and G16. All gener-

    ators are represented by full sixth order models. Generators

    G1–G8 are under slow DC excitation (IEEE-DC1A) while G9

    is equipped with a fast acting static exciter (IEEE-ST1A) and

    PSS. The remaining generators (G10–G16) are under constant

    manual excitation. Power system loads are modeled as constant

    impedance. Full system details, generator and exciter parame-

    ters are given in [17] with PSS settings for G9 taken from [18].

    All simulations are performed within the MATLAB/Simulink 

    environment making use of modified MATPOWER [19] to

     perform initial load  flows.

    With loading as given in [17], the NYPS area is heavilyimporting power from the surrounding areas due to an active

     power demand of 8.57 GW but generation of just 5.86 GW.

    Details of the active power import across inter-area ties are

    given in Table I. The infeed from the G16 area (along two

    lines) is largest, accounting for over half the total active power 

    import into NYPS.

    A VSC-HVDC link is introduced between the NYPS and

    G16 areas to support this power infeed. The link is installed

    in parallel with the most heavily loaded tie line, from bus 18

    to bus 50. Normal operating capacity for the HVDC link is se-

    lected as 400 MW. At this capacity, active power infeed from

    the G16 area to NYPS through AC tielines is reduced com-

     pared with those given in Table I: MW,

    and MW. Import from area G14 is also

  • 8/20/2019 HVDC Power Injection

    4/10

    PREECE et al.: DAMPING OF INTER-AREA OSCILLATIONS IN MIXED AC/DC NETWORKS 1163

    Fig. 2. 16-machine, 68-bus test system. Separate areas (NETS, NYPS, G14, G15, G16) shown with inter-area ties highlighted. VSC-HVDC link shown.

    TABLE III NTER -AREA  ELECTROMECHANICAL  MODE  DETAILS FOR  TEST  SYSTEM WITH  STANDARD LOADING

    slightly reduced with import from NETS largely unaffected. De-

    tails about the VSC-HVDC line parameters and controller gains

    are included in the Appendix.

    Small signal analysis of the linearized system including the

    VSC-HVDC reveals that with standard loading, as given in

    [17], four poorly damped oscillatory inter-area electro-mechan-

    ical modes are present which would benefit from improved

    damping. These modes are detailed in Table II. All these

    modes have damping factors, , below 5% which is considered

    unsatisfactory in terms of control design objectives [18]. The

    VSC-HVDC link is operating at 400 MW capacity with zero

    reactive power injection . All local electro-mechan-

    ical modes are adequately damped with .

    V. POD CONTROLLER  DESIGN

    The general control overview for the 16-machine, 68-bus

    network is shown in Fig. 3. For all designed supplementary

    HVDC POD controllers, output was limited to just one signal,

    . This signal is sent to the converter station regulating

    active power injection. In this study this is the inverter con-

    nected at bus 50. POD controller input signal transmission

    delays are experienced only for wide area, or global, signals as

    is discussed below.

     A. SISO PSS Structure

    1) Design Methodology:  The initial design process for the

    VSC-HVDC POD controller followed that of a conventional

    Fig. 3. Test network control overview.

    Fig. 4. PSS based SISO POD controller for HVDC.

    PSS. This has been used previously with HVDC power mod-

    ulation [20]–[23] and has been shown to be effective at modal

    damping. The design consists of a washout filter, lead-lag blocks

    and gain, as shown in Fig. 4. Controller parameters are depen-

    dent upon the modes of interest to be damped.

  • 8/20/2019 HVDC Power Injection

    5/10

    1164 IEEE TRANSACTIONS ON POWER SYSTEMS, VOL. 28, NO. 2, MAY 2013

    TABLE IIIR ESIDUE  A NGLES FOR  MODES OF I NTEREST FOR  HVDC PSS POD DESIGN

    The residue based tuning method is used [24]. First, the

    angles of the residue values, (of the open loop transfer 

    function between controller output and controller input), cor-

    responding to the critical eigenvalues are assessed. In this case

    the controller output will be the HVDC modulation reference,

    , and the controller input is selected as a local signal,

    the active power transfer into bus 50 from bus 18 through

    the parallel AC transmission line, . The phase com-

     pensation required by the PSS is then calculated simply as

    (to shift the eigenvalue further into the

    left half of the complex plane with no change in frequency). Pa-

    rameters for the lead-lag blocks are then determined according

    to the required phase compensation desired.

    Shown in Table III are the residue angles and required PSS

    compensation angles for the four poorly damped inter-area

    modes. As the lead-lag block is tuned for a specific compensa-

    tion angle at a specific frequency [24] it is clearly not possible

    to optimally tune the PSS for multiple modes.

