Advanced Control and Protection of Modular UPS
Advanced Control and Protection of Modular UPS
Baoze Wei
Energy Technology, Aalborg University, Aalborg East, Denmark
Xiaochao Hou
Tsinghua University, Beijing, China
Yao Sun
Information Science and Engineering, Central South University,
Changsha, China
This work is subject to copyright. All rights are solely and exclusively
licensed by the Publisher, whether the whole or part of the material is
concerned, speci ically the rights of translation, reprinting, reuse of
illustrations, recitation, broadcasting, reproduction on micro ilms or in
any other physical way, and transmission or information storage and
retrieval, electronic adaptation, computer software, or by similar or
dissimilar methodology now known or hereafter developed.
The publisher, the authors, and the editors are safe to assume that the
advice and information in this book are believed to be true and
accurate at the date of publication. Neither the publisher nor the
authors or the editors give a warranty, expressed or implied, with
respect to the material contained herein or for any errors or omissions
that may have been made. The publisher remains neutral with regard
to jurisdictional claims in published maps and institutional af iliations.
1. Introduction
Jinghang Lu1
(1) Harbin Institute of Technology (Shenzhen), Shenzhen, China
Jinghang Lu
Email: jinghang.lu@ieee.org
The major bene its of this topology include simple design, low cost,
and small size of the converter. On the other hand, the lack of isolation
between the load and the grid voltage, no output voltage regulation,
long transfer time from the normal mode to the back-up mode, and the
poor performances are the main defects when the nonlinear load is
connected.
Jinghang Lu
Email: jinghang.lu@ieee.org
(2.1)
where V gα and V gβ indicate the grid voltage in the α − β reference
frame, V tα and V tβ are the output voltage of the converter, i α and i β are
the output current in the α − β reference frame, respectively.
By balancing the injected power and the active power of the DC-link,
the DC-link voltage controller keeps the DC-link voltage constant. The
DC-link capacitor power balance in Fig. 2.1 can be expressed as follows:
(2.2)
(2.3)
2.2 The Output Loop RGPIO-Based DC-Link
Voltage Controller Design
2.2.1 The Design of RGPIO
A reduced-order generalized proportional-integral observer (RGPIO) is
designed in this section to observe and actively cancel the lumped
disturbance caused by external disturbances and internal system
parameter variations.
The expression can be derived as follows by taking the Laplace
transfer function of both sides of (2.3):
(2.4)
where and P s are the system state and control input, respectively,
and P ext(s) is the system’s disturbance input.
By rearranging (2.2), the following expression is obtained:
(2.5)
(2.6)
where , , u = P s,
, and
To enhance the accuracy of estimation and also to enable simpler
practical implementation, a new RGPIO is designed for DC-link voltage
control. Figure 2.2 shows that the control structure includes the outer
loop and inner loop control strategies. It is usually assumed that the
dynamic response of the current loop is much faster than that of the
voltage loop when designing the DC-link voltage control loop so as to
prevent dynamic interaction between the DC-link voltage control and
the inner loop current control [1], and thus the dynamic of the current
loop can be neglected when designing the DC-link voltage control loop.
Fig. 2.2 Reduced-order GPIO based block diagram of the dual-loop control structure for the
three-phase AC/DC converter
(2.7)
(2.8)
Therefore, the RGPIO is designed as
(2.9)
(2.10)
(2.11)
(2.14)
(2.15)
(2.16)
(2.17)
(2.19)
Fig. 2.4 Pole’s movement when the DC-link capacitance value changes from 1100 to 3300 μF
(2.20)
(2.21)
(2.22)
(2.24)
where is the stage vector, represents the control input,
σ(t, x) is the sliding variable, and and are the smooth
uncertain functions.
In the ST-SMC, the control goal is to force σ and its derivative to
zero. Hence, by differentiating the sliding variable σ(t, x) twice, the
following expression can be derived as
(2.25)
where and g(t, x, u) are bounded but unknown, and there
exist positive constant values Φ, Γm, ΓM, such that the following
conditions are satis ied:
(2.26)
(2.27)
Then, a differential inclusion can be expressed as
(2.28)
With the condition of (2.26) and (2.27), a differential inclusion is
obtained:
(2.29)
Thus, a control law based on a super-twisting algorithm (STA) can
be designed as
(2.30)
where λ and μ are the design parameters that can be determined by the
boundary conditions (2.26) and (2.27). To ensure that the sliding
variable σ can be converged to the sliding manifold in inite time, the
parameters can also be chosen as follows:
(2.31)
(2.33)
A candidate Lyapunov function [4] is introduced to derive the
control law:
(2.34)
(2.35)
(2.36)
(2.37)
(2.40)
(2.42)
Parameter Value
Filter inductor L 1.8 mH
DC-link voltage 400 V
Inductor’s parasitic resistance 0.04 Ω
Grid RMS voltage 120 V
Sampling frequency f s 10 kHz
DC-link capacitance 1100 μF
Parameter A 35
Parameter B 10,000
Parameter C 500
Proportional gain (PR for comparison) 35
Resonant controller (PR for comparison) 1000
Parameter Value
Proportional gain (DC-link with PI) 0.2
Integral gain (DC-link with PI) 1
Proportional gain (RGPIO) 20
Fig. 2.11 Experimental results for the DC-link voltage reference increase. (a) DC-link voltage
with PI control, (b) DC-link voltage with the proposed control strategy, (c) active power with PI
control, and (d) active power with the proposed control strategy
Fig. 2.12 Experimental results for the transient response with reactive power step up. (a) d-axis
current with the standard control strategy, (b) d-axis current with the proposed control strategy,
(c) q-axis current with the standard control strategy, and (d) q-axis current with the proposed
control strategy
Fig. 2.13 Experimental results of the steady-state current with reactive power command. (a)
Current waveform for the PR control strategy, (b) current waveform for the proposed control
strategy, (c) spectrum of the current with PR control strategy, and (d) spectrum of the current
with the proposed control strategy
2.5.1 Dynamic Performance Under Load Step Change
Figure 2.10 shows the dynamic behavior of the standard PI control
strategy and the proposed control strategy with a sudden change in
external load from light load (1500 Ω) to full load (150 Ω). To avoid the
interaction of the voltage loop with the current loop and avoid stability
issues, the voltage loop of the DC-link control is typically 0.05–0.2
times the bandwidth of the current loop. Current loops are usually
chosen to be 0.1 times as wide as the sampling frequency. Thus, the
voltage loop is usually selected to be 0.005–0.02 times the sampling
frequency. It should be noted that a very fast DC-link voltage is not
needed, since the DC-link capacitor is generally large, and consequently,
the DC-bus voltage cannot be changed rapidly. The bandwidth is the
same for both DC-link voltage controllers, so a fair comparison can be
made, and both control laws can achieve DC-link voltage regulation.
However, their dynamic performances are quite different. Speci ically,
it is shown that with the PI control strategy, the DC-link voltage
undershoots by 60 V, and it takes around 1s to track the reference
voltage after the load is varied. With the proposed control strategy, the
RGPIO actively cancels the disturbance caused by the load variation,
resulting in a DC-link voltage dip of only 30 V and a recovery time of 0.4
s. Figure 2.10c and d shows the phase A output current with the load
variation. Compared to the PI control strategy, current with the
proposed control strategy has a faster speed. Furthermore, the PI
control strategy in the experiment works in the operating point, which
does not lead to the windup phenomenon in the system. In the event
that the system is overloaded by the proportional controller, the
proposed control strategy will not lead to a windup phenomenon. On
the contrary, the PI control strategy needs to add an anti-windup
algorithm to prevent the windup [5, 6].
2.6 Conclusion
A three-phase AC/DC converter was proposed using an RGPIO-based
RST-SMC strategy in this chapter. Compared with the conventional
control strategy, the system’s dynamic response under the disturbance
was greatly improved, and the settling time was reduced by employing
the proposed RGPIO-based control strategy. Under the uncertainty of
system parameters, RST-SMC was proposed as a current control
strategy for tracking in an α − β frame. The experimental results
demonstrated the effectiveness of the proposed control strategy
compared to the standard control strategy.
References
1. A. Yazdani, R. Iravani, An accurate model for the DC-side voltage control of the neutral point
diode clamped converter. IEEE Trans. Power Delivery 21(1), 185–193 (2006)
[Crossref]
2. J.D.P.G.F. Franklin, M. Workman, Digital Control of Dynamic System (Addison Wesley Longman,
Boston, 1998)
3. R.I. Amirnaser Yazdani, Voltage-Sourced Converters in Power Systems : Modeling, Control, and
Applications (Wiley, Hoboken, 2010)
[Crossref]
4. C.E.Y. Shtessel, L. Fridman, A. Levant, Sliding Mode Control and Observation (Springer, New
York, 2014)
[Crossref]
6. F.D. Bosio, L.A.D.S. Ribeiro, F.D. Freijedo, M. Pastorelli, J.M. Guerrero, Discrete-time domain
modeling of voltage source inverters in standalone applications: enhancement of regulators
performance by means of smith predictor. IEEE Trans. Power Electron. 32(10), 8100–8114
(2017)
[Crossref]
Jinghang Lu
Email: jinghang.lu@ieee.org
(3.1)
(3.2)
(3.3)
Fig. 3.2 Block diagram of dual-loop control diagram for the three-phase AC/DC converter
Figure 3.2 shows a DC-link controller with an ESO for
detecting/compensating system disturbances/uncertainties and a
simple proportional gain for regulating its voltage [2, 3]. The irst step
is to present the ESO analysis and design.
By applying some simple mathematical manipulations, (3.2) can be
rewritten as
(3.4)
(3.5)
(3.6)
is the gain vector of ESO which is designed using the pole placement
method in the next section.
(3.7)
(3.11)
(3.12)
(3.13)
(3.14)
where is the output of the controller. It is
observed from (3.14) that within the ESO’s bandwidth ω 0, the system
transfer function is reduced to an ideal integrator and expressed as
(3.15)
Parameter Value
Sampling frequency f s 10 kHz
Fig. 3.5 Root loci of the system with the capacitance change from 0.011 F to 0.033 F
(3.17)
3.2.1 Test 1
In this test, the DC-link capacitance is 0.011 F; the disturbance is
generated by connecting a 230 Ω resistor in the DC-link to examine the
controller’s performance from no load to full load operating point. The
DC-link voltage performance under the traditional PI control strategy,
feedforward-based PI control strategy, and the proposed control
strategy is compared in Fig. 3.7. The traditional PI control technique
has the worst performance, and the proposed sensor-less controller
and the sensor-based PI feedforward controller have similar
performance. To be more exact from Fig. 3.8, the traditional PI control
method has a peak undershoot of 60 V and a settling time of 0.8 s, while
the proposed controller has an undershoot of 30 V and a response time
of 0.3 s. The overshoot in the feedforward-based PI control approach is
the same as the proposed controller. However, the dynamic response is
a little faster than the proposed control strategy.
Fig. 3.7 Comparative study of the PI control strategy, PI with feedforward control strategy, and
proposed ESO-based control strategy when the DC-link capacitor is 0.011 F
Fig. 3.8 Comparison of DC-link voltage performance in test 1 with the traditional PI controller, PI
with feedforward controller, and proposed control strategy
3.2.2 Test 2
Compared to test 1, the DC-link capacitance has increased to 0.022 F to
examine the controller’s performance under 100% increase of DC-link
capacitance. The control parameters are identical to test 1. The
experimental results are shown in Fig. 3.9 for the traditional PI control
strategy, the feedforward-based PI control strategy, and the proposed
control strategy [5]. Both test 1 and test 2 yield the same conclusions.
