You are on page 1of 13

This article has been accepted for publication in a future issue of this journal, but has not been

fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TPWRS.2019.2895081, IEEE
Transactions on Power Systems

Plug-and-Play Compliant Control for Inverter-Based


Microgrids
Po-Hsu Huang, Student Member, IEEE, Petr Vorobev, Member, IEEE, Mohamed Al Hosani, Member, IEEE,
James L. Kirtley, Life Fellow, IEEE, and Konstantin Turitsyn, Member, IEEE

Abstract—Control of inverter-based microgrids (IBMGs) with and-play functionality, and similar power control structure to
power sharing capability is a challenging task due to specific synchronous machine systems [1]–[4].
nature of microgrid instabilities. Specifically, droop-controlled The droop control is based only on local actions that allow
microgrids prone to become unstable when the X/R ratio
is near unity and the lines are short. Virtual impedance is desired power distribution among multiple inverters operating
thought to be one of the promising tools to prevent this type in grid-forming modes, which ensures proper power sharing
of instability allowing for a broad range of droop coefficients and system resiliency. Strong coupling of active power to
for efficient power sharing between inverters. However, proper frequency and reactive power to voltage due to high X/R
sizing of virtual impedance for particular microgrid configuration ratios allows such adoption for conventional power systems.
is still an open question. In this work, a systematic method
for determining the size of virtual impedance in a Plug-and- Similarly, such a strategy was proposed for inverters connected
Play (PnP) fashion is proposed for enhancing the stability of the to AC grids [5] or inverter-based microgrids (IBMGs) [1], [3],
droop modes. Based on the Lyapunov method, concise and simple [4]. However, due to much smaller X/R ratios, which are
stability certificates are derived to provide flexible choices of the around unity for IBMGs, the coupling between real power and
droop gains for different inverter ratings. The implementation of frequency becomes distorted, due to which studies have hy-
the virtual impedance with the proposed approach for stability
enhancement is verified by both numerical simulation through pothesized limited small-signal stability regions for allowable
MATLAB/Simulink and experiment via a small-scale prototype droop gains for the microgrid as compared to conventional
for different scenarios. grids [6]. To investigate this issue, most studies utilized a
Index Terms—Droop control, microgrids, virtual impedance, seem-to-be-reasonable assumption neglecting electromagnetic
Lyapunov function, small-signal stability. network dynamics - an approximation routinely used for con-
ventional power system - that was justified by a distinct time
scale separation between the network and power control modes
I. I NTRODUCTION [7]–[10]. However, both experimental and detailed numerical
Small-scale power systems, capable of autonomous oper- simulations for IBMGs have shown that the true stability
ation - microgrids (MGs) - have become feasible due to region is much smaller than the one obtained from such
growing production of cost-effective battery storage with quasi-stationary approximation, signifying the importance of
high-bandwidth inverter interface, making effective energy electromagnetic transients despite their small time-scale [1],
dispatch from uncontrollable renewable sources to maintain [11], [12]. In a recent work [12], the influence of these fast
system operation without grid connections. Therefore, control degrees of freedom on dynamics of slow power controller
of inverter-based or inverter-dominant systems is gaining a modes has been analytically demonstrated. In particular, for
lot of attention while posing different challenges as com- a typical IBMG, the small relative impedance of the network
pared to conventional power systems. Among different control (leading to strong coupling between different inverters) was
schemes, droop-based control is widely appreciated due to identified as one of the main reasons for instability, which is
many advantages such as power-sharing capability without fully consistent with the experimental results presented in [13].
communication, high expandability and reliability with plug- It is, therefore, well justified to adopt the virtual impedance
for stability enhancement of IBMGs.
This work was supported in part by the MUSES project and the Cooperative Before the recognition of the importance of electromagnetic
Agreement between the Khalifa University, Abu Dhabi, UAE and the Mas- transients, the idea of suppressing the natural strong coupling
sachusetts Institute of Technology (MIT), Cambridge, MA, USA - Reference
02/MI/MIT/CP/11/07633/GEN/G/00 and in part by the Skoltech-MIT Next in IBMGs had led to various studies aiming at the increase
Generation program. of effective line impedance by physical or virtual means [10],
P.-H.Huang was with Department of Electrical and Computer Engineering, [14]–[19]. Particularly, the emulation of inductive dynamics
Massachusetts Institute of Technology. Email: pohsu0113@gmail.com
P. Vorobev is with Skolkovo Institute of Science and Technology, Moscow, through a digital controller helps to save bulky and costly
Russia, was with Department of Mechanical Engineering, Massachusetts inductors with the enhancement of power sharing accuracy and
Institute of Technology. E-mail: p.vorobev@skoltech.ru stability. In general, the control is realized by manipulating
M. Al Hosani, currently with Abu Dhabi Distribution Company, Abu Dhabi,
UAE, was with Department of Electrical and Computer Engineering, Khalifa voltage set-points with measured output currents to mimic
University, Abu Dhabi, 54224, UAE. E-mail: malhosani85@hotmail.com the dynamic behavior of an impedance, so that an inverter
J. L.Kirtley is with Department of Electrical and Computer Engineering, system behaves like an internal voltage source connected to
Massachusetts Institute of Technology. Email: kirtley@mit.edu
K. Turitsyn was with Department of Mechanical Engineering, Mas- its terminal through a virtual impedance [14].
sachusetts Institute of Technology. E-mail: turitsyn@gmail.com However, the proper choice of the size of virtual impedance

0885-8950 (c) 2018 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TPWRS.2019.2895081, IEEE
Transactions on Power Systems

for droop instability is not well covered in literature. In


general, the conventional approaches rely on either detailed I f dq Vcedq I dq
e
P Q I dq Vs e j 
jT
models with eigenvalues extracted through linearization [13] Ve
or very simplified models, like Kuramoto’s oscillators [7]. 5I /I &I 5P /P 5 F5O / F/O
3&&
Working with detailed models of high orders is very non- ,QYHUWHUDQG/&)LOWHU (TXLYDOHQW&LUFXLW
intuitive and time-consuming while simplified models tend
D
to be erroneous in predicting the true stability boundaries.
Thus, there is a practical need to develop a guideline for Vd V Vdq 3RZHU P
9LUWXDO,PSHGDQFH
control engineers that maintains both simplicity and fidelity. 
9ROWDJH&RQWURO
I f  dq I dqe 0HDVXUHPHQW Q
Vq 
We note, that there could be different types of instability
Vcedq
Zset
/3)
present in microgrids, however, small-signal stability is the Feed-forward Terms

most fundamental requirement, which is mandatory for any P W s 
k pZ  Z
&XUUHQW DEF e /3)
microgrid type and configuration. I f  dq