    Tuning is carried out for the inter-area mode for which the

    magnitude of the residue value suggests greatest modal control-

    lability will occur given the controller input signal [11]. Residue

    magnitude values are also shown in Table III, normalized for the

    largest value. As Mode 1 is most controllable, tuning is carriedout for these modal characteristics at a standard VSC-HVDC

    operating capacity of 400 MW. Final controller parameters are

    given in the Appendix.

    2) Small Signal Analysis:   The improvement in modal

    damping with the PSS based SISO POD controller installed

    is shown in Fig. 6. As a local input signal is selected, it is as-

    sumed that there is no transport delay associated with receiving

    the signal . Similarly, it is assumed that the POD

    controller is located at the VSC-HVDC converter station con-

    trolling active power regulation, and no output signal transport

    delays are modeled.

    Very slight improvement is seen in all inter-area modes,

    though only Modes 1 and 2 cross the 5% damping thresholdwith damping coef ficient increasing to 7.08% and 5.44%,

    respectively. All local modes remain damped with .

    The small signal analysis performed suggests that the PSS-

     based POD has limited influence upon the critical inter-area

    electromechanical oscillations present within the test network.

    This may be in part due to limited modal observability [11]

    within the local input signal . This is a problem that

    can be overcome through use of global signals within a multiple

    input control structure.

     B. MISO Modal LQG Controller Design

    To allow targeted damping of oscillatory electromechanicalmodes, MLQG was selected as the controller design approach.

    Selection of inputs to the controller is determined through a

    full modal observability assessment for the network [11]. The

    number of required signals depends upon the number of crit-

    ical modes requiring additional damping and the observability

    of these modes within the available system signals. Networks

    containing a large number of PMUs will be able to exploit these

    available signals to gain a highly accurate representation of the

    network’s oscillatory nature following disturbances. Within thisstudy, active power   flow through lines from bus 45 to bus 51

    , bus 68 to bus 18 , bus 65

    to bus 17 , and from bus 67 to bus 42

    were determined to display greatest observability of 

    the critical Modes 1 to 4, respectively.

    A MISO MLQG power oscillation damping controller was

    designed, as described in Section II, for the standard operating

    conditions with a VSC-HVDC operational capacity of 400 MW.

    Controller inputs are as described above and controller output

    was limited to .

    Often when performing LQG control design it is necessary to

     perform initial model order reduction to avoid ill conditioningwhen solving high order matrix Riccati equations [6]. This

     problem was not experienced and so controller design was

    carried out using the full linearized system model.1) Incorporating Signal Delays:   As remote signals are

    assumed to be sent through pre-existing communications links

    (and not dedicated signal transmission hardware), delays be-

    tween the instant of measurement and the signal reaching the

    controller are experienced. Dependent upon various factors,

    including distance and communication protocols, these delays

    can be in the range of a few hundreds of milliseconds [25]. In

    the simulations performed, global network signals are assumed

    to have associated transport delays of 500 ms. These signal de-

    lays are modeled as 2nd order Padé approximations [26] duringmodel linearization and controller design. As with the PSS

     based POD controller, it is assumed that the POD controller 

    is located at the VSC-HVDC converter station and no output

    signal transport delays are modeled.

    The introduction of delays to the controller input signals af-

    fects the LTR process during the design of the LQG controller.

    The degree of recovery achieved can be assessed through

    comparison of the singular value plots for both and

    . With parameters , , , and set to

    fixed values as given in the Appendix, the LTR procedure was

    completed for , shown in Fig. 5. Increasing

     beyond 100 provides no further discernable improvement in

    recovery in the frequency range of interest. Fig. 5 demonstrates

    that with the signal transport delays included, full recovery of 

    the robustness properties of is not possible, even

    for large . For the rest of the design process, was   fixed

    at 1, providing a compromise between optimal recovery of 

    robustness properties and unacceptably high  filter gains.

    2) Small Signal Analysis and Controller Reduction:   Small

    signal analysis on the closed loop system (of Fig. 3) demon-

    strates the improvements seen in damping of all critical

    inter-area modes. This is shown in Fig. 6 where it can be

    seen that all critical modes now achieve damping greater than

    5%. Modes 1 to 4 are now damped with values of 19.39%,

    15.93%, 14.09%, and 11.63%, respectively. This is not onlygreatly improved over the “no POD” case but also over the

  • 8/20/2019 HVDC Power Injection

    6/10

    PREECE et al.: DAMPING OF INTER-AREA OSCILLATIONS IN MIXED AC/DC NETWORKS 1165

    Fig. 5. LTR process at plant input with varying valueswith delayed controller input signals.