In terms of setting time and disturbance rejection, the traditional PI
controller performs poorly. A forward-based PI controller and the
proposed control strategy have similar performance. Speci ically, in
Fig. 3.10, all three of these control strategies are better at rejection
disturbances. While the traditional PI method has an undershoot of 50
V and the settling time is still 0.8 s, the proposed method has an
undershoot of 20V and 0.3 s of settling time. As with the proposed
controller, the feedforward-based PI controller also undershoots, while
the settling time is 0.4 s.
Fig. 3.9 Comparative study of the PI control strategy, PI with feedforward control strategy, and
proposed ESO-based control strategy when the DC-link capacitor is 0.022 F
Fig. 3.10 Comparison of DC-link voltage performance in test 2 with the traditional PI controller,
PI with feedforward controller, and proposed ESO-based controller
3.3 Conclusion
Using an ESO-based DC-link voltage control strategy, a simple three-
phase AC/DC converter control strategy is presented. As a result of the
proposed control strategy, additional sensors were not required for
optimizing dynamic response to disturbances, making it possible to
expand UPS modules and microgrids via plug and play. DC-link voltage
controllers based on ESOs can mitigate the DC-link voltage variation as
sensor-based feedforward controllers, but they have a greater effect on
reducing the settling time. In the experiment, the robustness of the
controller was also examined. Experimental results con irmed the
effectiveness of the proposed control strategy.
References
1. R.I.A. Yazdani, Voltage-Sourced Converters in Power Systems: Modeling, Control, and
Applications (Wiley, Hoboken, 2010)
[Crossref]
3. Z. Song, C. Xia, T. Liu, Predictive current control of three-phase grid-connected converters with
constant switching frequency for wind energy systems. IEEE Trans. Ind. Electron. 60(6),
2451–2464 (2013)
[Crossref]
4. J.D.P.G.F. Franklin, M. Workman, Digital Control of Dynamic System (Addison Wesley Longman
Inc., Boston, 1998)
Baoze Wei
Email: bao@energy.aau.dk
(4.1)
(4.2)
where E and V pcc are the amplitude of the inverter output voltage and
the AC bus voltage, φ is the power angle of the inverter, and Z and θ
represent the amplitude and the phase of the output impedance. A
large ilter inductor connected to the AC bus typically produces a highly
inductive output impedance, which is traditionally assumed to be high.
In this case, assuming that the output impedance is purely inductive (θ
= 90∘ ), the active and reactive power expressions can be simpli ied as
(4.3)
(4.4)
(4.6)
in which ω ∗ and E ∗ represent the frequency and voltage amplitude
references, and m p and m q are the droop coef icients [5, 13, 14].
Similarly, for a highly resistive output impedance (θ = 0∘), the active
and reactive power can be calculated as
(4.7)
(4.8)
(4.10)
The active power can be controlled by the output-voltage amplitude
of the inverter, while the reactive power can be controlled by the
frequency of the inverter—this is the opposite strategy of conventional
droop, so it is called reverse droop. [5, 16–18], and [19] discuss in
greater detail the choice of droop function and the analysis of output
impedance.
A highly inductive output impedance is required to decouple the
in luence of P and Q from frequency and voltage amplitude in the
conventional droop control scheme, since the output impedance
in luences the choice of the droop function [5, 13]. Commonly, the LCL
ilter will be interconnected with the inverter. The output impedance
can be adjusted by adding virtual impedance. Due to the similarity in
operation between three-phase and single-phase inverters, Fig. 4.2 can
be simpli ied to represent parallel three-phase inverters considering
the output impedances.
(4.11)
(4.12)
(4.13)
(4.14)
Based on the former analysis, if the output voltages and the output
impedances of the parallel inverters are equal to each other,
respectively, the circulating current can be eliminated to obtain the
target of average power sharing. The equivalent circuit of two parallel-
connected inverters with virtual impedances is shown in Fig. 4.3, and Z
vir1 and Z vir2 are the virtual impedances.
(4.15)
Fig. 4.3 Equivalent circuit of two parallel inverters with virtual impedances
(4.16)
(4.17)
(4.18)
(4.19)
(4.20)
(4.21)
(4.25)
(4.26)
Thus, with (4.25) and (4.26), we can calculate the approximate control
parameters for the proposed DAVIC. It should be noticed that
optimized parameters will be obtained further based on the simulation
and experiments. The control parameters are shown in Table 4.1, and it
can be seen that the optimized parameters for the DAVIC are close to
the calculated values.
Table 4.1 Electrical and control parameters in simulations and experiments
Coef icient Parameter Value
Voltage reference (RMS) V ref 230 V/50 Hz
Filter inductor L 200 μH
Filter capacitor C 60 μF
Equivalent resistor of ilter inductor at 50 Hz rL 0.0628 Ω
Voltage proportional controller K PV 0.8 A/V
Voltage resonant controller K RV 1000 As/V
(4.28)
(4.30)
being
(4.31)
(4.32)
When considering the virtual impedance in the control loops in Fig.
4.7, where R D is the total virtual impedance that consists of the preset
virtual impedance and the adaptive virtual impedance, the closed-loop
transfer function will be
(4.33)
where
(4.34)
being
(4.35)
Fig. 4.8 Bode diagram of the output impedance, R D= 0 ∼ 1 Ω with increment of 0.1 Ω
Fig. 4.9 The zoomed-in Bode diagram of the output impedance around 50 Hz
4.3 Simulations
PLECS was used to build a UPS consisting of two inverter modules. The
model was used to test the effectiveness of the proposed adaptive
virtual impedance control. To analyze system dynamics when
considering unbalanced output impedances, different preset virtual
impedances were used in the simulation.
Fig. 4.14 Active and reactive power on Phase A of the two modules
Fig. 4.15 Total value of virtual impedance and the value of the adaptive virtual impedance for the
two modules
Voltage reference
Module 1 230∗1.01 V(RMS)
Module 2 230 V(RMS)
Fig. 4.23 Active and reactive power on Phase A of the two modules during the dynamic test
Fig. 4.24 The adaptive virtual impedance and the total virtual impedance of the two modules in
one phasek
Fig. 4.25 Voltage and current when there is communication delay between the two modules
Fig. 4.26 Active and reactive power of Phase A when there is communication delay between the
two modules
Fig. 4.27 Data communication delay happens between the two modules
4.4 Experimental Results
Table 4.3 shows the preset virtual impedances for the experiments
using a modular UPS platform. To verify the proposed control, two
unbalanced presets are used. Figure 4.28 illustrates the modular UPS
platform. In the control boards, the digital controller is a Del ino 32-bit
DSP from TI, TMS320F28377D.
4.5 Conclusion
Industrial applications with high levels of reliability are increasingly
using modular UPS systems, where the average power sharing
performance is an important factor. This section proposes an adaptive
virtual impedance control method to control modular UPS systems,
along with the procedure for designing control parameters for the
adaptive virtual impedance loop. A description of the selection criteria
for the virtual impedance is also included. PLECS was used to simulate
the performance and reliability of the proposed control using two
inverter modules. Experiments based on an industrial modular UPS
platform were performed. Comparatively to the conventional virtual
impedance control, the dynamic test can effectively suppress the
circulating current between parallel modules with different output
impedances, presenting a better average power sharing performance.
References
1. J.M. Guerrero, J.C. Vasquez, J. Matas, Control strategy for lexible microgrid based on parallel
line-interactive UPS systems. IEEE Trans. Ind. Electron. 56(3), 726–736 (2009)
[Crossref]
2. S.Y. Altahir, X. Yan, X. Liu, A power sharing method for inverters in microgrid based on the
virtual power and virtual impedance control, in Proceedings of the IEEE International
Conference on Compatibility, Power Electronics and Power Engineering, Cadiz (2017), pp.
151–156
3. J.M. Guerrero, L.G. Vicuñ a, J. Matas, M. Castilla, J. Miret, Output impedance design of parallel-
connected UPS inverters with wireless load-sharing control. IEEE Trans. Ind. Electron. 52(4),
1126–1135 (2005)
[Crossref]
4. C. Zhang, J.M. Guerrero, J.C. Vasquez, C.M. Seniger, Modular Plug’n’Play control architectures
for three-phase inverters in UPS applications. IEEE Trans. Ind. Appl. 52(3), 2405–2414
(2016)
[Crossref]
5. K. De Brabandere, B. Bolsens, J. Van den Keybus, A. Woyte, J. Driesen, R. Belman, A voltage and
frequency droop control method for parallel inverters. IEEE Trans. Power Electron. 22(4),
1107–15 (2007)
[Crossref]
6. K. De Brabandere, B. Bolsens, J. Van den Keybus, A. Woyte, J. Driesen, R. Belmans, K.U. Leuven,
A voltage and frequency droop control method for parallel inverters, in Proceedings of the
IEEE 35th Annual Power Electronics Specialists Conference, 20–25 June, 2004 (IEEE,
Piscataway, 2004), pp. 2501–2507
7. J.M. Guerrero, J. Matas, L. Vicuna, M. Castilla, J. Miret, Decentralized control for parallel
operation of distributed generation inverters using resistive output impedance. IEEE Trans.
Ind. Electron. 54(2), 994–1004 (2007)
[Crossref]
8. J. He, Y.W. Li, J.M. Guerrero, An islanding microgrid power sharing approach using enhanced
virtual impedance control scheme. IEEE Trans. Power Electron. 28(11), 5272–5282 (2013)
[Crossref]
11. A.R. Bergen, Power Systems Analysis (Prentice-Hall, Englewood Cliffs, 1986)
12. H. Mahmood, D. Michaelson, J. Jiang, Accurate reactive power sharing in an Islanded microgrid
using adaptive virtual impedances. IEEE Trans. Power Electron. 30(3), 1605–1617 (2015)
[Crossref]
13. C. Dou, Z. Zhang, D. Yue, M. Song, Improved droop control based on virtual impedance and
virtual power source in low-voltage microgrid. IET Gener. Transm. Distrib. 11(4), 1046–
1054, 3 9 (2017)
14. X. Sun, Y. Tian, Z. Chen, Adaptive decoupled power control method for inverter connected DG.
IET. Renew. Power Gener. 8(2), 171–82 (2014)
[Crossref]
15. A. Villa, F. Belloni, R. Chiumeo, C. Gandol i, Conventional and reverse droop control in islanded
microgrid: Simulation and experimental test, in Proceedings of the International Symposium
on Power Electronics, Electrical Drives, Automation and Motion (SPEEDAM), Anacapri
(2016), pp. 288–294
16. D. Wu, F. Tang, J.C. Vasquez, J.M. Guerrero, Control and analysis of droop and reverse droop
controllers for distributed generations, in Proceedings of the IEEE 11th International Multi-
Conference on Systems, Signals & Devices (SSD14), Barcelona (2014), pp. 1–5
17. J. Matas, M. Castilla, L. Vicuñ a, J. Miret, J.C. Vasquez, Virtual impedance loop for droop-
controlled single-phase parallel inverters using a second-order general-integrator scheme.