&RQWURO GT
V INV 
Q kq  V
W s 
In this paper, we focus on deriving small-signal stability
I ef dq Feed-foward
T s  Z
rules and stability enhancement methods for microgrids. The Terms Vset
problem is first investigated analytically and reduced-order ,QWHUQDO&RQWURO 'URRS3RZHU&RQWURO
models are presented to allow physical insights into the origin E F
of the droop instability. Then, Lyapunov function candidates
are proposed based on the reduced-order model of a simple Fig. 1. (a) System configuration of an inverter connected to the PCC; (b)
Internal control; (c) Droop (power) control.
two-bus example to obtain a concise stability criterion, fol-
lowed by further generalization to network settings. Finally,
Plug-and-Play (PnP) compliant design rules for choosing control as an equivalent series impedance (Rm , Lm shown
virtual and coupling impedances are obtained, offering great in Fig. 1(a)) connected between the inverter terminal and the
flexibility to incorporate the inverters of arbitrary ratings. The internal voltage sources (voltage reference) due to the slower
proposed approach allows for simple and easy integration of power control time-scale.
inverters into microgrids by satisfying only a set of local rules. It has been reported that setting high droop gains of the
The approach is then validated on a number of examples with power controller can lead to the instability of low frequency
models ranging from reduced-order to detailed ones. The main droop modes [1], [13], [22]. For a typical microgrid, the
contributions of this paper are summarized as follows: maximum allowed active power droop coefficient could be
1) Simple and reliable criteria for sizing of virtual significantly limited. Thus, efforts have been put into direct
impedance to achieve a desired stability region are eigenvalue analysis on the detailed models of high orders
derived and analyzed. to explore cumbersome and complicated stability conditions,
2) Plug-and-Play (PnP) compliant control rules for flexible where little insights were provided into the sources of insta-
integration of droop-controlled inverters into microgrids bility. In a previous work [12], the authors had revealed the
are developed. physical insights of the stability from a different perspective.
3) Comprehensive verification of the proposed approach Thus, in this section, we begin by providing a brief intro-
on detailed models is performed to demonstrate its duction of the basic framework to allow the development of
effectiveness and practicality. concise stability criteria for droop-controlled IBMGs in the
later section.
II. D ROOP I NSTABILITY OF I NVERTER -BASED M ICROGRID
A. Modeling of the Simple Two-Bus System
A typical configuration of a droop-controlled inverter con-
tains an external droop (power) controller, providing the Considering the point of common coupling (PCC) to be a
references for inverter internal control loops which regulate stiff voltage source as shown in Fig. 1(a), a droop-controlled
inverter voltage and frequency. The principle structure of such inverter can be described by the following set of ODEs with
an inverter is shown in Fig. 1 [1]. For the internal control respect to the synchronous reference frame of the stiff bus
loops, the measurement signals are refereed to the internal (Electromagnetic Model):
reference frame denoted by the superscript e ; the fast current θ̇ = ω − ω0 (1)
control loop performs high-bandwidth regulation of the filter
τ ω̇ = ωset − ω − kp ω0 P (2)
current; the filter current reference is generated from the
voltage controller that aims at regulating the terminal voltage. τ V̇ = Vset − V − kq Q (3)
In fact, various modified forms of voltage control have been LI˙d = V cos θ − Vs − RId + ω0 LIq (4)
proposed [14], [17], [20], [21]. Particularly, the emulation
LI˙q = V sin θ − RIq − ω0 LId (5)
of virtual impedance has been shown effective for different
purposes including stability enhancement, harmonics sharing, where θ is the relative angle of the internal voltage source with
reactive power sharing, etc. Along with the assumption of respect to the stiff bus, ω0 is the frequency of the stiff bus,
employing virtual impedance emulation for studying power τ = wc−1 is the time constant of the power filters, ωset and
sharing instability, it is sufficient to consider the internal Vset are the frequency and voltage set-points, respectively; ω

0885-8950 (c) 2018 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TPWRS.2019.2895081, IEEE
Transactions on Power Systems

and V are the actual frequency and voltage of the internal B. Stability Assessment
voltage source, respectively; P and Q are the active and A simple form of (6) allows for application of various
reactive power (before filtering), respectively; kp and kq are advanced stability assessment techniques. A method of Lya-
the per-unit frequency and voltage droop gains, respectively; punov functions is especially convenient - by finding a positive
L and R are the aggregations of the virtual, coupling and Lyapunov function V(x) > 0 for x 6= 0 (x being a vector of
line inductance (L = Lm + Lc + Ll ) and resistance (R = system states), the system stability can be certified if the time-
Rm + Rc + Rl ), respectively; Id and Iq are the inverter output derivative of this function is negative, i.e. V̇(x) < 0. Although
direct and quadrature currents in the synchronized reference obtaining an effective candidate function for arbitrary systems
frame, respectively. is usually a non-trivial task, for linear systems, quadratic
This 5th -order model assumes the combined LC filter Lyapunov function candidates can produce reasonable results.
and inverter internal control dynamics to be instantaneous. In order to get Lyapunov function candidates, we start by
Validation of this assumption will be provided in later sections. multiplying (6a) by (cτ ϑ̇ + ϑ) (c is some constant) to obtain:
Even this relatively simple 5th -order model is still not easy to ( )
analyze, therefore, in many works the left-hand sides of (4) d λp (τ ϑ̇ + ϑ)2 (cτ B − B 0 )ϑ2 (c − 1)τ 2 λp ϑ̇2
and (5) are set to zeros, resulting in a reduced-order model + +
dt 2 2 2
similar to Kuramoto’s oscillators which then can be analyzed
using a number of techniques. The usual justification for such + τ ((c − 1)λp − cB 0 )ϑ̇2 + Bϑ2 + Gϑρ − G0 ϑρ̇
an approximation is that the time constant of the network + cτ Gϑ̇ρ − cτ G0 ϑ̇ρ̇ = 0
dynamics (L/R) is typically much smaller (few milliseconds) (8)
than that of the power controller - τ - which is around Here, we have used the relations like 2ϑϑ̇ = d/dt(ϑ2 ) to
30ms. However, this type of approximation has been shown transfer some of the terms under the sign of full derivative.
to be inappropriate [11], [12], [22] for stability assessment Similarly, we multiply (6b) by (cτ ρ̇ + ρ) to obtain:
of IBMGs, where electro-magnetic phenomena play major
d ((c + 1)τ λq + cτ B − B 0 )ρ2
 
role, despite their small time-scale. Therefore, a more accurate 0
+ (cτ G + G )ϑρ
reduced-order model had been proposed in [12]: dt 2
(9)
+ (λq + B)ρ2 + c(τ 2 λq − τ B 0 )ρ̇2 − Gϑρ
λp τ ϑ̈ + (λp − B 0 ) ϑ̇ + Bϑ + Gρ − G0 ρ̇ = 0 (6a)
− G0 ϑρ̇ − cτ Gϑ̇ρ − 2cτ Gϑρ̇ + cτ G0 ϑ̇ρ̇ = 0.
(λq τ − B 0 ) ρ̇ + (λq + B)ρ − Gϑ + G0 ϑ̇ = 0 (6b)
Adding (8) and (9) together yields:
where ϑ , δθ, ρ , δV , λp = (kp ω0 )−1 , λq = kq−1 , and:
R X 1 dV
G= , B= 2 , (7a) = −Y (10)
R2 + X 2 R + X2 2 dt
2 2 where the Lyapunov function candidate V is a quadratic form
L(R − X ) 2LXR
G0 = 2 2 2
, B0 = . (7b) V = y T P y with y = [τ ϑ̇ + ϑ, ϑ̇, ϑ, ρ]T and matrix P being:
(R + X ) (R2 + X 2 )2
Model (6) allows one to see the role of electro-magnetic
 
λp 0 0 0
transients in instability - the term B 0 brings reduction to  0 (c − 1)τ 2 λ
p 0 0 
the system effective damping. A simple qualitative stability P =  (11)
 
0 0 cτ B cτ G
estimation can be inferred as λp − B 0 > 0, i.e. the sign of the

coefficient in front of the ϑ̇. A more rigorous expression will 0 0 cτ G (c + 1)τ λq + cτ B
be derived in the following sub-section. For the decay rate Y, likewise, we have the following quadratic
It should be noted that, as was demonstrated in [11], [12], form Y = z T Qz, where z = [ϑ, τ %̇, τ ϑ̇, %]T and matrix Q:
[23], for inverter-based microgrids, even significantly loaded  
conditions correspond to rather small angle differences and B −cG 0 0
voltage variations (due to small per unit values of network −cG cλ − c B 0
q 0 0 
Q = τ
 (12)
 
impedances). One of the consequences of this fact is that  0 0 c−1
λ − c 0
B 0
τ p τ

small-signal stability is almost independent of the operating
0 0 0 λq + B
point - this is in full agreement with the experimental results in
[13]. This fact allows for significant simplification of dynamic In the derivation of both V and Y, we have assumed that
equations and stability conditions - in fact equations (6) G0  τ G and B 0  τ B. In order for the system to be stable,
were derived using the condition of small angle differences both Lyapunov function V and decay rate Y have to be positive
and voltage variation for the inverter operating points. This definite, which is equivalent to positive definiteness of matrices
approximation is valid for any normal operating point and can P and Q from eqs. (11) and (12), respectively. Straightforward
be violated only for some extremely heavy loaded conditions calculations lead to the following conditions for P  0 (we
(line currents of several p.u. and more) that could possibly assume that c > 1):
appear in some emergency states of a microgrid. Consideration
of such abnormal states (not only from the point of view of λp > 0 (13a)
small-signal stability) is a separate problem that is beyond the c+1
λq + B − GB −1 G > 0 (13b)
scope of this work. c