    Fig. 6. Modal placement with no POD, PSS-based POD, full order MLQGPOD, and reduced order MLQG POD (dashed line signifies 5% damping).

    modal placements for the PSS-based POD controller. With the

    targeted modal damping from the MLQG controller affecting

    solely the inter-area Modes 1 to 4, all local modes remain

    unaffected and adequately damped with .

    The designed MISO MLQG controller as defined by (11) is

    of 182nd order, equal to the full linearized system model order.

    It is desirable to reduce this in order to decrease the online com-

     putational burden of the controller whilst still maintaining the

    improved critical mode damping.The balanced Schur reduction method [27] was applied to the

    controller, implemented within MATLAB. The limiting factor 

    upon the level of reduction permitted was chosen as the degra-

    dation seen in the damping of critical modes through the use of 

    the reduced order controller. Allowable degradation was set at

    5% of the full order controller values for each critical mode.

    With this methodology, the MLQG POD controller was reduced

    to 28th order. Small signal analysis of the closed loop system

    with the reduced order controller provides values of 18.90%,

    15.43%, 14.34%, and 11.70% for Modes 1 to 4, respectively,

    also shown in Fig. 6.

    The small signal analysis performed is dependent upon the

    linearization of the nonlinear power network and elements such

    as the signal transmission delays. Furthermore, the improve-

    ment in inter-area mode damping is “ideal” and based on the

    LQR state feedback control eigenvalue placement (which as-

    sumes full state knowledge). The true performance and robust-

    ness of the designed controller is dependent upon the Kalman

    filter state estimator and is, therefore, most readily assessed

    through transient simulations.

    VI. CONTROLLER  TRANSIENT PERFORMANCE

     A. Variation in System Operating Conditions

    Large disturbance transient studies have been performed for 

    two cases. These are:

    Fig. 7. For  base case, settling times for NYPS inter-area AC infeeds.

    Fig. 8. For  outage case, settling times for NYPS inter-area AC infeeds.

    1) The base case, theoperating point for which the controllers

    have been designed. System loading and generation is

    as previously described. The VSC-HVDC is operating at

    400-MW capacity with zero reactive power output. Alllines are in service. The system is subjected to a 100-ms

    self-clearing fault at bus 38 at a time of 0.5 s.

    2) The  outage case. Still with standard loading, the line be-

    tween bus 18 and bus 49 is removed from service. This line

     provided a path for some of the power  flow from the G16

    area to NYPS. As a consequence, power  flow through the

    line from bus 18 to bus 50 is increased. VSC-HVDC oper-

    ational capacity is increased slightly to 450 MW to aid this

    transmission, still with zero reactive power injection. The

    system is then subjected to a 100-ms fault near to bus 1 on

    the line from bus 1 to bus 30, at a time of 0.5 s, cleared by

    disconnecting the line.These transient studies were performed for the case with no

    POD controller installed, with PSS-based POD, and with the re-

    duced order MLQG POD including signal delays. VSC-HVDC

    modulation was limited to 100 MW.

    1) Transient Results and Discussion:   Settling times were

    recorded from the point of fault clearance to the time at which

    the power deviation was within 1% of the steady state value

    (or 1 MW, whichever was greatest). These times are shown

    for the NYPS inter-area AC infeeds (as detailed in Table I) in

    Figs. 7 and 8.

    Looking  first at the  base case  (Fig. 7), the inter-area infeed

     power   flows demonstrate the considerably improved modal

    damping of the MLQG controller. All infeeds settle in less than12 s, compared with 23 s with the PSS-based POD controller 

    installed, and 31 s with no damping controller.

    With the  outage case   (Fig. 8), the settling times of the AC

     NYPS infeeds show the robustness of the MISO MLQG con-

    troller to varying operating point. The improved performance

    over the PSS-based controller is still pronounced with all ties

    settling within 19 s (compared to 33 s for the PSS-based POD

    and 40 s when no POD controller is used).

    Plots of the active power injected at bus 9 from bus 8 through

    the AC infeed and the active power injected at bus 50 by the

    VSC-HVDC are shown in Figs. 9(a) and 10(a). These show that

    despite the separation between the point of control (the VSC-

    HVDC link between buses 18 and 50) and a relatively distant

    inter-area tie (bus 8 to bus 9), damping is still vastly improved.