IEEE Trans. Power Electron. 25(12), 2993–3001 (2010)
[Crossref]
18. A. Engler, N. Soultanis, Droop control in LV-Grids, in Proceedings of the IEEE International
Conference on Future Power Systems, Amsterdam (2005), pp. 1–6
19. K. Heuk, K.-D. Dettmann, Elektrische Energieversorgung, 3rd edn. (Vieweg, German, 2005)
[Crossref]
20. Y. Zhang, M. Yu, F. Liu, Y. Kang, Instantaneous current-sharing control strategy for parallel
operation of UPS modules using virtual impedance. IEEE Trans. Power Electron. 28(1), 432–
440 (2013)
[Crossref]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
J. Lu et al. (eds.), Advanced Control and Protection of Modular Uninterruptible Power Supply
Systems, Power Systems
https://doi.org/10.1007/978-3-031-22178-1_5
Baoze Wei
Email: bao@energy.aau.dk
where E ref and f ref are the references of voltage and frequency,
respectively. E i and f i are the voltage amplitude and frequency of the
ith module, respectively, K PEsec and K IEsec are the proportional and
integral coef icients of the voltage secondary controller, respectively,
and K Pfsec and K Ifsec are the coef icients of the frequency secondary
controller.
Fig. 5.1 The conventional distributed secondary control scheme
Fig. 5.2 Control diagram of n parallel modules using droop, virtual impedance, and secondary
control
Fig. 5.3 Data communication setup for the distributed secondary control
Fig. 5.6 The three-phase output value of the secondary control of the two parallel modules
By Fig. 5.6, the output voltage from the second control module was
different from 0.15 s, and it cannot return to the same point at 0.25 s,
since the output voltage of the module #2 was already equal to the
reference. As a result, the secondary control already recovered the
voltage and is not responsible for the power sharing. Under dynamic
conditions, the conventional secondary control cannot guarantee the
convergence of the output voltages of all the parallel-connected
modules to the same reference voltage, resulting in poor power sharing
performance. In this section, a new secondary control distributed
across the parallel power modules is presented, which not only allows
voltage and frequency to be recovered but also demonstrates excellent
performance in sharing power.
(5.3)
(5.4)
(5.7)
(5.10)
Fig. 5.10 Eigenvalue traces of the system (5.7) (b) for K IEsec is various from 1 to 20
Table 5.1 Main control and hardware parameters in simulations and experiments
Coef icient Parameter Value
Nominal output voltage V orms 230 V
Fundamental frequency fo 50 Hz
Filter capacitance Cf 60 μF
Filter inductance Lf 200 μH
Virtual impedance R vir1 = R vir2 0.5 Ω
Voltage proportional controller K PV 0.8 A/V
(5.16)
Fig. 5.12 Eigenvalues trace of the system (5.7). (b) K Ifsec from 1 to 10
5.2.3 Circulating Current Analysis Regarding the Power
Sharing Performance
The equivalent circuit of two parallel-connected inverters with virtual
impedances is shown in Fig. 5.4, in which Z vir1 and Z vir2 are the virtual
impedances, Z line1 and Z line2 are the line impedances, and E 1 and E 2
are the capacitor voltages to be controlled. I 1 and I 2 are the output
currents of the parallel-connected power modules, Z L is the total load
impedance, and I o is the total load current. Then, the circulating
current can be de ined as follows:
(5.17)
As shown in Fig. 5.13, the following equations can be written as
(5.18)
(5.19)
Fig. 5.16 Local and average values of the integral part of the secondary controller
The second dynamic test was performed by changing the load-step
from 0.5 p.u. to 1 p.u. and vice versa. The irst step is at 0.15 s, the load
power changed from 0.5 p.u. to 1 p.u., and the second step is at 0.4 s,
back to 0.5 p.u. The results are shown in Figs. 5.17, 5.18, 5.19, and 5.20.
During the transient time, a good dynamic performance is obtained
along with an excellent average power sharing performance. In Fig.
5.20, the output voltage frequency did not change much during the
transient period.
Fig. 5.19 Local and average values of the integral part of the secondary controller
Fig. 5.20 Output frequency of the two modules
Parameters Value
Bit rate 500 kbps
Frame length 44 (bits data handling) + 64 (bits data) = 108 bits
Frame rate 500 kbps/108 bits = 4629.629
Frame time 1/4629.629 = 216 μs
1 data sending time 216 μs
Modules in the modular UPS platform have their own physical
identi iers. The module with the lowest physical ID is assigned priority
to send data to the CAN bus, and once the module receives the data, it
is recorded into the register. The average value of power will be
obtained when all the modules inish sending data.
Fig. 5.28 Experimental results during the transient time with step-load: CH1: Phase A current of
module #1, CH2: Phase A voltage of module #1, CH3: Phase A current of module #2, and CH4:
Phase A voltage of module #2
Fig. 5.29 Experimental results under steady state with step-load: CH1: Phase A current of module
#1, CH2: Phase A voltage of module #1, CH3: Phase A current of module #2, and CH4: Phase A
voltage of module #2
Fig. 5.30 Dynamic test when the output terminals of the modules were directly connected: CH1:
Phase A current of module #1, CH2: Phase A voltage of module #1, CH3: Phase A current of
module #2, and CH4: Phase A voltage of module #2
Fig. 5.31 Current waveforms when four modules in parallel operation: CH1: Phase A current of
module #1, CH2: Phase A current of module #2, CH3: Phase A current of module #3, and CH4:
Phase A current of module #4
Fig. 5.32 Voltage waveforms when four modules in parallel operation: CH1: Phase A voltage of
module #1, CH2: Phase A voltage of module #2, CH3: Phase A voltage of module #3, and CH4:
Phase A voltage of module #4
5.5 Conclusion
We present for modular UPS systems a novel distributed secondary
control (DAISC). It improves the dynamic performance of parallel
operation of converter modules compared to existing distributed
secondary controls. As a result, a modular UPS system will attain
excellent power sharing capabilities, which is very important for a UPS
system. At the same time, the system ensures nominal output voltage
and frequency. In addition, a stability analysis is presented to verify the
availability of the DAISC. With PLECS simulations and experimental
results on a real modular UPS under dynamic load conditions with
resistive, capacitive, and inductive loads, the performance of the
improved distributed secondary control was demonstrated.
References
1. A. Villa, F. Belloni, R. Chiumeo, C. Gandol i, Conventional and reverse droop control in islanded
microgrid: simulation and experimental test, in International Symposium on Power Electronics,
Electrical Drives, Automation and Motion (SPEEDAM), Anacapri (2016), pp. 288–294
2. D. Wu, F. Tang, J.C. Vasquez, J.M. Guerrero, Control and analysis of droop and reverse droop
controllers for distributed generations, in IEEE 11th International Multi-Conference on
Systems, Signals & Devices (SSD14), Barcelona (2014), pp. 1–5
3. J.C. Vasquez, J.M. Guerrero, M. Savaghebi, J. Eloy-Garcia, R. Teodorescu, Modeling, analysis, and
design of stationary reference frame droop controlled parallel three- phase voltage source
inverters. IEEE Trans. Ind. Electron. 60(4), 1271–1280 (2013)
[Crossref]
4. J. Matas, M. Castilla, L. Vicuñ a, J. Miret, J.C. Vasquez, Virtual impedance loop for droop-
controlled single-phase parallel inverters using a second-order general-integrator scheme.
IEEE Trans. Power Electron. 25(12), 2993–3001 (2010)
[Crossref]
5. J.W. Simpson-Porco, Q. Sha iee, F. Dö r ler, J.C. Vasquez, J.M. Guerrero, F. Bullo, Secondary
frequency and voltage control of islanded microgrids via distributed averaging. IEEE Trans.
Ind. Electron. 62(11), 7025–7038 (2015)
[Crossref]
6. C.K. Sao, P.W. Lehn, Autonomous load sharing of voltage source converters. IEEE Trans. Power
Delivery 20(2), 1009–1016 (2005)
[Crossref]
7. J.M. Guerrero, J. Matas, L.G. Vicuñ a, M. Castilla, J. Miret, Decentralized control for parallel
operation of distributed generation inverters using resistive output impedance. IEEE Trans.
Ind. Electron. 54(2), 994–1004 (2007)
[Crossref]
6. Regeneration Protection in
Uninterruptible Power Supply
Jinghang Lu1
(1) Harbin Institute of Technology (Shenzhen), Shenzhen, China
Jinghang Lu
Email: jinghang.lu@ieee.org
(6.3)
(6.4)
where P and Q are, respectively, the active and reactive power injected
into the PCC, V droop is the voltage amplitude of the inverter, V pcc is the
voltage amplitude at PCC, and δ is the phase angle difference between
the V droop and V pcc.
Accordingly, from (6.3) and (6.4), it can be seen that the active
power can be controlled by regulating output voltage amplitude V droop,
and the reactive power can be controlled by regulating the phase angle
δ. However, the initial phase angle of the inverter is dif icult to obtain.
Hence, the angular frequency ω, instead of the phase angle, is regulated
to control the reactive power. So, the droop control strategy is given by
(6.5)
(6.6)
where ω ∗ and ω are the UPS nominal and reference angular frequency
and E ∗ and E are the UPS nominal and reference voltage amplitude. P
LPF and Q LPF are the output active and reactive power through a low-
pass ilter with cutoff frequency ω c, and D p and D Q are the droop
coef icients for regulating the UPS active power and reactive power,
respectively. In this chapter, the design of the droop coef icient is based
on the static deviation method, which ensures that the system is
stable. D Q and D P are de ined as and
, Δω and ΔV are the maximum frequency and voltage
deviations, and and are the maximum active and reactive
power delivered by the inverter, respectively [3].
For the power low through the feeder that consists of inductance
and resistance, the voltage drop on the impedance leads to the
expression [4]
(6.7)
where ΔV is the voltage drop on the impedance, P and Q are the active
and reactive power respectively, and R and X are the resistance and
inductance of the line feeder, respectively. In the UPS system, by
neglecting the inductance, the voltage drop on the resistance is
expressed as
(6.8)
From (6.8), the voltage drop on the resistance of UPS system is derived
as
(6.9)
(6.10)
(6.11)
(6.12)
(6.13)
(6.14)
where k pv and k pi are the proportional terms and k rV and k ri are the
resonant term coef icient at ω o = 314 rad/s. The inner current loop is
designed to provide suf icient damping and protect the inductor’s
current from overcurrent.
The regeneration protection solution is based on DCVP and active
power sharing control. In other words, R v is the ixed virtual resistance
to make sure that the system is stable even when the adjustable virtual
resistance is subtracted.
(6.15)
(6.16)
where P and Q are the instantaneous active and reactive power of the
UPS system that lows out of the general line impedance. E and V are
the amplitudes of the inverter output voltage and the common bus
voltage, respectively, and φ is the power angle. Z and θ are the
magnitude and phase of the output impedance.
Due to the fact that the virtual resistance in the UPS system is
added in the control system and that the distance from the load to the
UPS system is usually quite short, the line resistance can usually be
omitted. The UPS system’s power low is therefore expressed as
follows:
(6.17)
(6.18)
Hence, the active power and reactive power under the small-signal
disturbance of voltage amplitude and the phase angle is expressed as
(6.19)
(6.20)
(6.22)
(6.23)
(6.24)
where τ is the time constant of the low-pass ilter in the active and
reactive power calculation.
Considering that and by the simple manipulation of
(6.19)–(6.20), the dynamic performance of the UPS system yields the
following expression:
(6.25)
where, , ,
Figure 6.6 shows the root locus of the UPS system with different
values of the virtual resistance. Three eigenvalues are not zero and one
is zero in this system. In the meantime, only nonzero eigenvalues are
used in the study of dynamic response and stability [8]. It can be seen
from Fig. 6.6 that when virtual resistance increases, the dominant
eigenvalue of λ 3 determines the performance of the system. However,
even if the virtual resistance varies widely, system stability remains. It
stays far away from the imaginary axis as this eigenvalue is far from it.