0885-8950 (c) 2018 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TPWRS.2019.2895081, IEEE
Transactions on Power Systems

Likewise, matrix Q is positive definite if: 0.08 (6), 3rd-order Model


(1)-(5), 5th-order EM
c
B0

Voltage Droop k q
Fig. 1, Detailed Simulink Model
λp > (14a) 0.06
c−1 (17), Proposed Certificate
(16), c = 1.05
λq − B 0 /τ − cGB −1 G > 0 (14b) 0.04 (16), c = 1.5
(16), c = 3
λq + B > 0 (14c) (16), c = 10
0.02
If one assumes that λp > 0 and λq > 0 (which corresponds to
0
both frequency and voltage droop coefficients being positive), 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
the conditions in eqs. (13a) and (14c) are always satisfied. (a) Frequency Droop k p
Moreover, since c > 1, B > 0 and B 0 > 0, the condition in
0.08
(14b) is stricter than (13b). Therefore, positive definiteness of

Voltage Droop k q
both P and Q is guaranteed if two conditions, namely, eqs. 0.06
(14a) and (14b) are satisfied. Assuming that B 0  τ GB −1 G,
we can rewrite them in a simple way: 0.04

c
λp > B0, λq > cGB −1 G, (15) 0.02
c−1
In terms of frequency and voltage droop coefficients kp and 0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
kq , these relations are: (b) Frequency Droop k p
c−1 B
kp < , kq < , (16) 0.08
cω0 B 0 cG2
Voltage Droop k q
The effect of B 0 on the system stability is especially evident 0.06

from these equations, particularly, the active power-frequency 0.04


mode suffers significantly from the increase of B 0 . For every
value of constant c, equations (16) represent sufficient stability 0.02
condition corresponding to rectangular region on the plane of
0
kp and kq (see Fig. 2). Combining all those regions together - 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
varying c from 1 to infinity - one gets the following simplified (c) Frequency Droop k p
stability criterion:
Fig. 2. Small-signal stability regions of the two-bus system based on different
B 0 ω0 kp + Γkq < 1 (17) models with respect to the voltage and frequency droop gains (|Z| = 3%):
(a) X/R = 0.667; (b) X/R = 1; (c) X/R = 1.5.
where Γ = G2 /B.
The illustration of the above stability conditions is given
by Fig. 2 where the predicted stability regions for different C. Network Generalization
values of c from equations (16) as well as the total predicted The above derived framework can be naturally extended to
stability region from (17) are plotted in the coordinates of per- a network configuration. The most straightforward way to do
unit frequency and voltage droop coefficients. The stability this (detailed derivation can be found in [12]) is to switch to
boundaries obtained by direct numerical analysis of equations Laplace domain where network components are represented
(1) to (5) as well as (6) are also plotted. Moreover, we by their impedances as Zi = Ri + jXi + sLi and construct a
further provide the stability region predicted by the high-order Y-bus matrix of Kron-reduced system with only inverter nodes:
detailed model to illustrate the performance of the proposed
approximations (some parameters of the model are given in I(s) = Y(s)V(s) (18)
Section IV). Even though the kq predictions among all the
In time-domain, it can now be approximated:
cases are quite conservative, the practical reactive power droop
gains normally do not leverage the extended stability range I(t) = Y0 V(t) + Y1 V̇(t) (19)
(reactive power sharing suffers from unsymmetrical network
where Y0 = Y(s)|s=0 and Y1 = ∂Y(s)/∂s|s=0 . Then, the
configurations so that the droop gains are less effective to
relation (19) can be used to construct the generalized dynamic
allow designated sharing ratios). Also, it can be seen that
equations of a system with interconnected inverters [12]:
the reduced-order model along with the proposed stability
certificate tends to perform well when the X/R is around τ Λp ϑ̈ + (Λp − B0 )ϑ̇ + Bϑ + G% − G0 %̇ = 0 (20a)
unity. Therefore, in the later sections we will consider such 0 0
(τ Λq − B )%̇ + (Λq + B)% − Gϑ + G ϑ̇ = 0 (20b)
an observation for practical applications. Finally, the analysis
of (7) and (17) leads to a general conclusion, that the increase where ϑ, % are the vectorized states of inverter voltages and
of the impedance to the PCC extends the allowed region of angles; Λp and Λq are the diagonal  0 matrices containing
 0 0the
droop coefficients. In practice, the most binding term is B 0 inverse of droop gains; B = −Im Y ; G = Re Y ,B =
Im Y1 , and G0 = −Re Y1 . We note that B, G, and B0
 
- which we call transient susceptance - and it can result in
very tight restrictions on frequency droop gains for a typical are positive semi-definite matrices and G0 is a sign indefinite
microgrid configuration. matrix. It should be noted that in [24] it was concluded that the

0885-8950 (c) 2018 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TPWRS.2019.2895081, IEEE
Transactions on Power Systems

B' ,QY
50 0
y s
40
Value

30 0* ,QY1
20 y s yN s
1 2 3 4 5
,QY
Line Length (km)
y s
,QY
Fig. 3. Variation of B 0 ω0 and Γ with respect to the line length (Xmc /Rmc =
1 and Xl /Rl = 0.5). Fig. 4. Star network of a microgrid by aggregating the lines into a single
node.

typical load participation into the droop instability for IBMGs


is limited due to short lines of low impedance as compared operating point. For the reactive power droop gain, it is not
to much larger shunt components. We simplify the network much of an issue as can be seen from Fig. 2. In fact, the
representation by neglecting the shunt elements due to their proposed stability captures the kq bound by using Γ ∝ |Z|−1 :
limited influences in nominal operating conditions.
Starting from (20), the Lyapunov function and its decay rate G2 R2
Γ= = (25)
for the two-bus case can be generalized for networks - detailed B X|Z|2
derivations can be found in [25]. Thus, we obtain the Lyapunov
Although the predicted kq region is relatively conservative,
function as: V(y) = 0.5yT Py with y = [τ ϑ̇ + ϑ, ϑ̇, ϑ, %]T
high values of kq may be considered uncommon due to
and the decay rate Y(z) = zT Qz with z = [ϑ, τ %̇, τ ϑ̇, %]T .
induced voltage drops. This practical consideration enables
The corresponding P and Q matrices are given by:
the use of the proposed certificate that balances the active and
 
Λp 0 0 0 reactive droop gains in the same scale. To extend the stable
 0 (c − 1)τ 2 Λ 0 0  regions of the droop gains, adding a large amount of physical
p
P =
 
 impedance into the system is not practical due to the size and
 0 0 cτ B cτ G 
cost of required inductors/resistors. Therefore, emulation of
0 0 cτ G (c + 1)τ Λq + cτ B virtual components (e.g. inductance, reactance, or resistance)
(21) via high-bandwidth digital controllers have been commonly
  seen in literature. Consider now the representation of Z is
B −cG 0 0 reformulated as Z = Zmc + Zl where Zmc = Zm + Zc will
−cG cΛ − c B0 0 0  be referred to as controlled impedance - aggregation of the
q τ
Q =
 