  • 8/20/2019 HVDC Power Injection

    7/10

    1166 IEEE TRANSACTIONS ON POWER SYSTEMS, VOL. 28, NO. 2, MAY 2013

    Fig. 9. For   base case: (a) Active power injected at bus 9 (NYPS) from bus8 (NETS), and (b) Active power injected at bus 50 by VSC-HVDC inverter station. : fault occurs, : fault cleared, PSS POD power  modulation evident, : MLQG POD power modulation evident.

    Fig. 10. For  outage case: (a) Active power injected at bus 9 (NYPS) from bus8 (NETS), and (b) Active power injected at bus 50 by VSC-HVDC inverter station. : fault occurs, : fault cleared, PSS POD power  modulation evident, : MLQG POD power modulation evident.

    This is still evident forthe outage case, demonstrating the ability

    of the MLQG controller to perform well as operating conditions

    change.

    The plots of the active power injected by the inverter,Figs. 9(b) and 10(b), present the control action of the various

    POD controllers. The forced modulation of active power  flow

    through the VSC-HVDC link is used to stabilize the network.

    With no POD it can be seen that the VSC-HVDC returns to

    its steady state power injection setpoint rapidly. This can also

     be seen to occur with the MLQG POD controller following

    the disturbance at 0.5 s (labeled t1). It is not until 500 ms after 

    the fault instant (at ) that the controller input signals

    display a disturbance and the HVDC active power modulation

     begins. The active power modulation of the PSS POD using

    signals with no delay is evident as soon as the fault is cleared

    (at ).

    Further investigation has been made into the effects of 

    varying operating conditions including generator and key tie

    Fig. 11. Deterioration in damping of critical modes by MLQG controller withincreasing signal delay.

    line outages with both the PSS-based POD and MLQG POD

    controller, available in [28]. It should be noted that the MLQG

    controller significantly outperformed the PSS-based controller 

    across wide ranging operating scenarios.

     B. Variation in Wide Area Signal Delay and Loss of Signals

    The use of wide area measurements has been demonstrated as

    enabling a better, more robust controller design. These signals

    are becoming more prevalent in modern power systems, pro-viding reliable real-time data which can improve many aspects

    of system performance. However, these signals will often be

    sent through pre-existing satellite communication links (as ded-

    icated hard-wired links may prove prohibitively expensive) and

    as such are potentially subject to increased delay or even com-

     plete loss. If faster communication channels (e.g.,   fi ber optic

    links) are available then the shorter associated delays should be

    included in the controller design stage. However, controller per-

    formance will only improve as signal delays shorten [9].

    The effects of signal latency and mitigation techniques for 

    use with WAMS based controllers has been a topic of much in-

    terest. Readers are directed towards [29]–[34] for comprehen-sive analyses of these effects and novel mitigation techniques

    to improve controller performance in the presence of signal la-

    tency. For the novel MLQG controller, the effects of increased

    signal delays and the complete loss of signals have been investi-

    gated in order to demonstrate the controller’s robustness to these

     problems.

    The controller’s local signal to the VSC-HVDC converter sta-

    tion is assumed to be hard-wired and not subject to these issues.

    In addition to the results investigating the robustness of the con-

    troller to varying system operating conditions, the MLQG con-

    troller has also been tested to determine the robustness to vari-

    ation in signal delay and complete loss of input signals. Fig. 11

    shows the deterioration seen in the damping of all critical modesas the delays are increased to 1100 ms for the  base case  oper-

    ating point. There is a gradual degradation in controller perfor-

    mance as the signals are delayed for longer than the 500 ms

    assumed during the design process.

    Table IV details the maximum permissible delay tolerances

     beyond which the MLQG controller will exhibit worse perfor-

    mance than with the PSS-based POD controller, or no POD con-

    troller, installed. These delays can increase to 766 ms (53.2%

    higher than the already pessimistic 500 ms assumed) before the

    controller displays worse performance than the PSS-based POD

    controller taking local signals. With respect to the complete loss

    of signals, this has been tested for the loss of up to two of the

    four input signals. As the signals selected all contain some in-

    formation about all the critical modes, there is some redundancy

  • 8/20/2019 HVDC Power Injection

    8/10

    PREECE et al.: DAMPING OF INTER-AREA OSCILLATIONS IN MIXED AC/DC NETWORKS 1167

    TABLE IVMLQG CONTROLLER  SIGNAL  DELAY TOLERANCES

    TABLE VDAMPING FACTOR OF CRITICAL  MODES WITH LOSS OF I NPUT SIGNALS

    Fig. 12. Active power injected at bus 9 (NYPS) from bus 8 (NETS) for  basecase  operating point with and signal failures.

    within them. The damping of the four critical modes for the var-

    ious cases is shown in Table V.