Fig. 6.6 Root locus when the virtual resistance changes from 0.3 to 0.8 Ω
Parameter Value
ESR of inductor 0.02 ohm
Fundamental frequency f o 50 Hz
Filter capacitance C f 27 μF
Proportional K pi 5.6
Resonant gain K ri 500
Fig. 6.10 DC-link voltage and active power during the activation transient time
Fig. 6.13 Parallel UPS power sharing performance during the current sharing process
Finally, in the steady state, the active and reactive power and output
current of parallel UPS system are shown in Fig. 6.14, where it is
observed that the current is equally shared and there is no circulating
current due to the active power difference.
Fig. 6.14 UPS current and their errors at steady state after activation of current sharing control
strategy. (Time Scale: 0.02s/div and Current Scale: 4A/div)
6.4 Conclusion
Using DDB’s online UPS system, this section proposes a regeneration
protection method. First, a power regeneration protection technique
based on virtual resistance is proposed to prevent parallel UPSs from
having excessive DC-link voltages. To mitigate the circulating current
caused by the power back-feeding, a virtual resistance-based power
sharing control strategy is also proposed. We have described the
principle of the proposed control scheme and analyzed the dynamic
response and stability of the proposed control strategy with small-
signal modeling. The experimental results support the effectiveness of
the proposed control strategy.
References
1. H.P. Glauser, M. Keller, A. Pluss, M. Schwab, R. Scherwey, New inverter module with digital
control for parallel operation, in TELESCON 2000. Third International Telecommunications
Energy Special Conference (IEEE Cat. No.00EX424) (2000), pp. 265–269
2.
J.M. Guerrero, L. Hang, J. Uceda, Control of distributed uninterruptible power supply systems.
IEEE Trans. Ind. Electron. 55(8), 2845–2859 (2008)
[Crossref]
4. J.D. Glover, M.S. Sarma, T. Overbye, Power System Analysis and Design, 5th edn. (Cengage
Learning, 2011)
5. Q. Sha iee, J.M. Guerrero, J.C. Vasquez, Distributed secondary control for islanded microgrids—
a novel approach. IEEE Trans. Power Electron. 29(2), 1018–1031 (2014)
[Crossref]
6. D. He, D. Shi, R. Sharma, Consensus-based distributed cooperative control for microgrid voltage
regulation and reactive power sharing, in IEEE PES Innovative Smart Grid Technologies,
Europe (2014), pp. 1–6
7. H. Jinwei, L. Yun Wei, F. Blaabjerg, An enhanced islanding microgrid reactive power, imbalance
power, and harmonic power sharing scheme. IEEE Trans. Power Electron. 30(6), 3389–3401
(2015)
[Crossref]
8. H. Jinwei, L. Yun Wei, Analysis, design, and implementation of virtual impedance for power
electronics interfaced distributed generation. IEEE Trans. Ind. Appl. 47(6), 2525–2538 (2011)
[Crossref]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
J. Lu et al. (eds.), Advanced Control and Protection of Modular Uninterruptible Power Supply
Systems, Power Systems
https://doi.org/10.1007/978-3-031-22178-1_7
Jinghang Lu
Email: jinghang.lu@ieee.org
As speci ied in the IEC 62020-3:2011 standard [1], when the UPS
system is operating normally, the battery is fully charged and in
standby mode. From the grid to the DC link, the SCR delivers active
power in a unidirectional direction (Fig. 7.1 brown arrows). At the
same time, the inverter is capable of operating two-dimensionally and
bidirectionally (see Fig. 7.1 blue arrows). In Fig. 7.1, the UPS module
with the higher output voltage may be able to feed effective power to
the other UPS module under light load, depending on whether the
inverter has a low output voltage or an output fault (red arrow in Fig.
7.1). If the feeding power leads to an increase in the DC-link voltage (in
this case V DC1) and the DC-link voltage exceeds its upper limit without
the DCVP, then system protection is triggered. Additionally, when UPS
operates in the eco-mode, the bypass switch must be closed. If the grid
suffers a temporary overvoltage or amplitude synchronization is not
performed, the grid voltage may be higher than the voltage at the PCC.
Figure 7.1 green arrows: in this situation, both UPS modules might be
powered by grid power, leading to excessive DC-link voltages on both
UPS modules. Accordingly, excessive DC-link voltages cause inverter
operation to be interrupted. Therefore, DCVP control algorithms
should be investigated.
(7.3)
where ω ∗ and ω are the UPS nominal and
reference angular frequency, and E ∗ and E are the UPS nominal
and reference voltage amplitude. P LPF and Q LPF are the output active
and reactive power through a low-pass ilter with cutoff frequency
ω c. D p and D Q are the droop coef icients for regulating the UPS
active power and reactive power, respectively. In this chapter, the
design of the droop coef icient is based on the static deviation method,
D Q and D P are de ined as and ,
Δω and ΔV are the maximum
frequency and voltage deviations, and and are the
maximum active and reactive power delivered by the inverter,
respectively.
As shown in Fig. 7.3, the parallel UPS modules are modeled as
controlled voltage sources (V ref) with virtual resistances R v,f at
fundamental frequency. The resistance of the line at the fundamental
frequency is indicated by R line,f. There is no way to omit the R Line,f even
though it is small. R L is modeled passively using the PCC load.
Normally, for the power low through the feeder that consists of
inductance and resistance, the voltage drop on the impedance leads to
the following expression:
(7.4)
where P and Q are the active and reactive instantaneous power that
lows out of the general line impedance, and R and X are the
corresponding resistive and inductive component of the line
impedance. E ∗ is the nominal voltage magnitude, and
ΔV is the voltage magnitude drop on the impedance.
In the UPS system, by neglecting the line inductance, the voltage
drop on the resistance is expressed as
(7.5)
(7.6)
(7.7)
(7.8)
It can be observed from (7.8) that the active power sharing error is
related to two factors that include the total resistance difference (R e1
and R e2) of the two UPS system and the voltage magnitude difference
(V ref1 and V ref2). If the circulating current is caused by the difference
between V ref1 and V ref2 or line resistance mismatching (R Line1 and R
Line2), the best way to mitigate the circulating current is to adjust each
UPS’s virtual resistance (R v,f1 and R v,f2).
Meanwhile, the equivalent circuit at the harmonic frequencies
when connecting a nonlinear load to PCC is shown in Fig. 7.4. The
harmonic resistance is expressed as
(7.9)
where R line,H and R v,H are the physical feeder and virtual resistance at
the harmonic frequencies, respectively. The controlled harmonic
voltage source is considered zero in the system at harmonic frequency,
since the voltage source is considered short circuit at harmonic
frequency. According to Fig. 7.4, the decrease of UPS equivalent
resistance results in the increase of the corresponding harmonic
current and vice versa. As a result, by properly controlling the virtual
harmonic resistance, the circulating harmonic current caused by line
resistance mismatch can be mitigated.
(7.10)
(7.12)
(7.13)
where E ∗ is the rated UPS phase voltage. and are the
negative sequence component of UPS ifth harmonic current. and
are the positive sequence component of UPS seventh harmonic
current. Harmonic power is calculated by using only low-order
harmonic currents, as 5th and 7th harmonic currents comprise the
main components of harmonic orders. The SOGI-based sequence
decomposition method described in Fig. 7.6 can also be used to
separate fundamental and harmonic current of UPS. There are two
SOGI quadrature signal generators (SOGI-QSGs), each of which consists
of a harmonic decoupling network, a frequency locked loop (FLL), and
multiple frequency-locked loops. In highly distorted current conditions,
the SOGI-based sequence extraction method shows fast and accurate
results for detecting fundamental frequency and harmonic frequency
currents.
Fig. 7.6 Decomposition of fundamental sequence and harmonic component of the current
Fig. 7.7 Implementation of virtual resistance at the fundamental sequence and harmonic
frequencies
(7.14)
(7.15)
(7.16)
where δ ij(k) stores the cumulative difference between the two
agents.
In this chapter, x i(k) may indicate the frequency and voltage
amplitude. Figure 7.8 shows that the saturation block is inserted
between x i(k) and in order to prevent frequency or voltage
amplitudes from exceeding the permitted value. Even though the
uncertainty block is only added to the consensus algorithm, it will still
take a long time to wind up. A back calculation method [10] is adopted
to calculate the x i(k + 1), which prevents windup
and slows down dynamic response. Hence, x i(k + 1)
is updated as
(7.17)
From Eq. (7.17), it is seen that in steady state x i(k) does not exceed
the saturation value, , and therefore, Eq. (7.17) equals
with Eq. (7.15). However, in dynamic restoration process, x i(k)
exceeds the saturation block value, and the anti-windup part will force
the x i(k) to be lower than upper limitation value of saturation block to
alleviate the frequency and voltage amplitude overshoot.
Figure 7.9 contains the complete control diagram, including the
primary and secondary control levels. In addition to the droop control
strategy and DCVP method, voltage and current loops are included in
the primary control level. A secondary control level implements the
fundamental and harmonic current sharing strategies, as well as the
ADCA strategy. The section utilizes a dual loop control strategy with
harmonic compensators in order to achieve voltage tracking. To
regulate the output capacitor’s voltage, the outer loop voltage
controller is implemented. The inverter side current is regulated by an
inner loop current controller that is nestled inside the voltage control
loop. The voltage controller and the current controller are expressed as
follows:
(7.18)
(7.19)
Parameter Value
ESR of inductor 0.02 ohm
Fundamental frequency f o 50 Hz
Filter capacitance C f 27 μF
Filter inductance L f 1.8 mH
Fig. 7.22 Comparison of voltage amplitude deviation with ADCA and DCA
7.4 Conclusion
As part of the modular UPS system, this section discussed DC-link
protection and control. We irst proposed a DCVP strategy for
protecting inverters from excessive DC-link voltage. Through DCVP
control, the power back-feeding problem can be signi icantly reduced.
To mitigate the impact of line resistance mismatches and power back-
feeding, a strategy for sharing fundamental and harmonic currents in
the UPS system is proposed. To prevent voltage and frequency
overshoots, a dynamic consensus algorithm is presented as an anti-
windup solution. Experiments have demonstrated the ef icacy of the
methods presented.
References
1. IEC_62040-3-2011 Uninterruptible power systems (UPS)—part 3: method of
specifying the performance and test requirements
2.
C. Zhang, J.M. Guerrero, J.C. Vasquez, E.A.A. Coelho, Control architecture for parallel-connected
inverters in uninterruptible power systems. IEEE Trans. Power Electron. 31(7),
5176–5188 (2016)
3. C. Li, E.A.A. Coelho, T. Dragicevic, J.M. Guerrero, J.C. Vasquez, Multiagent-based distributed
state of charge balancing control for distributed energy storage units in AC microgrids. IEEE
Trans. Ind. Appl. 53(3), 2369–2381 (2017)
[Crossref]
4. Y. Xu, Z. Li, Distributed optimal resource management based on the consensus algorithm in a
microgrid. IEEE Trans. Ind. Electron. 62(4), 2584–2592 (2015)
[MathSciNet][Crossref]
5. H.P. Glauser, M. Keller, A. Pluss, M. Schwab, R. Scherwey, New inverter module with digital
control for parallel operation, in TELESCON 2000. Third International Telecommunications
Energy Special Conference (IEEE Cat. No.00EX424) (2000), pp. 265–269
6. E.A.A. Coelho, P.C. Cortizo, P.F.D. Garcia, Small-signal stability for parallel-connected inverters
in stand-alone AC supply systems. IEEE Trans. Ind. Appl. 38(2), 533–542 (2002)
[Crossref]
7. E.A. Alves Coelho, P.C. Cortizo, P.F.D. Garcia, Small signal stability for single phase inverter
connected to stiff AC system, in Industry Applications Conference, 1999. Thirty-Fourth IAS
Annual Meeting. Conference Record of the 1999 IEEE, vol. 4 (1999), pp. 2180–2187
8. J. Schiffer, T. Seel, J. Raisch, T. Sezi, Voltage stability and reactive power sharing in inverter-
based microgrids with consensus-based distributed voltage control. IEEE Trans. Control Syst.