 0 0 c−1
Λp − τc B0 0 
 virtual and coupling impedances (since they are the known
τ
and controllable parameters) and Zl is the line impedance to
0 0 0 Λq + B
the PCC. By choosing a particular Xmc /Rmc , it is possible
(22)
to show that B 0 and Γ are strictly decaying given the Xl /Rl
By demanding P  0 and Q  0, generalization of (15) is ratio of the cable/line. An example of using a certain line
obtained: characteristic is demonstrated in Fig. 3 [26]. This idea is very
c interesting and will serve as the core for deriving useful and
Λp − B0  0 (23a)
c−1 concise approximation for the stability certificate.
Λq − cGBg G  0 (23b)
where Bg denotes the generalized inverse of B with the A. Approximated Stability Rules
constraint (I − BBg )G = 0 satisfied.
Proceeding with the stability criteria of (23), one can search
for a set of virtual impedances for stable operation given the
III. S TABILITY E NHANCEMENT desired droop gains of an IBMG. However, these conditions
It can be seen in Fig. 2 that the system suffer from limited in (23) are not very intuitive for control engineers to tune
active power droop gain kp , which is mainly constrained by the desired virtual impedances. Therefore, in this section, an
transient susceptance B 0 . Thus, inspection of (7) shows that, in approximation is intended to obtain simple and concise rules
general, the transient susceptance B 0 is inversely proportional for tuning the virtual impedances. Since we know that it is
to the total impedance, B 0 ∝ (ω0 |Z|)−1 , and is given by: possible to choose a particular X/R ratio of the controlled
impedance to achieve the decaying property of B 0 and Γ with
2LXR 2X 2 R
B0 = = (24) respect to the additional line length (as illustrated by the Fig.3),
(X 2 + R2 )2 ω0 |Z|4
one can design the virtual impedances by assuming a worst-
This means that B 0 increases when the coupling between the case scenario, which is a set of inverters i ∈ N connected
inverter and PCC becomes stronger (e.g. connection with a to a PCC node only with their controlled impedances and no
short line) and a specific increase in the value of coupling lines, as shown in Fig. 4. Furthermore, we impose all inverters
impedance could be required to guarantee stability of the to have the same X/R ratio of their controlled impedances.

0885-8950 (c) 2018 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TPWRS.2019.2895081, IEEE
Transactions on Power Systems

Under this condition, the controlled admittance of inverter i B. Plug-and-Play Compliant Control Rules
can be written as: Equations (33) and (34) imply that the certification of
αi stability is approximated from network matrices to a set of
yi (s) = αi y(s) = (26)
Rmc + jXmc + sLmc local conditions for individual inverters. That is, this set of
distributed rules are given to allow the selection of the virtual
where y(s) is some reference admittance common for all
impedance independent of the network and other inverters.
inverters, and αi is the scaling factor for inverter i. Then, the
We refer to such functionality as the Plug-and-Play (PnP)
admittance matrix of the original network (inverters connected
compliant control - introduction of new inverters only needs
to PCC) can be represented as an (N + 1) × (N + 1) matrix
to satisfy these approximated local rules for droop stability.
YP CC (s) = y(s)ΨP CC , where ΨP CC denotes the following
Similar to the derivation of (17), equations (33) and (34)
matrix (we use the subscript P CC to refer to unreduced
produce the following criterion:
system):
g2
b0 ω0 αi kp,i +
αi kq,i < 1, ∀i ∈ N (35)
 
α1 · · · 0 −α1 b
 . .. .. .. 
 .. . . .  With the particular choice of Xmc = ω0 Lmc = Rmc (
ΨP CC =    (27)
 0 · · · αn −αn 
 Xl /Rl = 0.5), substituting b0 , g, and b in eqs. (29)-(31), (35)
can be further simplified to:
−α1 · · · −αn A
kp,i + kq,i < 2Xmc,i , ∀i ∈ N (36)
Here, the top-left block represents a diagonal N ×N matrix
with elements αi (we will denote it as Φ), and A is the sum where Xmc,i = Xmc /αi and the equation becomes fairly
of αi over all i. Since the Kron reduction only affects matrix simple to use for sizing the virtual impedances. Also, the
ΨP CC , the reduced N × N admittance matrix Y(s) becomes: choice of Xmc /Rmc = 1 can further benefit the performance
of the stability rules and the fidelity of the reduced-order model
as depicted in Fig. 2.
Y(s) = y(s)Ψ; Ψ = Φ − A−1 aaT (28) To verify the proposed control rules, a simple five-inverter
to PCC system is tested with arbitrary line lengths as shown
where a = [α1 · · · αN ]T . The convenience of representation
in Table I. The projected stability region with respect to the
(28) lies in the fact that all the matrices B, G and B0 from
universal droop gains (kp and kq are the same for all inverters)
(23) are proportional to the same matrix Ψ:
are shown in Fig. 5. First, it can be identified that case 1 with
Xmc the greatest scale gives the largest projected stability region;
B = bΨ; b= 2 + X2
(29)
Rmc mc however, whenever there are short paths (between inverter 1
and inverter 4) the projected stability region becomes bottle-
Rmc necked by such connections so that case 3 and case 2 are
G = gΨ; g= 2 + X2
(30) almost identical. In fact, this issue has been identified in [22]
Rmc mc
for multi-microgrid stability and referred as ”critical links.” It
should also be emphasized that Fig. 5 may not be sufficient to
2Lmc Xmc Rmc
B0 = b0 Ψ; b0 = 2 + X 2 )2
(31) reflect the actual multi-dimensional stability region (volume)
(Rmc mc since the droop gains are constrained to be equal for all
Therefore, we have GBg G = (g 2 /b)Ψ in (23b). Additional inverters.
simplification comes from the fact that the matrix aaT is Therefore, in Fig. 6 the stability boundaries from different
always positive semi-definite, therefore, conditions (23) can perspectives are plotted with respect to the active power droop
be substituted by more conservative ones: gains of inverter 1 and inverter 4 (kp1 , kp4 ) to show the
effect of the critical links: the remaining droop gains are
cb0 fixed as described and one can imagine slicing Fig. 5 based
Λp − Φ0 (32a)
c−1 on different kq gains. First, significant reduction of stability
cg 2 regions is mainly affected by the distances between inverter
Λq − Φ  0. (32b)
b 1 and inverter 4 by comparing case 1 and case 2 (zeroing
All matrices in equations (16) are diagonal and the positive line 1 and line 4), whereas the region reduction becomes
definiteness can be certified by considering separate conditions insignificant from shorting other lines (between case 2 and
for every diagonal element, i.e: case 3), indicating the effect of the critical links. Similar to
Fig. 6, the proposed PnP rules show further conervativeness
cb0 as kq increases. In Fig. 7 (kq,i = 0%), case 4 demonstrates
λp,i ≥ αi , ∀i ∈ N (33)
c−1 the greatest area due to infinite lengths of lines 2, 3 and 5
(such a condition deduces to a two-inverter system). By adding
cg 2 inverters back and reducing their lines to the PCC (case 2
λq,i ≥ αi , ∀i ∈ N (34)
b and case 3, sequentially), it can be seen that the stability
where every condition is directly analogous to equations (15) region reduces and is eventually of a square shape. In fact,
for the two-bus case. the proposed PnP rules predict the same boundary as case

0885-8950 (c) 2018 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TPWRS.2019.2895081, IEEE
Transactions on Power Systems

0.07 0.08
Case 1, Eq. (20) Case 2, (20)
0.06 Case 3, (20)
Case 2, Eq. (20) 0.06
Case 4, (20)
Voltage Droop k q