    The shaded cells represent the cases when the modal damping

    factor drops below that seen with the local PSS-based POD con-

    troller installed. It is evident, for example, that the loss of input

    signal always results in the damping factor of Mode 3 being

    heavily reduced to less than 5%. This signal was initially se-

    lected for the high observability of Mode 3, so this result is

    not unexpected. Damping factors can be seen to still be higher than with the PSS-based POD installed for the majority of cases

    when signals fail.

    Even with the (unlikely) loss of two input signals, the MLQG

    controller can often maintain relatively high damping factors

    on some modes. Looking at case 5 (highlighted in Table V),

    even though damping of Mode 2 has dropped to “lower than

    PSS POD” levels, the damping of the remaining low frequency

    modes are still in the range of 8.87%–9.60%. Due to this fact,

    the transient performance of the controller is still highly com-

     petitive, with all infeeds settling within 16 s for the  base case

    scenario when signals and are lost. The oscillationspresent

    on the tie line between buses 8 and 9 are shown in Fig. 12, the

    robustness of the MLQG controller to the failure of wide area

    signals is clearly visible with the oscillations quickly damped.

    Fig. 13. Active power injected at bus 9 (NYPS) from bus 8 (NETS) for  basecase  with and without reactive power modulation.

    C. Incorporating Reactive Power Modulation

    The studies presented have been concerned with the modu-

    lation of active power   flow through the parallel VSC-HVDC

    link in order to stabilize post-disturbance system oscillations.

    One of the stated advantages of using VSC-HVDC over classic

    LCC-HVDC is the availability of “four quadrant” operation of 

    the converters, allowing the generation or consumption of reac-

    tive power at each converter station.The MLQG design approach can be readily extended to in-

    clude multiple controller outputs into a MIMO structure. In ad-

    dition to the single signal, signals at each con-

    verter station were incorporated and the MLQG controller de-

    sign was completed once more.

    Fig. 13 shows the oscillations present on the tie line between

     buses 8 and 9 for the base case operating scenario both with, and

    without, reactive power modulation included. It can be seen that

    the additional reactive power modulation provides very limited

    improvement in the system response (just 0.4 s improvement in

    settling time).

    With little benefit achieved, it is unlikely that VSC-HVDC

    reactive power output would be modulated for power oscillationdamping purposes. Perhaps more probable would be the use of 

    fast reactive power modulation to ensure quickly stabilized bus

    voltages at the points of interconnection with the VSC-HVDC

    link during the post-disturbance oscillations.

     D. Variation in Modulation Capacity

    The simulations presented within this paper have assumed

    a generous allowance of 100 MW for power oscillation

    damping (equating to 25% of the VSC-HVDC link operating

    capacity for the base case   scenario). This modulation capacity

    is the same for both designed POD controllers.

    In a practical installation, the limit of available modulationcapacity will be determined by the system operator. The bene-

    fits of reserving this capacity for modulating purposes following

    system disturbances must be compared with the costs of re-

    ducing the power transfer capability through the HVDC link.

    Converter ratings will set the upper bound on possible power 

    transfer. As the required modulations in active power are rela-

    tively slow (typically about 1 Hz), DC voltage limit violations

    should not be an issue provided the converter regulating DC

    voltage has been designed to be suitably fast.

    The effect of limiting the modulation capacity (from the pre-

    viously considered 25%) to 10% and 5% has been investigated.

    Fig. 14 shows the oscillations between buses 8 and 9 and the

    controlled variations in the power injected by the inverter station

    for the   base case  with varying limits on modulation capacity

  • 8/20/2019 HVDC Power Injection

    9/10

    1168 IEEE TRANSACTIONS ON POWER SYSTEMS, VOL. 28, NO. 2, MAY 2013

    Fig. 14. (a) Active power injected at bus 9 (NYPS) from bus 8 (NETS) and (b)Active power injected at bus 50 by VSC-HVDC inverter station for  base caseoperating point with differing modulation capacity limits.

    with the initially designed MLQG controller (varying only ac-

    tive power injection).

    It can be seen that the settling time for this line increases

    slightly as less capacity is reserved for POD action. The same

    result is true of the PSS-based controller. With the exception

    of the line 18–50 (which always sees improved settling times

    with the PSS POD due to the local signal selection), restricting

    the capacity reserved for POD to just 5% with the MLQG con-

    troller still results in improved settling times over the PSS-based

    POD controller operating with 25% modulation capacity. For 

    the MLQG controller 10% modulation capacity results in key

    tie line settling times increasing by 1.9–2.7 s (to a maximum of 

    14.1 s); and 5% modulation capacity results in key tie line set-tling times increasing by 2.7–4.7 s (to a maximum of 16.1 s);

    when compared with the initial 25% modulation capacity.