Technol. 24(1), 96–109 (2016)
[Crossref]
10. A.R. Teel, Windup in control: its effects and their prevention (by Hippe, P.; 2006). IEEE Trans.
Autom. Control 53(8), 1976–1977 (2008)
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
J. Lu et al. (eds.), Advanced Control and Protection of Modular Uninterruptible Power Supply
Systems, Power Systems
https://doi.org/10.1007/978-3-031-22178-1_8
Jinghang Lu
Email: jinghang.lu@ieee.org
Fig. 8.2 The control diagram considering overload and short-circuit protection
Figure 8.3 shows the simpli ied linear control model for the voltage
inverter in the UPS, in which G V(s) represents the transfer function of
the voltage controller; G I(s) represents the transfer function of the
current controller, and G PWM(s) represents the transfer function of the
inverter. De ining that:
(8.1)
where K PV and K RV are the proportional and resonate coef icients of
the voltage PR controller, respectively; K PI and K RI are the proportional
and resonate coef icients of the current PR controller, respectively; and
for the three-level voltage-source inverter, the gain K PWM equals to V
dc∕2. Moreover, in order to suppress the in luence of the luctuation of
the DC bus voltage, the DC bus voltage will always be considered in the
control, and the PWM signal will be divided by V dc∕2 before sending to
the driver circuit. So that the gain of the path from the output of the
current controller to the inverter output, from the point PWM* to the
point V XO(s)(X = A, B, C) in Fig. 8.2, will be one. Accordingly, a unit
feedforward control is applied in this chapter, F(s) = 1.
(8.2)
(8.3)
After the resonant part of the PR controller, see y KR(s) in Fig. 8.4 and
applying (8.3), it yields to
(8.4)
(8.5)
(8.6)
(8.7)
If the resonant controller is well tuned, the inal value of the
resonant controller will be a constant with a zero steady-state error if
applied in a control architecture. An error signal triggers the resonant
controller to operate. The resonant controller can also be observed as
only working when the input signal is non-zero. The result of this is
that over time, the output of the resonant controller will accumulate
with the error signal. If given enough time, a small error can turn into a
large one. As the error accumulates, the system is forced to correct it
until the steady-state error is zero. An integral or resonant controller
will always overshoot the controlled signal. Some tools can be applied
to evaluate a controller’s effectiveness in a control diagram.
Control schemes are often tested for effectiveness and stability
with step responses [43–48]. The power system must be capable of fast
dynamic response in order to reduce deviations in output voltage or
current during transient events [43]. Power electrical system unit step
response can be de ined as the time behavior of the system when the
input changes from 0 p.u. to 1 p.u. within a short amount of time [49].
As a practical matter, knowledge of how the system responds to a
sudden input is essential because large deviations from the long-term
steady state may have extreme effects on the controller, as well as on
other parts of the system that are dependent on it. In addition, the
overall system cannot act until the controller’s output reaches a point
near its inal state, so it takes longer for the overall system to respond.
It gives information on the stability of a dynamical system and on how
fast a system can reach a stationary state (if it starts from another
point of operation) if known the step response of the system [49].
Normally, we can obtain the following information from the system
step response: overshoot, settling time, and steady-state information.
In the event of an overload or a short circuit, the system moves to a
new operation point, from a load perspective, and it creates a large
load-step change, which requires more current to be supplied by the
converter. Besides reducing overshoot and settling times, we also need
to achieve an accurate current limit protection in order to protect the
hardware. By examining the step-response analysis of a closed-loop
system as well as the control strategy, the control can be improved in a
systematic way.
For the purpose of analyzing the step-response behavior of the
control architecture, three different con igurations of the control
architecture are considered:
(8.8)
(8.9)
(8.10)
in which
(8.11)
Table 8.1 Parameters of the controllers and hardware circuits in the experiments
Figures 8.9 and 8.10 depict the waveforms when the UPS switches
from its normal working condition to overload mode. The simulation
shows the same phenomenon as the reset operation of the resonant
regulator of the current controller in the DSP when overload occurs.
Besides, since the current limit is set to 20 A, it shows an accurate
current limit. In the transient phase, the waveform looks good and the
presented control scheme does a good job.
Fig. 8.9 Single phase from normal to overload condition. CH1: Output voltage; CH2: Resonant
output of current controller in DSP; CH3: Inductor current; CH4: Output current
Fig. 8.10 Overload happens with single-phase load. CH1: Output voltage; CH2: Resonant output
of current controller in DSP; CH3: Inductor current; CH4: Output current
Figures 8.11 and 8.12 give the experimental results when phase-to-
ground short-circuit fault happens. It shows the dynamic performance
of the cases from normal to short circuit and the reverse. After clearing
the fault, the output current could be limited quickly to the preset value
and returned to normal operation as expected.
Fig. 8.11 Single phase from normal to phase-to-ground short circuit. CH1: Output voltage; CH4:
Output current
Fig. 8.12 Phase-to-ground short-circuit condition switch to normal working condition on single
phase. CH1: Output voltage; CH4: Output current
Figure 8.13 shows the results when the UPS was operated under
normal working conditions with a three-phase load. Figures 8.14 and
8.15 illustrate the waveforms of symmetric overload occurring in three
phases. By using the proposed method, it is still possible to limit the
current quickly and accurately from no load to overload condition
directly. These results demonstrate the usefulness of the control
method presented here. As there are no more channels on one
oscilloscope, Phase C results were not shown in the experimental
results. Figures 8.16, 8.17, and 8.18 show the experimental results
when a phase-to-phase short circuit occurs. When a fault occurs on one
of the phases, it does not affect the normal operation of the other
healthy phase, and it can quickly return to normal operation when it is
ixed. Very similar results are obtained as shown in the simulation.
Furthermore, the consistency of the simulation and experimental
results further supports the effectiveness and applicability of the
proposed method.
Fig. 8.13 Three-phase load normal working condition. CH1: Phase A output voltage; CH2: Phase
A output current; CH3: Phase B output current; CH4: Phase C output current
Fig. 8.14 Overload happens with three-phase load. CH1: Phase A output voltage; CH2: Phase B
output voltage; CH3: Phase A output current; CH4: Phase B output current
Fig. 8.15 Three-phase overload switch to no load condition. CH1: Phase A output voltage; CH2:
Phase B output voltage; CH3: Phase A output current; CH4: Phase B output current
Fig. 8.16 Normal operation to phase-to-phase short circuit. CH1: Phase A output voltage; CH2:
Phase A output current; CH3: Phase B output current; CH4: Phase C output current
Fig. 8.17 Phase-to-phase short circuit to normal operation. CH1: Phase A output voltage; CH2:
Phase A output current; CH3: Phase B output current; CH4: Phase C output current
Fig. 8.18 Normal operation to phase-to-phase short circuit. CH1: Phase A output voltage; CH2:
Phase B output voltage; CH3: Phase A output current; CH4: Phase B output current
8.5 Conclusions
A control method is presented in this chapter to protect voltage-source
inverters-based UPS systems from overloads and short circuits. As a
result of the proposed control method, UPS output current can be
limited to a reference value within the safe operation area, so that a
voltage drop occurs when an overload or short circuit occurs. In
comparison with the state-of-the-art solutions, the proposed solution
has greater accuracy and validity for a greater range of overcurrent
issues, while being compatible with advanced current/voltage tracker
regulators based on resonant controllers. Moreover, the proposed
method is robust to not only symmetric faults in three-phase systems
but also asymmetric overloads and short circuits in single-phase
systems, and its healthy phase will not be disturbed by faults. Another
feature of the proposed controller is that it does not have to switch
between operation modes, which can result in instabilities and
oscillatory transients. This is in contrast to conventional methods in
which the control switches from voltage to current mode once
overloading or short-circuiting occurs and then back to voltage mode
once the fault has been cleared. As a result of this approach, once the
fault is resolved, the UPS system can return to normal operation
without changing the control mode. It would continue to act as a
voltage-controlled source. Testing of the proposed control method has
been conducted on a commercial UPS platform.
References
1. I. Lizama, R. Alvarez, S. Bernet, M. Wagner, A new method for fast short circuit protection of
IGBTs, in Proc. IEEE IECON 2014 (2014), pp. 1072–1076
2. J.H. Choi, K.J. Hong, J.S. Park, B.G. Gu, C.Y. Won, A new short circuit protection scheme for small
inverters, in Proc. IEEE VPPC 2012 (2012), pp. 1544–1547
3. X. Zhang, M. Chen, N. Zhu, D. Xu, A self-adaptive blanking circuit for IGBT short-circuit
protection based on VCE measurement, in Proc. IEEE ECCE 2015 (2015), pp. 4125–4131
4. T. Horiguchi, S. Kinouchi, H. Urushibata, S. Okamoto, S. Tominaga, H. Akagi, A short circuit
protection method based on a gate charge characteristic, in Proc. IEEE ECCE-Asia 2014
(2014), pp. 2290–2296
6. A.K. Irodi, N.R. Srinivas, Active overcurrent protection schemes in bridge inverters operating
under shoot-through conditions, in Proc. IEEE PICC 2015 (2015), pp. 1–5
8. Y. Wang, C. Lee, P. Kuo, Y. Lin, Overcurrent protection design, failure mode and effect analysis
of an electric vehicle inverter, in Proc. IEEE ICIT 2016 (2016), pp. 1287–1292.
9. X. Pei, Y. Kang, Short-circuit fault protection strategy for high-power three-phase three-wire
inverter. IEEE Trans. Ind. Inform. 8(3), 545–553 (2012)
[Crossref]
10. F.B. Costa, A. Monti, S.C. Paiva, Overcurrent protection in distribution systems with distributed
generation based on the real-time boundary wavelet transform. IEEE Trans. Power Delivery
32(1), 462–473 (2017)
[Crossref]
11. M.E. Santos, B.J.C. Filho, Short circuit and overcurrent protection of IGCT-based three-level
NPC inverters, in Proc. IEEE PESC 2004 (2004), pp. 2553–2558
12. N. Bottrell, T.C. Green, Comparison of current-limiting strategies during fault ride-through of
inverters to prevent latch-up and wind-up. IEEE Trans. Power Electron. 29(7), 3786–3797
(2014)
[Crossref]
13. B. Lu, S.K. Sharma, A literature review of IGBT fault diagnostic and protection methods for
power inverters. IEEE Trans. Ind. Appl. 45(5), 1770–1777 (2009)
[Crossref]
14. M. Oinonen, Short circuit protection for a power transistor. Master of Science Thesis, Aalto
University, Finland, 2013, p. 81
15. A. Volke, M. Hornkamp, IGBT Modules, Technologies, Driver and Application (In ineon
Technologies AG, 2011)
18. A.S. Noghabi, H.R. Mashhadi, J. Sadeh, Optimal coordination of directional overcurrent relays
considering different network topologies using interval linear programming. IEEE Trans.