0.05 Case 3, Eq. (20)


Proposed, Eq. (36) Proposed, (36)

kp4
0.04 0.04

0.03
0.02
0.02
0
0.01 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
0 kp1
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07
Frequency Droop k p Fig. 7. Comparison of projected stability regions with respect to kp1 , kp4 :
Xmc,i = Rmc,i = 2%, kp2 = kp3 = kp5 = 4%, kq,i = 0% (line
Fig. 5. Comparison of low-dimensional projected stability regions of (35) and parameters are shown in Table I).
(20) for a microgrid with five inverters (n = 5, star topology) of different
scales with symmetric droop gains and controlled impedances: kp,i = kp ,
kq,i = kq , Xmc,i = Rmc,i = 2% (scale parameters are shown in Table I). &XUUHQW VRXUFH Zt s Vv I
Vc VPCC Z t s   VPCC
)HHGIRUZDUG &RXSOLQJDQG/LQH

I I If  Ic
 Cif s Z s s
f
u   Vc
If Z s s
TABLE I &XUUHQW 6KXQW &DS
D ISTANCES IN K ILOMETERS TO PCC OF F IVE -I NVERTER T EST C ASES : &RQWUROOHU
Z s  
Xl = 0.28%km−1 , Rl = 0.56%km−1
)HHGIRUZDUG
D E

Case line 1 line 2 line 3 line 4 line 5


1 0.5 3.0 2.5 0.7 1.5 Fig. 8. (a) Filter current injection into the PCC; (b) Block diagrams of the
2 0.0 3.0 2.5 0.0 1.5 internal control.
3 0.0 0.0 0.0 0.0 0.0
4 0.0 ∞ ∞ 0.0 ∞
C. Power Sharing of Multiple Inverters
The power sharing for multiple inverters depends on their
relative droop gains. While the reactive power sharing is sen-
0.08
Case 1, (20) Case 2, (20) Case 3, (20) Proposed, (36) sitive to the network characteristics, the active power sharing,
kq ,i = 0% in general, depends on (2) since the frequency is a universal
0.06
signal. Therefore, in steady-state, one can obtain (neglecting
k p4

0.04
non-ideal frequency errors):
0.02

0
kp,1 P1 = kp,2 P2 = ... = kp,i Pi , ∀i ∈ N (37)
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08
(a) k p1 It can be seen that the power sharing is inversely proportional
to the droop gain. By defining kp,i = kp /si with kp being
0.08
kq ,i = 1% the common droop base and si denoting the sharing ratio, the
0.06
equality reduces to:
k p4

0.04
P1 P2 Pi
0.02 = = ... = , ∀i ∈ N (38)
s1 s2 si
0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 In fact, this notation of sharing ratio si becomes convenient
(b) k p1
for investigating the stability region. If we choose the notation:
0.08 kp,i = kp /si , kq,i = kq /si , and Xmc,i = Rmc,i = Xmc /si ,
kq ,i = 2%
0.06 where kp , kq , and Xmc denote the parameters bases. Then,
(36) can be rewritten as:
k p4

0.04

0.02 kp /si + kq /si < 2Xmc /si , ∀i ∈ N (39)


0
0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 The above equation indicates that a smaller sharing ratio
(c) k p1 results in larger droop gains and therefore requires larger
virtual impedance. Therefore, in the later sections, we are
Fig. 6. Comparison of projected stability regions with respect to kp1 , kp4 :
Xmc,i = Rmc,i = 2%; (a) kp2 = kp3 = kp5 = 4%, kq,i = 0%; (b)
to abuse the usage of si for convenience due to difficulty in
kp2 = kp3 = kp5 = 3%, kq,i = 1%; (c) kp2 = kp3 = kp5 = 2%, visualizing multi-dimensional stability regions.
kq,i = 2% (line parameters are shown in Table I).
IV. I MPLEMENTATION OF V IRTUAL I MPEDANCES
A. Adding Impedance to Inverter Control Loops
4 indicating conservative predictions against the instability Virtual impedance control has been thoroughly investigated
resulted from the critical links. in literature [14], [17], [20], [21], [27]. In this work, we

0885-8950 (c) 2018 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TPWRS.2019.2895081, IEEE
Transactions on Power Systems

&XUUHQW VRXUFH Zt s
Z m s Vc

Mangnitude
100
9LUWXDO
,PSHGDQFH
I I
Vv
Vc Vc I sC f 0 km 2 km 4 km

10-2
-5 -4 -3 -2 -1 0 1 2 3 4 5

200

Angle (degree)
Fig. 9. Ideal equivalent circuit of the virtual impedance control.
0

aim to focus on the control structure that provides direct -200


emulation of dynamic behavior of impedances without an -5 -4 -3 -2 -1 0 1 2 3 4 5
sgn( )log10(| |) (rad/s)
explicit voltage control loop. Since d and q axes are coupled,
it is not straightforward to analyze each axis separately. Thus, Fig. 10. Bode plots of the transfer function from I ∗ to Vc with different line
it is convenient to employ complex transfer functions such lengths.
that both voltage and current become phasors and transfer-
functions of control system are now complex [28]. We denote
that the impedance of a typical RL component is written in 1
0.985
the form:

Voltage (pu)
0.968
0.95
Z(s) = R + jωL + sL (40) X m c = R m c = 2%
X m c = R m c = 4%
where by setting s to zero, Z(0) indicates the steady-state X m c = R m c = 6%
0.87
impedance. First, we consider that equivalent circuit of the
filter current injection into the PCC, as shown in Fig. 8(a), 0 1 2 3 4 5 6 7 8 9 10
Time (ms)
where Zt (s) = Zc (s) + Zl (s) denotes the aggregated coupling
and line impedance, and Zs (s) = (sCf + jωCf )−1 denotes Fig. 11. Output voltage (Vc ) responses for a sudden load change (1 pu
the shunt impedance (filter capacitors). The filter current If resistive) with different controlled impedances.
is generally regulated with high bandwidth current controller;
thus, it is convenient and simpler to assume If = If∗ for
derivations so that the transfer function from If∗ to Vc becomes: (illustrated in Fig. 9), the output current reference can be
computed by measuring the terminal voltage Vc :
Zt (s)Zs (s) ∗
Vc = I (41) Vc∗ − Vc
Zt (s) + Zs (s) f I∗ = (44)
Zm (s)
Then, the filter current reference can given as:
where Zm (s) = sLm + Rm + jXm denotes the virtual
If∗ = u + σv I + Zs (0)−1 Vc (42)
impedance. It should be emphasized that there are many pos-
where u denotes the temporary reference input, σv is the out- sible alternative forms in the literature [14], [20], [21]. In this
put current feed-forward gain, and Zs (0) = (jωCf )−1 is the work, we focus on direct emulation of the dynamical responses
steady-state shunt impedance of the capacitor filter (physical of the impedance to serve for the proposed framework. Also, it
dampers can also be compensated whenever deployed). The is very common to employ virtual damper circuits (along with
last two terms of (42) are normally referred as the feed- physical damper circuits) to mitigate the resonance of the CL
forward compensation, which aims at canceling the known network. The general idea is to add an RC filter in parallel to
disturbances of output current and the capacitive current in the output capacitor by adding the damping current into the
synchronized reference frame. Thus, it is interesting to explore current references:
the open-loop characteristics from u to Vc , which can be s/Rvd
depicted using the block diagram shown in Fig. 8(b). By ∆I ∗ = Vc (45)
s + 1/(Rvd Cvd )
using the standard loop deduction technique, we can obtain
the synthesized form of: where Rvd and Cvd are the virtual resistance and capacitance,
respectively. With the virtual damper control and considering
Zs0 (s)Zt (s) ∗ Z 0 (s)
Vc = 0
I + 0 s VP CC (43) the filter current regulation, the bode plots of the transfer
Zs (s) + Zt (s) Zs (s) + Zt (s) function from I ∗ to Vc based on different line lengths are
where Zs0 (s) = (1 − σv )/sCf and I ∗ = u/(1 − σv ) is the shown in Fig. 10. Since we are mainly interested in the low
output current reference. In fact, (43) represents an equivalent frequency droop mode of around 10 Hz, the shunt elements
circuit shown in Fig. 9. are quite insignificant in affecting droop stability due to their
The above investigation shows that the control structure greater values as compared to the series elements around the
allows to regulate the output current into the PCC (the shunt frequency of the droop mode. Effectively, we consider only
element becomes infinite in steady-state). Thus, to emulate the series impedances for stability enhancement, which will be
dynamic behavior of the RL component connected in series validated in the later section.