    An idea of the likely availability of this modulation capacity

    can be sourced from the publicly available data on the usage

    of the 1 GW Britned HVDC link between July and December 

    2011 [35]. This data is taken for an HVDC link which does not

    reserve capacity for modulation. During this six month period

    (ignoring periods with no power transmission): 73.8% of the

    time at least 5% link capacity was spare; 70.6% of the time at

    least 10% link capacity was spare; and 63.1% of the time at least

    25% link capacity was spare.

    It is clear that there will be periods when large amounts of 

    HVDC link capacity may be available for active power modu-lation for system stabilizing purposes. At these times it would

     be advisable to more fully exploit the damping capabilities of 

    the HVDC link as higher modulation capacities result in faster 

    system settling times. This may require   flexible modulation

    limits dependent upon system operating conditions.

    VII. CONCLUSIONS

    A MISO Modal LQG power oscillation damping controller 

    for VSC-HVDC lines has been shown to be effective at

    damping inter-area electromechanical oscillations within a

    large heavily meshed network. Furthermore, it has been shown

    that even when accounting for transmission delays on wide

    area controller input signals, the MISO controller is able to

    vastly outperform a standard PSS based SISO POD controller 

    dependent upon local signals. Furthermore, the effects of 

    variation in the transmission delays of the wide area signals

    the MLQG controller receives has been demonstrated. Not

    only can the controller tolerate delays over 50% longer than

    designed for, but it can even continue to outperform the PSS

     based controller with the loss of half of its wide area inputs. It

    has been shown that the addition of reactive power modulationat each VSC-HVDC converter station offers little benefit with

    respect to improved system settling times. The damping of 

     power oscillations is dominated by the active power modulation

    through the HVDC link.

    Furthermore, the effects of limiting the capacity available for 

    active power modulation have been demonstrated. The MLQG

    controller performance was reduced as the capacity reserved for 

    POD controller action is also reduced; however the controller 

    maintained superior performance compared with the PSS-based

    controller. Analysis of operational data for the Britned link 

    shows the availability of modulation capacity and suggests

    that use of   fl

    exible controller limits may be advantageous infully exploiting the VSC-HVDC link’s ability to stabilize the

    network. This paper has shown that active power modulation of 

    HVDC lines is effective at damping multiple inter-area modes

    within a large, heavily meshed network. Also, the MLQG

    control methodology has been shown to be implementable with

    a VSC-HVDC application. The targeted damping of critical

    electromechanical oscillatory modes afforded by the MLQG

    POD controller design may be of particular interest in large

     power systems where very selective additional damping is

    desired.

    Finally, It should be pointed out that although use of an LQG

    controller synthesis approach cannot intrinsically guarantee ro-

     bustness properties [36], robust controller performance can still be achieved. Conversely, the use of controller synthesis tech-

    niques that guarantee the robustness of the controller can only

    ensure this robustness within the bounds of the defined uncer-

    tainties. If these uncertainties are poorly formalized then the

    controller’s practical ef ficacy may not match its intended math-

    ematical performance. In either case, the performance and sta-

     bility of the final controller must be thoroughly assessed through

    multiple means (including nonlinear simulations) in order to es-

    tablish its true robustness.

    APPENDIX

    HVDC SYSTEM AND CONTROLLER  DATA

    VSC-HVDC Line Parameters (on 600-MVA HVDC base):

    VSC-HVDC Controller Parameters:

    SISO POD Control Parameters (on 100-MVA base):

    Fixed Parameters during LTR Tuning:

  • 8/20/2019 HVDC Power Injection

    10/10

    PREECE et al.: DAMPING OF INTER-AREA OSCILLATIONS IN MIXED AC/DC NETWORKS 1169

    R EFERENCES

    [1] J. Arrillaga, Y. H. Liu, and N. R. Watson , Flexible Power Transmis- sion: The HV DC Op tions. Chichester, U.K.: Wiley, 2007.

    [2] National Grid Electricity Transmission plc, Offshore Development In-formation Statement, Sep. 2010. [Online]. Available: http://www.na-tionalgrid.com.