Power Delivery 25(3), 1348–1354 (2010)
[Crossref]
19. R.M. Chabanloo, H.A. Abyaneh, A. Agheli, H. Rastegar, Overcurrent relays coordination
considering transient behaviour of fault current limiter and distributed generation in
distribution power network. IET Gen. Transm. Distrib. 5(9), 903–911 (2011)
[Crossref]
20. M.J. Newman, D.G. Holmes, An integrated approach for the protection of series injection
inverters. IEEE Trans. Ind. Appl. 38(3), 679–687 (2002)
[Crossref]
21. T. Horiguchi, S.Kinouchi, Y. Nakayama, H.Akagi, A fast short-circuit protection method using
gate charge characteristics of SiC MOSFETs, in Proc. IEEE ECCE 2015 (2015), pp. 4759–4764
22. R.S. Chokhawala, J. Catt, L. Kiraly, A discussion on IGBT short circuit behavior and fault
protection schemes. IEEE Trans. Ind. Appl. 31(2), 256–263 (1995)
[Crossref]
24. B.G. Park, J.B. Lee, D.S Hyun, A novel short-circuit detecting scheme using turn-on switching
characteristics of IGBT, in Conf. Rec. IEEE IAS Annual Meeting (2008)
25. S. Musumeci, R. Pagano, A. Raciti, G. Belverde, M. Melito, A new gate circuit performing fault
protections of IGBTs during short circuit transients, in Conf. Rec. IEEE IAS Annual Meeting,
vol. 3 (2002), pp. 2614–2621
26. K. Yuasa, S. Nakamichi, I. Omura, Ultra high speed short circuit protection for IGBT with gate
charge sensing, in Conf. Rec. IEEE ISPSD (2010), pp. 37–40
27. T. Tanimura, K. Yuasa, I. Omura, Full digital short circuit protection for advanced IGBTs, in
Conf. Rec. IEEE ISPSD (2011), pp. 60–63
29. K. Ishikawa, K. Suda, M. Sasaki, H. Miyazaki, A 600 V driver IC with new short circuit
protection in hybrid electric vehicle IGBT inverter system, in Conf. Rec. IEEE ISPSD (2005),
pp. 59–62
30.
B.G. Park, J.B. Lee, D.S. Hyun, A novel short-circuit detecting scheme using turn-on switching
characteristic of IGBT, in Industry Applications Society Annual Meeting (IEEE IAS, 2008), pp.
1–5
31. K. Yuasa, S. Nakamichi, I. Omura, Ultra high speed short circuit protection for IGBT with gate
charge sensing, in 22nd International Symposium on Power Semiconductor Devices IC’s
(ISPSD) (2010), pp. 37–40
32. A. Radun, An alternative low-cost current-sensing scheme for high-current power electronics
circuits. IEEE Trans. Ind. Electron. 42(1), 78–84 (1995)
[Crossref]
33. F. Huang, F. Flett, IGBT fault protection based on di/dt feedback control, in Power Electronics
Specialists Conference (IEEE PESC) (2007), pp. 1478–1484
34. Z. Wang, X. Shi, L.M. Tolbert, A fast overcurrent protection scheme for IGBT modules through
dynamic fault current evaluation, in IEEE APEC (2013), pp. 577–583
35. Z. Hu, C. Mao, J. Lu, S. Fan, Fuse protection of IGCTs against rupture in three-level commutated
inverter, in Proc. IEEE ICPSC 2002 (2002), pp. 611–615
36. D. Liu, A. Hu, G. Wang, J. Guo, Triple-loop-controlled overload or short circuit current limiter
and protection for three phase inverter, in Proc. IEEE ASEMD 2009 (2009), pp. 201–205
37. Z. Chen , X. Pei , M. Yang , L. Peng, P. Shi, A novel protection scheme for inverter-interfaced
microgrid (IIM) operated in islanded mode. IEEE Trans. Power Electron. 33(9), 7684–7697
(2018)
38. R. Guzman, L.G. Vicuna, J. Morales, M. Castilla, J. Miret, Model-based active damping control for
three-phase voltage source inverters with LCL ilter. IEEE Trans. Power Electron. IEEE Early
Access (2016). https://doi.org/10.1109/TPEL.2016.2605858
39. M.S. Moon, R.W. Johnson, DSP control of UPS inverter with over-current limit using droop
method, in Proc. IEEE 30th Annual IEEE Power Electronics Specialists Conference
(Charleston, 1999), pp. 552–557
40. Z. Liang , X. Lin , Y. Kang, B. Gao, H. Lei, Short circuit current characteristics analysis and
improved current limiting strategy for three-phase three-leg inverter under asymmetric short
circuit fault. IEEE Trans. Power Electron. 33(8), 7214–7228 (2018)
[Crossref]
41. H. Wang, X. Pei, Y. Chen, Y. Kang, Y.F. Liu, Short-circuit fault protection strategy of parallel
three-phase inverters, in Proc. IEEE Energy Convers. Congr. Expo. (Montreal, 2015), pp.
1736–1742
43. D. Segaran, D.G. Holmes, B.P. McGrath, Enhanced load step response for a bidirectional DC–DC
converter. IEEE Trans. Power Electron. 28(1), 371–379 (2013)
[Crossref]
44.
C.A. Reis, N.A.P. Silva, Suf icient conditions for analysis of extrema in the step-response of the
linear control systems. IEEE Latin Am. Trans. 9(7), 1025–1031 (2011)
[Crossref]
45. F. Zhang, Y. Yan, Start-up process and step response of a DC–DC converter loaded by constant
power loads. IEEE Trans. Ind. Electron 58(1), 298–304 (2011)
[Crossref]
46. J. Kim, J. Lee, S. Cho, C. Hwang, C. Yoon, J. Fan, Analytical probability density calculation for
step pulse response of a single-ended buffer with arbitrary power-supply voltage
luctuations. IEEE Trans. Circuits Syst. I: Regul. Pap. 61(7), 2022–2033 (2014)
[Crossref]
47. T.J. Freeborn, B. Maundy, A.S. Elwakil, Measurement of supercapacitor fractional-order model
parameters from voltage-excited step response. IEEE J. Emerg. Sel. Topics Circuits Syst. 3(3),
367–376 (2013)
[Crossref]
48. H. Oka, Y. Baba, M. Ishii, N. Nagaoka, A. Ametani, Parametric study on unit step responses of
impulse voltage measuring systems based on FDTD simulations. IEEE Trans. Power Delivery
28(1), 376–382 (2013)
[Crossref]
49. C. Benjamin Kuo, F. Golnaraghi, Automatic Control Systems, 8th edn. (Wiley, New York, 2003),
p. 253. ISBN 0-471-13476-7
Part III
Renewable Modular UPS System
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
J. Lu et al. (eds.), Advanced Control and Protection of Modular Uninterruptible Power Supply
Systems, Power Systems
https://doi.org/10.1007/978-3-031-22178-1_9
Xiaochao Hou
Email: houxiaochao@tsinghua.edu.cn
Figure 9.2 shows the schematic diagram for the on-line UPS system
and its associated control strategy. There are two cascaded control
loops in the control diagram: the outer DC-link voltage control loop
described in this section and the inner current control loop. PV,
batteries, and inverter units are represented by the general load. The
DC-link voltage is regulated by the AC/DC converter. Therefore, a
robust control strategy for the DC-link voltage is quite important for a
stable UPS system. The proposed two-port IMC-based DC-link
controller will be illustrated in the following sections.
Fig. 9.2 Schematic diagram of the AC/DC converter with the proposed two-port IMC-based DC-
link voltage control in PV-aided UPS system
(9.1)
(9.2)
(9.3)
In Fig. 9.4, the control diagram is shown when the 2-DOF IMC-based
strategy is applied to the DC-link voltage controller. This control
structure consists of an inner current loop for controlling the recti ier
output current and an outer voltage loop for regulating the DC-link
voltage. It is often considered that the bandwidth of the inner loop
should be much larger than that of the outer loop in order to prevent
dynamic interactions between them. It implies that, during the DC-link
controller design, the dynamics of the current regulation loop can be
neglected, and assume that ref, where and P S are
the reference and output active powers, respectively (see Fig. 9.4).
Fig. 9.4 The 2-DOF internal model controller structure
From Fig. 9.4, the closed-loop transfer function for the DC-link
voltage is expressed as
(9.4)
(9.6)
(9.8)
When there is the difference between the plant G P(s) and the
reference model G m(s), E( s) provides the information of the
disturbance and the plant mismatch. The robustness can be obtained
by compensating for the deviation appropriately.
In this chapter, the actual plant is expressed as: , and
the reference model is expressed as: , as the R p is considered
as the uncertainty in the system.
The parameter λ 1 of C 1(s) mainly decides the DC-link voltage
closed-loop bandwidth. In our system, the DC-link closed-loop
bandwidth is chosen to be 20 rad/s; hence, λ 1 = 1∕20. In addition, the
parameter λ 2 of Fr(s) is usually chosen to be much smaller than λ 1 to
make sure disturbance rejection controller’s bandwidth is large enough
to disturbance. But too large bandwidth may interact
with current controller’s bandwidth and, as a result, may cause stability
issue due to the coupling. Moreover, too large bandwidth may bring the
noise as well. Considering the above requirement, λ 2 = 1∕200[3–6].
(9.13)
(9.15)
If the internal mode is accurate, i.e., G P(s) = G m(s), from (9.15), the
closed-loop transfer function for the DC-link voltage is expressed as
(9.16)
(9.17)
(9.19)
Parameter Value
Sampling frequency f s 10 kHz
Filter inductor L f 1.8 mH
Parameter Value
Switching loss R p 1000 Ω
Fig. 9.7 Power low in the PNM of the on-line UPS system
(1) Control Strategy for DC/AC Inverters in on-line UPS System Figure
9.8 shows the control strategy of the UPS system in the PNM operation.
The droop control strategy is implemented for the parallel-connected
inverter modules. Due to the short distance from the inverter to the
load in the UPS system, the line impedance is considerably small. And a
virtual resistance R v is embedded in the control diagram to increase
the system stability. As the output impedance shows a resistive nature
by adding the virtual resistance, Q ∼ ω and P ∼ E droop control strategy
is implemented in the control loop [11] and is expressed as
(9.21)
(9.22)
where ω ∗ and ω are the inverter nominal and reference frequencies. E ∗
and E are the inverter nominal and reference voltage magnitudes. D p
and D q are the droop coef icients for controlling the output active
power P LPF and reactive power Q LPF through a low-pass ilter,
respectively. The calculated active power P LPF and reactive power Q LPF
are expressed as
(9.23)
(9.24)
where τ is the time constant of the low-pass ilter. V cα and V cβ are the
output voltages through Clarke transformation. I oα and I oβ are the
output currents through Clarke transformation.
Fig. 9.8 The control strategy of the PNM in on-line UPS system
(9.25)
(9.26)
where k pv and k pi are the proportional terms, and k rV and k ri are the
resonant term coef icients at ω. k vh is the resonant coef icient term for
the hth harmonics . The inner current loop is designed to
provide a suf icient damping and protect the inductor’s current from
an overcurrent. To increase the harmonic rejection ability of the output
voltage, the harmonic compensation device should be added to the
voltage controller of the DC/AC converter if nonlinear loads are
connected as sensitive loads[12]. Figure 9.2 illustrates the control
scheme for the recti ier, and it will not be repeated in this section.