0885-8950 (c) 2018 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TPWRS.2019.2895081, IEEE
Transactions on Power Systems

B. Voltage and Power Quality


,QYHUWHU
Employing large virtual impedance helps to mitigate the
droop instability and allows more space in terms of tuning
frequency and voltage droop coefficients. However, the steady- Rf Lf Rc Lc
Cf /RDG
state voltage variation is also affected (the transient voltage Rl 
dips are more relevant to the internal inverter control and filter
,QYHUWHU Ll 
parameters). Based on the parameters shown in Table II, the
output voltage responses for a single inverter under resistive
load stepping of 1 pu for different controlled impedances Rf Lf
(Zmc = Zm +Zc ) are shown in Fig. 11. It is evident from Fig. C f Rc Lc /RDG
Rl 
11 that the values of transient dip are similar for all three cases
- around 0.87 pu. The steady-state voltage drops, however, are ,QYHUWHU Ll 
significantly different (98.5%, 96.8%, 95%) and depend on the
amount of virtual impedance. Therefore, without secondary
compensations, compromise needs to be made between sharing Rf Lf Rc Lc
Cf /RDG
accuracy, droop stability, and voltage/frequency variations,
which are typical for droop-controlled sharing schemes.
Under heavy loading conditions, the maximum amount of
virtual impedance can be limited which leads to certain limits Fig. 12. Three-inverter test system.
on the maximum values of droop coefficients. Since the power
sharing is determined by the ratio between droop coefficients 0.07
according to (38), but not on their absolute values, it is possible 0.06
Reduced-order Model, eq. (20)
Simulation Model
to achieve desired sharing by a simultaneous decrease of all Voltage Droop k q
Proposed PnP Rules
0.05
kp gains. However, as the droop coefficients get smaller, the Experiment
0.04 Sampled Points
errors of frequency references can become crucial. This could
lead to inaccurate power sharing between devices according to 0.03

P = ω /kp , where P denotes the power sharing errors and A


0.02
w the frequency references errors. For example, for a small 0.01
droop gain of 0.5%, the frequency reference error of 0.05- B
0
0.01% (application dependent) results in active power sharing 0 0.01 0.02 0.03 0.04 0.05 0.06 0.07
deviations of 10-2%. Therefore, there is a certain limit on the Frequency Droop k p
minimum values of the droop coefficients based on the desired
power sharing accuracy. In general, for voltage/frequency Fig. 13. Comparison of projected stability regions: kp,i = kp /si , kq,i =
kq /si , Xmc,i = Rmc,i = 2/si %.
sensitive applications, the secondary compensation is recom-
mended to resolve such an issue [3].
V. VALIDATION (resistive load), 0.4 + j0.71 (inductive and resistive load),
In this section, the simulation models with the detailed 0.57 + j0.0 (constant active power load) for loads 1, 2 and
implementation of control loops in MATLAB/Simulink along 3, respectively. The projected stability region with respect
with a miniature experimental prototype are used to verify to the common droop bases kp and kq are shown, with
the effectiveness of the proposed approach. For computing the individual droop gains and controlled impedance scaled
the stability region of the Simulink model, linmod and trim according to their sharing ratios (kp,i = kp /si , kq,i = kq /si ,
functions are used to extract linearized transition matrices and Xmc,i = Rmc,i = 2/si %). It should be emphasized that the
eigenvalues. In addition, the experimental prototype was tested proportional scaling of the droop gains and virtual impedances
by using ST STEVAL-IHM028V2 inverters, self-designed are just for plotting convenience according to (39) and one
filter boards, and the control boards with TI C28379D micro- may choose any values within the upper bounds of the droop
controllers. Also, the constant power load is also considered gains (lower bounds of the virtual impedances when fixing
by using the commercial PFC power supplies along with droop gains, vice versa). The results show that the predicted
transformers for voltage adjustment since the base power and boundaries between the reduced-order model, detailed simula-
voltage are both scaled down for conceptual validation. Both tion model, and experiment coincide closely with very small
the simulation and experiment shares the same base impedance deviations. Moreover, we further verify the effect of critical
and therefore the parameters are intended to be identical in pu link by plotting the stability region with respect to the active
system. The system configuration is shown in Fig. 12 with the power droop gains of inverter 1 and inverter 2, shown in
basic settings and parameters described in Table II. Fig. 14. In this case, other parameters are fixed as described
in the caption with the only variables being kp1 and kp2 .
A. Stability Regions Similarly, the predicted boundaries are very close among all
In Fig. 13, we consider that the sharing ratios are: s1 : s2 : models, verifying the investigation in the previous section. It
s3 = 1 : 0.67 : 0.33 and the load types (pu): 1.0 + j0.0 should also be emphasized that the reduced-order model does

0885-8950 (c) 2018 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TPWRS.2019.2895081, IEEE
Transactions on Power Systems

0.1 k q,i = 0%
0.08
0.06
k p2

0.04
0.02

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08


(a) k p1

0.1 k q,i = 2/s i%


Reduced-order Model, (20)
0.08 Simulation Model
0.06 Proposed PnP Rules
k p2

Experiment
0.04

0.02

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08


(b) k p1

Fig. 14. Comparison of projected stability regions with respect to kp1 and
kp2 : s1 : s2 : s3 = 1 : 0.67 : 0.33, Xmc,i = Rmc,i = 2/si %; (a)
kq,i = 0, kp3 = 12.12%; (b) kq,i = 2/si %, kp3 = 6.06%.

not consider load participation and can still generate accurate


predictions.
The above results validate that: (1) load participation under
small virtual impedance is limited; (2) the projected stability
region is constrained by the short link between bus 1 and
bus 2; (3) the reduced-order model achieves high prediction
accuracy; (4) the proposed PnP control rule demonstrates
conservativeness when kq increases (Fig. 14(b)).

B. Time-domain Analysis
In addition to the validation of stability regions, the time-
domain analysis is carried out based on both experiment and
simulation. Particularly, we choose a relatively large droop
gains to show that the virtual impedance method can effec-
tively damp the droop oscillations. Under the same loading
condition as the previous subsection, the system dynamical
responses after a step change of sharing ratios from 1 : 1 : 1
to 1 : 0.67 : 0.33 are shown in Fig. 15. In this case, both
droop gains and virtual impedances are of a step change at
0.1s - kp,i = kq,i = 2/si % and Xmc,i = Rmc,i = 2/si %.
It can be seen that active and reactive power sharing changes
according to the sharing ratios; the oscillation is damped due
to the conservative nature of the proposed PnP rule; the voltage
and frequency drop due to the overall increase of the virtual
impedances and droop gains; the dynamic responses between Fig. 15. Time-domain results for a step change of sharing ratios, si , form 1 :
1 : 1 to 1 : 0.67 : 0.33: kp,i = kq,i = 2/si %; Xmc,i = Rmc,i = 2/si %;
the simulation and experiment match closely despite some DC (a) Output voltages (Vc ) ; (b) Filtered active power; (c) Filtered reactive power;
deviations due to calibration errors and un-modeled component (d) Frequency.
tolerances.
Fig. 16 shows the oscillations around two sampled critically
stable operating conditions (sampled points, A and B, in In Fig. 17, voltage responses for a sudden resistive load
Fig. 13, difference of the critical droop gains between the stepping of 2 pu based on different values of controlled
simulation and experiment is ∼ 0.1%). It can be seen that impedances are demonstrated. Similarly, discrepancies be-
choosing higher active power droop gains result in higher tween the simulation and experiment can be observed due to
frequencies of oscillations. Also, the critical link between measurement noises, calibration errors, and un-modeled digital
inverters 1 and 2 leads to a dominant magnitude of oscillation delays. In general, the transient voltage dips for few hundred
between them as compared to inverter 3. microseconds in simulation are around 0.88 to 0.94 pu (0.85 to