    [3] D. Dotta, A. S. e Silva, and I. C. Decker, “Wide-area measurements- based two-level control design considering signal transmission delay,”

     IEEE Trans. Power Syst., vol. 24, pp. 208–216, 2009.[4] T. Michigami, M. Terasaki, N. Sasazima, K. Hayashi, and T. Okamoto,“Developmentof a newadaptiveLQG system generatorfor high-speeddamping control techniques of power system oscillation,”  Elect. Eng.

     Jpn., vol. 142, pp. 30–40, 2003.[5] A. C. Zolotas, B. Chaudhuri, I. M. Jaimoukha, and P. Korba, “A study

    on LQG/LTR control for damping inter-area oscillations in power sys-tems,” IEEE Trans. Control Syst. Technol., vol. 15, pp. 151–160, 2007.

    [6] K. M. Son and J. K. Park, “On the robust LQG control of TCSC for damping power system oscillations,” IEEE Trans. Power Syst. , vol. 15, pp. 1306–1312, 2000.

    [7] A. M. D. Ferreira, J. A. L. Barreiros, J. W. Barra, and J. R.Brito-de-Souza, “A robust adaptive LQG/LTR TCSC controller applied to damp power system oscillations,”  Elect. Power Syst. Res.,vol. 77, pp. 956–964, 2007.

    [8] S. Skogestad and I. Postlethwaite , Multivariable Feedback Control: Analysis and Design. Chichester, U.K.: Wiley, 1996.

    [9] A. Almutairi, “Enhancement of power system stability using wide areameasurement system based damping controller,” Ph.D. dissertation,Univ. Manchester, Manchester, U.K., 2010.

    [10] G. H. Golub and C. F. V. Loan , Matrix Computations. Baltimore,MD: The John Hopkins Univ. Press, 1989.

    [11] P. Kundur  , Power System Stability and Control . London, U.K.: Mc-Graw-Hill, 1994.

    [12] P. E. Bjorklund, K. Srivastava, and W. Quaintance, “Hvdc light mod-eling for dynamic performance analysis,” in  Proc. PSCE ’06 , 2006.

    [13] H. F. Latorre, M. Ghandhari, and L. Söder, “Active and reactive power control of a VSC-HVdc,” Elect. Power Syst. Res., vol. 78, pp.1756–1763, 2008.

    [14] R. Preece and J. V. Milanovic, “Comparison of dynamic performanceof meshed networks with different types of HVDC lines,” in  Proc. IET 

     ACDC , London, U.K., 2010.[15] N. Flourentzou, V. G. Agelidis, and G. D. Demetriades, “VSC-based

    HVDC power transmission systems: An overview,”   IEEE Trans. Power Electron., vol. 24, pp. 592–602, 2009.

    [16] R. Preece, A. M. Almutairi, O. Marjanovic, and J. V. Milanovic,“Damping of electromechnical oscillations by VSC-HVDC active power modulation with supplementary WAMS based modal LQGcontroller,” in Proc. IEEE PES General Meeting , Detroit, MI, 2011.

    [17] B. Pal and B. Chaudhuri , Robust Control in Power Systems. NewYork: Springer, 2005.

    [18] G. Rogers , Power System Oscillatio ns. Norwell, MA: Kluwer, 2000.[19] R. D. Zimmerman, C. E. Murillo-Sanchez, and R. J. Thomas, “MAT-

    POWER: Steady-state operations, planning, and analysis tools for  power systems research and education,” IEEE Trans. Power Syst., vol.26, pp. 12–19, 2011.

    [20] C. Zheng, X. Zhou, and L. Ruomei, “Dynamic modeling and transientsimulation for VSC based HVDC in multi-machine system,” in  Proc.

     PowerCon ’06 , 2006.[21] Y. Pipelzadeh, B. Chaudhuri, and T. C. Green, “Wide-area power os-

    cillation damping control through HVDC: A case study on Australianequivalent system,” in Proc. IEEE PES General Meeting , 2010.

    [22] H. Jingbo, L. Chao, W. Xiaochen, W. Jingtao, and T. S. Bi, “Designand experiment of heuristic adaptive HVDC supplementary dampingcontroller based on online Prony analysis,” in  Proc. IEEE Power En-

     gineering Society General Meeting , 2007.[23] D. V. Hertem,R. Eriksson,L. Soder, andM. Ghandhari, “Coordination

    of multiple power  flow controlling devices in transmission systems,”in Proc. IET ACDC , London, U.K., 2010.

    [24] J. V. Milanovicand S. K. Yee, “Roadmap fortuning power systemcon-trollers,” in  Proc. 3rd IASTED Int. Conf. Power and Energy Systems,Marbella, Spain, Sep. 2–4, 2003.