Fig. 9.9 Power low in the EEM of the on-line UPS system
(9.28)
(9.29)
Fig. 9.10 The control strategy of the EEM operation in on-line UPS system
Fig. 9.11 The power low in the BTM operation for the on-line UPS system. (a) No load
condition. (b) Local load condition
The proposed control strategy of the BTM operation is depicted in
Fig. 9.12. Notice that the control strategy is very similar to that of the
EEM. The expression of the droop control strategy in this mode is
shown as
(9.30)
(9.31)
where P ref,i and Q ref,i are the reference active and reactive powers,
respectively. The integral term forces the output active power to
follow the reference one; meanwhile, the reactive power reference for
each inverter is determined by calculating the load reactive power and
has the relationship . It is pointed out that at no load
condition, reactive power reference should be set to zero .
From the discussion above, the BTM can be easily realized by changing
the reference of droop control from the TNM as well. In this way, the
seamless transfer among different modes can be realized without
affecting the output voltage of the UPS system.
Fig. 9.12 The control strategy of the BTM operation in the on-line UPS system
(9.32)
(9.33)
where P and Q are the instantaneous active and reactive powers of the
UPS system that lows out of the general line impedance. E and V are
the amplitudes of the inverter output voltage and the common bus
voltage, respectively, and φ is the power angle. Z and θ are the
magnitude and phase of the output impedance.
As the line impedance in UPS applications usually exhibits a
resistive characteristic, the power low of the UPS can be expressed as
follows:
(9.34)
(9.35)
(9.36)
(9.37)
where the operator Δ indicates a small-signal perturbation around the
UPS’s operating equilibrium point.
When there are some power luctuation during the EEM or the BTM,
expanding the proposed control strategy in (9.28)–(9.29) or (9.30)–
(9.31) results in the small-signal responses of the UPS voltage, which
are expressed as
(9.38)
(9.39)
(9.40)
(9.41)
where τ is the time constant of the low-pass ilter in the active and
reactive power calculation.
Considering that , and by the simple manipulation of
(9.36)–(9.37), the dynamic performance of the UPS system in the EEM
or BTM operation yields the following expression:
(9.42)
where
(9.43)
Fig. 9.13 Root locus with the proposed control strategy when the parameter changes: 0.0033 < k
< 0.0213
Table 9.2 Comparison among the different control strategy in response to the load step
Fig. 9.16 Active and reactive powers in transfer between the TNM and PNM operations
Fig. 9.17 Load voltage in transfer between the TNM and PNM operations
(2) Transition between the TNM and EEM Operations of the On-line UPS
System In this test, the seamless transition between the TNM and the
EEM is evaluated. As shown in Fig. 9.18a, when the operation mode of
the UPS shifts from TNM to EEM, the load active power provided by the
inverter drops to zero (blue waveform), and the 520 W of active power
is directly provided by the grid through the bypass switch (purple
waveform). Meanwhile, the reactive power is still powered by the
inverter (cyan waveform). On the contrary, when the operation mode
of the UPS shifts from the EEM to the TNM, the load active power is
delivered from the inverter, and the active power goes through the
bypass switch decreased to zero. Figure 9.18b and c show the load
voltage waveforms during the transition, it is indicated that the
seamless.
Fig. 9.18 Waveform of transition between the traditional normal mode and the enhanced eco-
mode. (a) Overall waveform; (b) zoomed-in waveform from the traditional normal mode to the
enhanced eco-mode; (c) zoomed-in waveform from enhanced eco-mode to the traditional normal
mode
(3) Transition Between the Traditional Normal Mode and the Burn-in Test
Mode of the On-line UPS System
(a)
Case 1: No Load Condition.
Under no load conditions, the UPS system should feed back the energy
drawn from the grid by the inverter when switching from normal to
burn-in test mode. In case 1, the inverter’s output active power is set to
500 W (blue waveform), and since all loads are connected to the UPS
system, the 500 W of output active power is fed back into the grid by
the bypass switch (Fig. 9.19a). Meanwhile, the load voltage during the
transition is shown in Fig. 9.19b and c, where it is observed that the
seamless transition is realized as the output voltage has no abrupt
change in voltage amplitude and frequency.
(b)
Case 2: R–L Load Is Connected During the Transition
Fig. 9.19 Waveform of transition between the traditional normal mode and the burn-in
test mode at no load condition. (a) Overall waveform; (b) zoomed-in waveform from the
traditional normal mode to the burn-in test mode; (c) zoomed-in waveform from the burn-
in test mode to the traditional normal mode
When the R–L load with 550 W of active power and 120 Var of reactive
power connecting at the UPS system, as can be seen in Fig. 9.20 that the
inverter provides the load power. When the UPS system transfers its
working mode from TNM to BTM, the active power drawn by the
inverter is set to be 850 W (blue waveform), and the extra 300 W of
output active power is fed back into the grid through the bypass switch
(Fig. 9.20a purple waveform). Meanwhile, the load voltage during the
transition is shown in Fig. 9.20b and c, where it is observed that the
seamless transition is realized as the output voltage has no abrupt
change in voltage amplitude and frequency.
Fig. 9.20 Waveform of transition between the traditional normal mode and the burn-in test mode
at R–L load condition. (a) Overall waveform; (b) zoomed-in waveform from the traditional
normal mode to the enhanced eco-mode; (c) zoomed-in waveform from enhanced eco-mode to
the traditional normal mode.
9.4 Conclusion
An on-line UPS system with multiple modes of operations is presented
in this chapter. The sensor-less IMC-based DC-link voltage control
strategy has been presented to provide a robust power delivery in
multi-mode operation. To achieve seamless transition among the
different modes of operations, we present some strategies for PV-aided
normal mode, enhanced eco-mode, and the burn-in test mode. By
utilizing these three modes, the cost charged from the grid can be
greatly reduced. Additionally, these modes of operations can be easily
achieved by adjusting the reference value of the inverter within the on-
line UPS system. Experiments proved that the proposed strategies are
effective.
References
1. Q. Zhu, Z. Yin, Y. Zhang, J. Niu, Y. Li, Y. Zhong, Research on two-degree-of-freedom internal
model control strategy for induction motor based on immune algorithm. IEEE Trans. Ind.
Electron. 63(3), 1981–1992 (2016)
[Crossref]
2. J. Yang, W.H. Chen, S. Li, L. Guo, Y. Yan, Disturbance/uncertainty estimation and attenuation
techniques in PMSM drives: a survey. IEEE Trans. Ind. Electron. 99, 1–1 (2016)
3. X. Sun, Z. Shi, L. Chen, Z. Yang, Internal model control for a bearingless permanent magnet
synchronous motor based on inverse system method. IEEE Trans. Energy Convers. 31(4),
1539–1548 (2016)
[Crossref]
4. W. Tan, C. Fu, Linear active disturbance-rejection control: analysis and tuning via IMC. IEEE
Trans. Ind. Electron. 63(4), 2350–2359 (2016)
5. C. Xia, Y. Yan, P. Song, T. Shi, Voltage disturbance rejection for matrix converter-based PMSM
drive system using internal model control. IEEE Trans. Ind. Electron. 59(1), 361–372 (2012)
[Crossref]
6. S. Li, H. Gu, Fuzzy adaptive internal model control schemes for PMSM speed-regulation
system. IEEE Trans. Ind. Inform. 8(4), 767–779 (2012)
[Crossref]
7. G. Stephanopoulos, H.-P. Huang, The 2-port control system. Chem. Eng. Sci. 41(6), 1611–1630
(1986)
[Crossref]
10. S. Kamran, H. Lennart, N. Hans-Peter, N. Staffan, T. Remus, Design, Control and Application of
Modular Multilevel Converters for HVDC Transmission Systems (Wiley-IEEE Press, 2016), p.
416
11. Y. Chen, J.M. Guerrero, Z. Shuai, Z. Chen, L. Zhou, A. Luo, Fast reactive power sharing,
circulating current and resonance suppression for parallel inverters using resistive-capacitive
output impedance. IEEE Trans. Power Electron. 31(8), 5524–5537 (2016)
[Crossref]
12. J. Lu et al., DC-link protection and control in modular uninterruptible power supply. IEEE
Trans. Ind. Electron. 99, 1–1 (2017)
13. M.A.G. de Brito, L.P. Sampaio, G. Luigi, G.A. e Melo, C.A. Canesin, Comparative analysis of MPPT
techniques for PV applications, in 2011 International Conference on Clean Electrical Power
(ICCEP) (2011), pp. 99–104
14. F. Khosrojerdi, S. Taheri, A.M. Cretu, An adaptive neuro-fuzzy inference system-based MPPT
controller for photovoltaic arrays, in 2016 IEEE Electrical Power and Energy Conference
(EPEC) (2016), pp. 1–6
15. J.C. Vasquez, J.M. Guerrero, M. Savaghebi, J. Eloy-Garcia, R. Teodorescu, Modeling, analysis, and
design of stationary-reference-frame droop-controlled parallel three-phase voltage source
inverters. IEEE Trans. Ind. Electron. 60(4), 1271–1280 (2013)
[Crossref]
16. Q. Sun, J.M. Guerrero, T. Jing, J.C. Vasquez, R. Yang, An islanding detection method by using
frequency positive feedback based on FLL for single-phase microgrid. IEEE Trans. Smart Grid
8(4), 1821–1830 (2017)
[Crossref]
17. S.D. Kermany, M. Joorabian, S. Deilami, M.A.S. Masoum, Hybrid islanding detection in
microgrid with multiple connection points to smart grids using fuzzy-neural network. IEEE
Trans. Power Syst. 32(4), 2640–2651 (2017)
[Crossref]
18. M. Bakhshi, R. Noroozian, G.B. Gharehpetian, Novel islanding detection method for multiple
DGs based on forced Helmholtz oscillator. IEEE Trans. Smart Grid 99, 1–1 (2017)
19. J.M. Guerrero, L. Hang, J. Uceda, Control of distributed uninterruptible power supply systems.
IEEE Trans. Ind. Electron. 55(8), 2845–2859 (2008)
[Crossref]
© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
J. Lu et al. (eds.), Advanced Control and Protection of Modular Uninterruptible Power Supply
Systems, Power Systems
https://doi.org/10.1007/978-3-031-22178-1_10
Xiaochao Hou
Email: houxiaochao@tsinghua.edu.cn
(10.2)
where ω i and V i are the angular frequency and voltage amplitude
references of an inverter-based i-th DG, respectively. ω ∗ and V ∗ imply
the values of ω and V at no load. P i and Q i are output active power and
reactive power of i-th DG. m and n are droop coef icients of P − ω and Q
− V control, respectively.
Figure 10.4 illustrates a typical control scheme of an inverter-based
DG, which consists of three control loops: the dip loop, the voltage loop,
and the current loop. In addition, a virtual-impedance control is used to
ensure a largely inductive output impedance [6].
Fig. 10.4 A typical control scheme of an inverter-based DG
(10.4)
(10.6)
2.
If microgrid works in grid-connected (GC) mode, STS is 1.
3. Microgrid works by mode transitioning from GC mode to IS mode
by disconnecting microgrid from utility grid due to grid fault,
switching STS from 1 to 0.
4.
As soon as the microgrid is connected back to the utility grid after
fault clearing, STS changes from 0 to 1.
Control Compensation Signals Δ ω ∗ , Δ V ∗ in IS Mode.