0885-8950 (c) 2018 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TPWRS.2019.2895081, IEEE
Transactions on Power Systems

PV

PV

Fig. 16. Droop oscillations beyond critically stable conditions: sharing ratios
1 : 0.67 : 0.33, Xmc,i = Rmc,i = 2/si %; (a) kp,i =∼ 2.8/si %, kq,i =
2/si %; (b) kp,i =∼ 4.3/si %, kq,i = 0/si %. Fig. 17. Dynamic responses of output voltages (Vc ) during load stepping:
sharing ratios 1 : 0.67 : 0.33, kp,i = 2/si %, kq,i = 2/si %; (a) Xmc,i =
Rmc,i = 2/si % (b) Xmc,i = Rmc,i = 5/si %.
0.93 pu for experiment) and are not sensitive to the controlled
impedance since it is mainly affected by the internal control,
network characteristics, and types and locations of the load.
In steady-state, it can be seen that the voltage variations are
proportional to the value of controlled impedance as expected.
The filtered active power responses for the same test is shown
in Fig. 18. The responses are almost identical regardless of the
discrepancies of voltage responses due to large filtering time
constants.

VI. C ONCLUSION
Specific nature of microgrids where per-unit line impedance
values are very small and X/R ratio is around unity leads
to instabilities when operating in droop-controlled mode. The
instabilities are mainly caused by electro-magnetic delays in
lines and are especially pronounced with reduction in network
size. We analyzed in details a possible solution for this issue in
the form of virtual impedance emulated by inverter controls.
The sizing of the virtual impedance given a choice of droop
gains is derived and investigated to ensure the stability of the
sharing operations. It was shown that the stability criteria for
an IBMG can be approximated into a set of local conditions,
allowing Plug-and-Play compliant control and local decision
making for sizing the virtual impedance. Finally, our analytical
and numerical findings are validated through detailed models
in MATLAB/Simulink and experiment. Fig. 18. Dynamic responses of filtered active power during load stepping:
sharing ratios 1 : 0.67 : 0.33, kp,i = 2/si %, kq,i = 2/si %; (a) Xmc,i =
R EFERENCES Rmc,i = 2/si % (b) Xmc,i = Rmc,i = 5/si %.
[1] N. Pogaku, M. Prodanović, and T. C. Green, “Modeling, analysis and
testing of autonomous operation of an inverter-based microgrid,” IEEE
Trans. Power Electron., vol. 22, no. 2, pp. 613–625, 2007.

0885-8950 (c) 2018 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TPWRS.2019.2895081, IEEE
Transactions on Power Systems

TABLE II low-voltage multibus microgrid,” IEEE Trans. Power Electron., vol. 24,
PARAMETERS OF T HREE I NVERTER TEST SYSTEM no. 12, pp. 2977–2988, 2009.
[16] J. M. Guerrero, L. G. de Vicuna, J. Matas, M. Castilla, and J. Miret,
“Output impedance design of parallel-connected ups inverters with
Parameter Description Value wireless load-sharing control,” IEEE Trans. Ind. Electron., vol. 52, no. 4,
Vb Base Voltage (simulation) 381 V pp. 1126–1135, Aug. 2005.
Sb Base Power (simulation) 5 kVA [17] J. He, Y. W. Li, J. M. Guerrero, F. Blaabjerg, and J. C. Vasquez,
Vb Base Voltage (prototype) 71 V “An islanding microgrid power sharing approach using enhanced virtual
Sb Base Power (prototype) 174 VA impedance control scheme,” IEEE Trans. Power Electron., vol. 28,
Zb Base Impedance 29 Ω no. 11, pp. 5272–5282, Nov. 2013.
[18] H. Tian, Y. W. Li, and P. Wand, “Hybrid ac/dc system harmonics control
ω0 Nominal Frequency 2 π×50 rad/s through grid interfacing converters with low switching frequency,” IEEE
wc Frequency Filter Constant 2 π×5 rad/s Trans. Ind. Electronic, in press.
fsw Switching Frequency 10 kHz [19] W. Yao, M. Chen, J. Matas, J. M. Guerrero, and Z.-M. Qian, “Design and
σi voltage feed-forward gain 1 analysis of the droop control method for parallel inverters considering
σv current feed-forward gain 0.5 the impact of the complex impedance on the power sharing,” IEEE
Lc Coupling Inductance 0.35 mH Trans. Ind. Electron., vol. 58, no. 2, pp. 576–588, 2011.
[20] X. Wang, Y. W. Li, F. Blaabjerg, and P. C. Loh, “Virtual-impedance-
Rc Coupling Resistance 80 mΩ
based control for voltage-source and current-source converters,” IEEE
Lf Filter Inductance 1 mH Trans. Power Electron., vol. 30, no. 12, pp. 7019–7037, 2015.
Cf Filter Capacitance 30 µF [21] J. Kim, J. M. Guerrero, P. Rodriguez, R. Teodorescu, and K. Nam,
Rvd Virtual Damping Resistance 5 ohm “Mode adaptive droop control with virtual output impedances for an
Cvd Virtual Damping Capacitance 0.25 mF inverter-based flexible ac microgrid,” IEEE Trans. Power Electron.,
kpc , kic Current Loop PI Gains 8, 18 000 vol. 26, no. 3, pp. 689–701, 2011.
Ll Line Inductance 0.26 mH km−1 [22] I. P. Nikolakakos, H. H. Zeineldin, M. S. El-Moursi, and N. D. Hatziar-
gyriou, “Stability evaluation of interconnected multi-inverter microgrids
Rl Line Resistance 165 mΩ km−1
through critical clusters,” IEEE Trans. Power Syst., vol. 31, no. 4, pp.
l12 Line Length (Bus 1 to 2) 0.6 km 3060–3072, 2016.
l23 Line Length (Bus 2 to 3) 3.25 km [23] X. Guo, Z. Lu, B. Wang, X. Sun, L. Wang, and J. M. Guerrero,
“Dynamic phasors-based modeling and stability analysis of droop-
controlled inverters for microgrid applications,” IEEE Trans. Smart Grid,
vol. 5, no. 6, pp. 2980–2987, 2014.
[24] N. Bottrell, M. Prodanovic, and T. C. Green, “Dynamic stability of a
[2] D. E. Olivares et al., “Trends in microgrid control,” IEEE Trans. Smart microgrid with an active load,” IEEE Trans. Power Electron., vol. 28,
Grid, vol. 5, no. 4, pp. 1905–1919, 2014. no. 11, pp. 5107–5119, 2013.
[3] J. M. Guerrero, J. C. Vasquez, J. Matas, L. G. De Vicuña, and [25] P. Vorobev, P.-H. Huang, M. Al Hosani, J. L. Kirtley, and K. Turitsyn,
M. Castilla, “Hierarchical control of droop-controlled ac and dc mi- “A framework for development of universal rules for microgrids stability
crogrids—a general approach toward standardization,” IEEE Trans. Ind. and control,” in 2017 IEEE 56th Annual Conference on Decision and
Electron., vol. 58, no. 1, pp. 158–172, 2011. Control (CDC). IEEE, 2017, pp. 5125–5130.
[4] E. Planas, A. Gil-de Muro, J. Andreu, I. Kortabarria, and I. M. [26] ——, “Towards plug-and-play microgrids,” in IECON 2018-44th Annual
de Alegrı́a, “General aspects, hierarchical controls and droop methods Conference of the IEEE Industrial Electronics Society. IEEE, 2018,
in microgrids: A review,” Renewable and Sustainable Energy Reviews, pp. 4063–4068.
vol. 17, pp. 147–159, 2013. [27] P.-H. Huang, P. Vorobev, M. Al Hosani, J. L. Kirtley, and K. Turitsyn,
[5] M. C. Chandorkar, D. M. Divan, and R. Adapa, “Control of parallel “Systematic design of virtual component method for inverter-based
connected inverters in standalone ac supply systems,” IEEE Trans. Ind. microgrids,” in Power & Energy Society General Meeting, 2017 IEEE.
Appl., vol. 29, no. 1, pp. 136–143, 1993. IEEE, 2017, pp. 1–5.
[6] N. Hatziargyriou, Microgrids: architectures and control. John Wiley [28] S. Gataric and N. R. Garrigan, “Modeling and design of three-phase
& Sons, 2013. systems using complex transfer functions,” in 30th Annual IEEE Power
[7] J. W. Simpson-Porco, F. Dörfler, and F. Bullo, “Droop-controlled in- Electronics Specialists Conference. IEEE, 1999, pp. 691–697.
verters are kuramoto oscillators,” IFAC Proceedings Volumes, vol. 45,
no. 26, pp. 264–269, 2012.
[8] Y. Zhang and L. Xie, “Online dynamic security assessment of microgrid
interconnections in smart distribution systems,” IEEE Trans. Power Syst.,
vol. 30, no. 6, pp. 3246–3254, 2015.
[9] E. Coelho, P. Cortizo, and P. Garcia, “Small-signal stability for parallel-
connected inverters in stand-alone ac supply systems,” IEEE Trans. Ind.
Appl., vol. 38, no. 2, pp. 533–542, 2002.
[10] J. M. Guerrero, L. G. De Vicuna, J. Matas, M. Castilla, and J. Miret, “A
wireless controller to enhance dynamic performance of parallel inverters
in distributed generation systems,” IEEE Trans. Power Electron., vol. 19,
no. 5, pp. 1205–1213, 2004.
[11] V. Mariani, F. Vasca, J. Vásquez, and J. Guerrero, “Model order reduc-
tions for stability analysis of islanded microgrids with droop control,”
IEEE Trans. Ind. Electron., vol. 62, no. 7, pp. 4344–4354, 2015.
[12] P. Vorobev, P.-H. Huang, M. Al Hosani, J. L. Kirtley, and K. Turitsyn,
“High-fidelity model order reduction for microgrids stability assess-
ment,” IEEE Trans. Power Syst., vol. 33, no. 1, pp. 874–887, 2018.
[13] E. Barklund, N. Pogaku, M. Prodanovic, C. Hernandez-Aramburoand,
and T. C. Green, “Energy management in autonomous microgrid using
stability-constrained droop control of inverters,” IEEE Trans. Power
Electron., vol. 23, no. 5, pp. 2346–2352, 2008.
[14] J. He and Y. W. Li, “Analysis, design, and implementation of virtual
impedance for power electronics interfaced distributed generation,” IEEE
Trans. Ind. Appl., vol. 47, no. 6, pp. 2525–2538, Nov. 2011.
[15] Y. W. Li and C. N. Kao, “An accurate power control strategy for
power-electroncis-interfaced distributed generation units operating in a