    [25] R. Majumder, B. Chaudhuri, and B. C. Pal, “Implementation and testresults of a wide-area measurement-based controller for damping in-terarea oscillations considering signal-transmission delay,”  IET Gen.,Transm., Distrib., vol. 1, pp. 1–7, 2007.

    [26] J. Lam, “Model reduction of delay systems using Pade approximants,” Int. J. Control , vol. 57, pp. 377–391, 1993.

    [27] M. G. Safonov and R. Y. Chiang, “A Schur method for balanced-trun-cation model reduction,”   IEEE Trans. Autom. Control , vol. 34, pp.729–733, 1989.

    [28] R. Preece, A. M. Almutairi, O. Marjanovic, and J. V. Milanovic, “Ef-fectiveness of a supplementary MLQG power oscillation damping con-troller installed at an HVDC line within a meshed network,” in  Proc.Cigre Int. Symp.: The Electric Power System of the Future , Bologna,Italy, 2011.

    [29] J. W. Stahlhut, T. J. Browne, G. T. Heydt, and V. Vittal, “Latencyviewed as a stochastic process and its impact on wide area power system control signals,”  IEEE Trans. Power Syst., vol. 23, pp. 84–91,2008.

    [30] W. Hongxia, K. S. Tsakalis, and G. T. Heydt, “Evaluation of timedelayeffects to wide-area powersystem stabilizer design,” IEEE Trans.

     Power Syst., vol. 19, pp. 1935–1941, 2004.[31] B. Mirkinand P. O. Gutman, “Robust output-feedback model reference

    adaptive control of SISO plants with multiple uncertain, time-varyingstate delays,”  IEEE Trans. Autom. Control , vol. 53, pp. 2414–2419,2008.

    [32] B. Chaudhuri, R. Majumder, and B. C. Pal, “Wide-area measurement- based stabilizing control of power system considering signal transmis-sion delay,” IEEE Trans. Power Syst., vol. 19, pp. 1971–1979, 2004.

    [33] R. Majumder, B. Chaudhuri, B. C. Pal, and Z. Qing-Chang, “A unified

    Smith predictor approach for power system damping control designusing remote signals,” IEEE Trans. Control Syst. Technol., vol. 13, pp.1063–1068, 2005.

    [34] N. R. Chaudhuri, S. Ray, R. Majumder, and B. Chaudhuri, “A new ap- proach to continuous latency compensation with adaptive phasor power oscillation damping controller (POD),”  IEEE Trans. Power Syst., vol.25, pp. 939–946, 2010.

    [35] National Grid Transmission, “Metered Half-Hourly ElectrictyDemands: July–December 2011,” 2011. [Online]. Available:http://www.nationalgrid.com.

    [36] J. Doyle, “Guaranteed margins for LQG regulators,”   IEEE Trans. Autom. Control , vol. 23, pp. 756–757, 1978.

    Robin Preece (S’10) received the B.Eng degree in electrical and electronic en-gineering in 2009 from the University of Manchester, Manchester, U.K., where

    he is currently pursuing the Ph.D. degree.

    Jovica V. Milanović   (M’95–SM’98–F’10) received the Dipl.Ing. and M.Sc.degrees from the University of Belgrade, Yugoslavia, the Ph.D. degree fromthe University of Newcastle, Australia, and his Higher Doctorate (D.Sc. de-gree) from The University of Manchester, Manchester, U.K., all in electricalengineering.

    Currently, he is a Professor of electrical power engineering and Director of External Affairs in the School of Electrical and Electronic Engineering at TheUniversity of Manchester (formerly UMIST), Visiting Professor at the Univer-sity of Novi Sad, Novi Sad, Serbia, and Conjoint Professor at University of  Newcastle, Newcastle, Australia.

    Abddulaziz M. Almutairi (S’06–M’11) received the B.Sc. degree from KuwaitUniversity, Kuwait, the M.Sc. degree from the University of North Carolina atCharlotte, and the Ph.D. degree from the University of Manchester, Manchester,U.K., all in electrical engineering.

    Currently he is an Assistant Professor of electrical power engineering in theCollege of Technological Studies, PAAET, Kuwait.

    Ognjen Marjanovic  (M’08) received the First Class honors degree from theDepartment of Electrical and Electronic Engineering, Victoria University of Manchester, U.K., and the Ph.D. degree from the School of Engineering, Vic-toria University of Manchester, U.K.

    Currently he is a Lecturer in the School of Electrical and Electronic Engi-neering at The University of Manchester, U.K.