In IS mode, the compensation signals Δω ∗ and ΔV ∗ are set as
follows in order to restore the voltage/frequency amplitudes
(10.7)
(10.8)
(10.9)
(10.10)
(10.11)
(10.12)
where V i and δ i are output voltage amplitude and angle of DG-i. V c and
δ c are voltage amplitude and angle of PCC. X i is the linking reactance
between i-th DG and PCC.
In (10.10), Δω i is presented as follows from (10.4):
(10.13)
(10.15)
(10.17)
where
(10.18)
(10.21)
(10.22)
where
(10.23)
(10.24)
(v) System State-Space Model of P-ω Control For the tracking phase
synchronization, a new state variable is set to facilitate the stability
analysis.
(10.25)
The system linearization is given by combining (10.17)–(10.24)
(10.26)
where
(10.27)
(10.29)
(10.30)
where
(10.31)
(3) Eigenvalue Analysis of Aω To test the stability of the proposed
active power-frequency control scheme, the eigenvalues analysis of
matrix A ω is applied. According to the simulation system described in
Sect. 10.5, the root-locus plots are studied by varying the secondary
control gain k ω of (10.4) and tertiary PI control coef icients k pω, k iω of
(10.23).
Figure 10.8a shows the root-locus diagram as k ω increases from
0.1 to 10. As seen, the poles λ 1 and λ 2 are gradually moved to the
imaginary axis, which might lead to a poor dynamic response and even
instability. Therefore, k ω should choose a relatively small value.
Fig. 10.8 Root locus of system matrix A ω. (a) 0.1 < k ω < 10; (b) 0.3k pω < 10; (c) 0k iω < 10
(10.32)
where
(10.33)
where V i and V c are the voltage amplitudes of i-th DG and PCC. X i is the
line reactance between i-th DG and PCC.
In (10.32), ΔV i is presented as follows from (10.6)
(10.34)
In (10.32), ΔV ∗ is the voltage control compensation signal, derived
from the tertiary mode-supervisory controller (10.7)–(10.9)
(10.35)
where
(10.36)
(10.39)
where
(10.40)
(ii) Linearization of Secondary Distributed Control (10.34)
(10.41)
where
(10.42)
(10.43)
where
(10.44)
(10.48)
where
(10.49)
(v) System State Space of Q–V Control The entire system linearization
is given from (10.38)–(10.49)
(10.50)
In this chapter, given
(10.51)
(10.52)
where
(10.53)
(10.54)
Fig. 10.10 Simulated microgrid physical and communication models
Table 10.1 Comparison among the different control strategy in response to the load step
From Fig. 10.11a–b, the injected active and reactive powers into the
utility grid are P g = P ∗ = 10 kW and Q g = Q ∗ = 3 kVar at the beginning of
GC mode. After t = 1 s, the islanded microgrid is formed
and . In IS mode, Fig. 10.11d and f reveals that a nominal
system operation frequency and rated voltage amplitude
of are guaranteed in steady state. Moreover, Fig.
10.11c presents the grid-injected current, and Fig. 10.11e gives the
instantaneous voltage of PCC during t = 0.6 s∼1.4 s.
The simulation results of four DGs are shown in Fig. 10.12. From
Fig. 10.12a–b, DG1 ∼ DG4 achieve a seamless transition between GC
and IS modes with a satis ied voltage and frequency transition process.
Moreover, as seen in Fig. 10.12c–d, there is an accurate power sharing
between four DGs whether in GC or IS mode.
According to case 1, the proposed distributed hierarchical control
achieves all the required control targets and performances in both GC
and IS modes and under mode transitions from GC to IS.
10.4.2 Case 2: Active Synchronization from IS Mode to GC
Mode
In case 2, the proposed distributed hierarchical control scheme is
validated during the active synchronization (AS) from IS mode to GC
mode. The simulation results for case 2 are shown in Fig. 10.13.
Fig. 10.13 Results in case 2: (a) DGs voltage amplitudes; (b) DGs frequencies; (c) voltage
difference between grid and PCC; (d) voltage amplitudes of grid and PCC; (e) frequencies of grid
and PCC; (f) difference of frequency; (g) difference of phase angle; (h) grid current
Fig. 10.15 Results of ive DGs in case 4: (a) output active power; (b) output reactive power; (c)
frequencies; (d) voltage amplitudes
10.5 Conclusions
Our study presents a distributed hierarchical control scheme for AC
microgrids that takes into account all possible operating modes. It can
operate in both grid-connected modes, islanded mode and seamless
transition modes between them. The proposed control combines the
primary droop control, secondary distributed leader–follower control,
and tertiary mode-supervisory control. The mode-supervisory
controller at the top of the mode hierarchy provides compensation
signals to only a few DGs near the PCC. By using distributed consensus
protocol, the rest of the follower DGs would exchange information with
their neighbors. Finally, all DGs reach an agreement and achieve their
group targets in different ways. Overall, all major control targets are
met, and this method offers a cost-effective approach to building a
practical microgrid.
References
1. J. He, Y. Li, Analysis, design, and implementation of virtual impedance for power electronics
interfaced distributed generation. IEEE Trans. Ind. Appl. 47(6), 2525–2538, Nov. 2011.
[Crossref]
2. H. Han et al., Review of power sharing control strategies for islanding operation of AC
microgrids. IEEE Trans. Smart Grid 7(1), 200–215 (2016)
[Crossref]
3. Y. Deng et al., Enhanced power low control for grid-connected droop-controlled inverters
with improved stability. IEEE Trans. Ind. Electron. 64(7), 5919–5929 (2017)
[MathSciNet][Crossref]
4. F. Tang, J.M. Guerrero, J.C. Vasquez, D. Wu, L. Meng, Distributed active synchronization
strategy for microgrid seamless reconnection to the grid under unbalance and harmonic
distortion. IEEE Trans. Smart Grid 6(6), 2757–2769 (2015)
[Crossref]
5. Y. Sun, X. Hou, J. Yang, H. Han, M. Su, J.M. Guerrero, New perspectives on droop control in AC
microgrid. IEEE Trans. Ind. Electron. 64(7), 5741–5745 (2017)
[Crossref]
7. H. Han, Y. Liu, Y. Sun, M. Su, J.M. Guerrero, An improved droop control strategy for reactive
power sharing in islanded microgrid. IEEE Trans. Power Electron. 30(6), 3133–3141 (2015)
[Crossref]
8.
J.M. Guerrero, J.C. Vasquez, J. Matas, Hierarchical control of droop-controlled AC and DC
microgrids—a general approach toward standardization. IEEE Trans. Ind. Electron. 58(1),
158–172 (2011)
[Crossref]
9. H. Geng, S. Li, C. Zhang, G. Yang, L. Dong, B.N. Mobarakeh, Hybrid communication topology and
protocol for distributed-controlled cascaded H-bridge multilevel STATCOM. IEEE Trans. Ind.
Appl. 53(1), 576–584 (2017)
[Crossref]
10. X. Sun, Y. Tian, Z. Chen, Adaptive decoupled power control method for inverter connected DG.
IET Renew. Power Gener. 8(2), 171–182 (2014)
[Crossref]
11. M. Debasish, Power System Small Signal Stability Analysis and Control, ch. 5 (Academic, 2014,
New York), pp. 119–143
12. A. Bidram, A. Davoudi, Hierarchical structure of microgrids control system. IEEE Trans. Smart
Grid 3(4), 1963–1976 (2012)
[Crossref]
13. D.E. Olivares, A. Mehrizi-Sani, et al., Trends in microgrid control. IEEE Trans. Smart Grid 5(4),
1905–1919 (2014)
[Crossref]
14. J. Xiao, P. Wang, L. Setyawan, Hierarchical control of hybrid energy storage system in DC
microgrids. IEEE Trans. Ind. Electron. 62(8), 4915–4924 (2015)
[Crossref]
15. J. Wang, C. Jin, P. Wang, A uniform control strategy for the interlinking converter in
hierarchical controlled hybrid AC/DC microgrids. IEEE Trans. Ind. Electron. 65(8), 6188–
6197 (2018)
[Crossref]
17. M.H. Cintuglu, T. Youssef, O.A. Mohammed, Development and application of a real-time testbed
for multiagent system interoperability: a case study on hierarchical microgrid control. IEEE
Trans. Smart Grid 9(3), 1759–1768 (2018)
[Crossref]
18. Y. Han, H. Li, P. Shen, E.A.A. Coelho, J.M. Guerrero, Review of active and reactive power sharing
strategies in hierarchical controlled microgrids. IEEE Trans. Power Electron. 32(3), 2427–
2451 (2017)
[Crossref]
19. C. Cho, J. Jeon, J. Kim, S. Kwon, K. Park, S. Kim, Active synchronizing control of a microgrid.
IEEE Trans. Power Electron. 26(12), 3707–3719 (2011)
[Crossref]
Index
A
Active power sharing 57, 83, 109–111, 113–115,
121–122, 129, 130, 136
Adaptive virtual impedance 45–76, 110, 111
Anti-windup 28, 85, 132–135, 142, 148, 149, 153, 176
B
Burn-in test mode 182–183, 192–194
C
CAN bus 79, 85, 95, 98–99
Circulating current 47, 57, 59, 64, 75, 91–92, 109, 113114,
121, 122, 128, 129, 138, 139
Consensus control 132, 133, 198, 200, 220
Current sharing 101, 122, 128–132, 134, 135,
138–141
D
DC-link voltage 6, 11–18, 20–24, 26–28,
31–42, 109–111, 119–122,
125–135137–139, 141, 169–176,
186–189, 194
DC-link voltage protection (DCVP) 109–111, 114, 115, 120,
121, 125–135, 137–139, 141
Distributed coordination 200
Distributed secondary control 79–83, 85, 93, 98, 102, 104
Disturbance rejection 20, 39, 171, 173, 174, 188
Droop control 45–48, 111–113, 121, 128, 129, 134,
178, 180182, 199–202, 206, 211, 220
E
Enhanced eco-mode 180–182, 191, 193, 194
Enhanced state observer 32–34
H
Hierarchical control 199–205, 213, 216–221
I
Internal model control (IMC) 169–176, 187–189, 194
M
Microgrid 3–4, 42, 79–106, 132, 197–221
Modular uninterruptible power supply (UPS) system 1, 5–7,
45–76, 79–106, 109–122, 125–142
O
Overcurrent protection 145, 153
Overload 5, 27, 125, 135, 145–163
P
Parallel inverters 47, 48, 92
Power sharing 47, 55, 57, 58, 63, 68, 71–73, 7583, 85, 93, 98,
99, 101, 104, 109111–115, 118, 119, 121–122, 129,
130, 135, 136139, 141, 199, 201, 202, 217, 221
PV-aided normal mode 178–180, 194
R
Reduced-order generalized proportional-integral observer
(RGPIO) 12–17, 20–22, 27, 29
Robustness 16, 36, 42, 173
S
Seamless transfer 182
Seamless transition 4, 177, 180–184, 191, 192, 194,
197–221
Short circuit 130, 145–163
Sliding mode control 11–29
T
Three-phase converter 12, 21, 29, 31–45, 47, 84, 95, 139146,
148, 154, 155, 159, 160, 163, 169
U
Uninterruptible power supply (UPS), v 1–7, 16, 42,
45–76, 79–106, 109–122, 125–142,
145–163169–194, 197–199
V
Virtual impedance 45–76, 80, 81, 91, 92, 147, 200
Virtual resistance 48, 110–117, 119, 122, 128–132,
137, 178
Voltage control 11, 13, 16, 17, 31–42, 125, 135,
147169–176, 186–189, 194, 201, 202, 209, 215
Voltage source inverter (VSI) 45–76, 145–163