0885-8950 (c) 2018 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Citation information: DOI 10.1109/TPWRS.2019.2895081, IEEE
Transactions on Power Systems

Po-Hsu Huang (SM’12) received the B.Sc. degree Konstantin Turitsyn (M‘09) received the M.Sc.
from National Cheng-Kung University, Tainan, Tai- degree in physics from Moscow Institute of Physics
wan, and his first M.Sc. degree from National Tai- and Technology and the Ph.D. degree in physics
wan University, Taipei, Taiwan, in 2007 and 2009, from Landau Institute for Theoretical Physics,
respectively, both in electrical engineering. He also Moscow, in 2007. He was an Associate Profes-
attained his second M.Sc. degree from Department sor at the Mechanical Engineering Department of
of Electrical Power Engineering, Masdar Institute of Massachusetts Institute of Technology (MIT), Cam-
Science and Technology, Abu Dhabi, United Arab bridge. Before joining MIT, he held the position of
Emirates. He received the Ph.D. degree from the Oppenheimer fellow at Los Alamos National Lab-
Department of Electrical Engineering and Computer oratory, and Kadanoff–Rice Postdoctoral Scholar at
Science at Massachusetts Institute of Technology, University of Chicago. His research interests encom-
USA, in 2019. His current interests include photovoltaic power systems, pass a broad range of problems involving nonlinear and stochastic dynamics
DC/AC Microgrids, power electronics, wind power generation, linear/non- of complex systems. Specific interests in energy related fields include stability
linear system dynamics, power system stability, and control. and security assessment, integration of distributed and renewable generation.

Petr Vorobev (M‘15) received his Ph.D. degree in


theoretical physics from Landau Institute for Theo-
retical Physics, Moscow, in 2010. Currently, he is
an Assistant Professor at Skolkovo Institute of Sci-
ence and Technology (Skoltech), Moscow, Russia.
Before joining Skoltech, he was a Postdoctoral As-
sociate at the Mechanical Engineering Department
of Massachusetts Institute of Technology (MIT),
Cambridge. His research interests include a broad
range of topics related to power system dynamics
and control. This covers low frequency oscillations
in power systems, dynamics of power system components, multi-timescale
approaches to power system modelling, development of plug-and-play control
architectures for microgrids.

Mohamed Al Hosani (S‘10M‘13) received the


B.Sc. degree in electrical engineering from the
American University of Sharjah, UAE, in 2008,
and the M.Sc. and the Ph.D. degrees in electrical
engineering from the University of Central Florida,
Orlando, FL, USA, in 2010 and 2013, respectively.
He was an Assistant Professor in the Department
of Electrical and Computer Engineering, Khalifa
University, Abu Dhabi, UAE, in 2014-2018, and
a Visiting Assistant Professor at the Massachusetts
Institute of Technology, Cambridge, MA, USA, for
8 months in 2015-2016. He is currently the Department Manager of Demand
Side Management (DSM) at Abu Dhabi Distribution Company, Abu Dhabi,
UAE. His interests demand side management, distributed generation protection
and control, modeling and stability analysis of microgrid and smart grid.

James L. Kirtley Jr. (F’90) is of Professor of


Electrical Engineering at the Massachusetts Institute
of Technology. He was with General Electric, Large
Steam Turbine Generator Department, as an Elec-
trical Engineer, for Satcon Technology Corporation
as Vice President and General Manager of the Tech
Center, as Chief Scientist and as Director Dr. Kirtley
was Gastdozent at the Swiss Federal Institute of
Technology, Zürich (ETH). Dr. Kirtley attended MIT
as an undergraduate and received the degree of Ph.D.
from MIT in 1971. Dr. Kirtley is a specialist in
electric machinery and electric power systems. He served as Editor in Chief
of the IEEE Transactions on Energy Conversion from 1998 to 2006. Dr.
Kirtley was made a Fellow of IEEE in 1990. He was awarded the IEEE
Third Millennium medal in 2000 and the Nikola Tesla prize in 2002. Dr.
Kirtley was elected to the United States National Academy of Engineering in
2007. He is a Registered Professional Engineer in Massachusetts.

0885-8950 (c) 2018 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission. See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.

You might also like