Journal of Modern Power Systems and Clean Energy

ISSN 2196-5625 CN 32-1884/TK

网刊加载中。。。

使用Chrome浏览器效果最佳,继续浏览,你可能不会看到最佳的展示效果,

确定继续浏览么?

复制成功,请在其他浏览器进行阅读

Intelligence-driven Grid-forming Converter Control for Islanding Microgrids  PDF

  • Issarachai Ngamroo 1
  • Tossaporn Surinkaew 2
  • Yasunori Mitani 3
1. Department of Electrical Engineering, School of Engineering, King Mongkut’s Institute of Technology Ladkrabang, Bangkok 10520, Thailand; 2. School of International and Interdisciplinary Engineering Programs, School of Engineering, King Mongkut’s Institute of Technology Ladkrabang, Bangkok 10520, Thailand; 3. Kyushu Institute of Technology, Fukuoka 804-8550, Japan

Updated:2025-07-23

DOI:10.35833/MPCE.2024.001157

  • Full Text
  • Figs & Tabs
  • References
  • Authors
  • About
CITE
OUTLINE

Abstract

In modern microgrids (MGs) with high penetration of distributed energy resources (DERs), system reconfiguration occurs more frequently and becomes a significant issue. Fixed-parameter controllers may not handle these tasks effectively, as they lack the ability to adapt to the dynamic conditions in such environments. This paper proposes an intelligence-driven grid-forming (GFM) converter control method for islanding MGs using a robustness-guided neural network (RNN). To enhance the adaptability of the proposed method, traditional proportional-integral controllers in the GFM primary control loops are entirely replaced by the RNN. The RNN is trained by a robustness-guided strategy to replicate their robust behaviors. All the training stages are purely data-driven methods, which means that no system parameters are required for the controller design. Consequently, the proposed method is an intelligence-driven model-less GFM converter control. Compared with traditional methods, the simulation results in all testing scenarios show the clear benefits of the proposed method. The proposed method reduces overshoots by more than 71.24%, which keeps all damping ratios within the stable region and provides faster stabilization. In comparison to traditional methods, at the highest probability, the proposed method improves damping by over 14.7% and reduces the rates of change of frequency and voltage by over 59.97%. Additionally, the proposed method effectively suppresses the interactions between state variables caused by inverter-based resources, with frequencies ranging from 1.0 Hz to 1.422 Hz. Consequently, these frequencies contribute less than 19.79% to the observed transient responses.

I. Introduction

GRID-FORMING (GFM) converters are crucial for grid voltage and frequency regulation, especially in 100% power electronics-based systems [

1], [2]. The GFM converter control strategy addresses various challenges in low-inertia/inertia-less microgrids (MGs) such as supporting grid frequency and voltage, providing pseudo-inertia, operating in standalone/islanding mode, and enabling black-start capability [1], [2]. However, traditional GFM controllers, specifically proportional-integral (PI) controllers embedded in tertiary, secondary, and primary control loops, encounter unsolvable and challenging problems when applied in weak MGs [3]. Additionally, these controllers face significant challenges in systems with high information uncertainties [4]. Research works including [1] and [5] have noted that fixed-parameter model-based controllers often struggle with adaptability and performance due to the reliance on accurate and up-to-date system models [1], [2]. Inadequate gain selection can negatively impact MG stability [1], [2], and the complexity of traditional GFM controllers can hinder their efficiency and practicality, which requires laborious tuning and optimization procedures [5]. These challenges become more pronounced in MGs with high penetration of distributed energy resources (DERs). Alternative methods such as intelligence-driven methods present effective solutions to these limitations. Intelligence-driven methods show promising capabilities for enhancing adaptability and optimizing performance in DER-based MG [6].

In [

7], a bi-objective control technique for MG load-frequency control is used to address uncertainties from renewable energy sources and optimize battery storage. This strategy can improve the characteristics of stability and frequency. However, the simplified load-frequency control model in [7] might neglect the impact of reactive power. A multi-layer interactive control strategy for MGs with DERs in [8] ensures the stability using optimal control parameters, but does not explicitly address DER intermittency. In [9], a novel virtual damping stabilizer in a battery enhances frequency stability and reduces capacity requirements, though potential DER intermittency drawbacks are not discussed. In [10], a control strategy is presented for managing transient power among inverter-based distributed generators, which ensures quick integration, and does not explicitly address the intermittency of DERs. Reference [11] develops a dynamic model for photovoltaic (PV)-based MGs to enhance stability, yet thoroughly explores PV generation intermittency. In [12], a settling-angle-based stability criterion is proposed, which uses the phase information of converter impedance and combines grid-following (GFL) and GFM modes. This criterion might overlook the impacts of reactive power. Reference [13] compares the control loops of GFL and GFM converters. A time-domain simulation of a 1.5 kW converter is used to analyze stability under various conditions. The results show that GFL control suits stiff grids, while GFM converter control is better for weak grids. In [14], a dual-port GFM converter control for power systems with AC/DC transmission is presented, which reduces the complexity by integrating both GFM and GFL functions and includes modeling and stability analysis for both emerging and legacy technologies. In [15], an MG control strategy that eliminates the need for a centralized GFM converter is proposed, which uses improved centralized control and droop-based power loop in DERs. The strategy supports both grid-connected and islanded modes, which ensures power sharing and smooth transitions without the need for islanding detection. In [16], an autonomous control architecture for grid-interactive inverters in MGs during utility outages is proposed, which highlights key tasks such as load sharing and voltage/frequency restoration when the MG isolates from the grid. The control architecture allows inverters to detect reconnection to the utility grid without communication, which enables a seamless transition from GFM mode to GFL mode. In [17], a GFM converter control method for solid oxide fuel cells to address voltage and frequency regulation is presented, which incorporates an incremental impedance technique to track the maximum power point and adjust output frequency during load transients.

As can be observed, traditional methods are mainly applied in previous research on stability and control. There is little focus on intelligence-driven methods like machine learning or neural networks, which could enhance system stability, decision-making, and real-time control and highlight a gap in the application of advanced techniques to power systems. As a result, the limitations of traditional methods emphasize the need for intelligence-driven methods, which could provide an advanced level of control for modern power systems [

18], [19]. From this perspective, recent research works show promising intelligence-driven methods as follows. In [20], a decentralized control strategy for PV-based DC MGs is presented, which enables the cooperation among multiple PV sources without communication. A predictive control strategy is introduced for GFM PV units, eliminating the need for external PI controllers, enhancing MG stability, reliability, and economy, while considering various operating conditions. Reference [21] uses an intelligence-driven controller for MG frequency and power sharing, outperforming traditional methods. Reference [22] introduces an adaptive control strategy for damping oscillations in DC MGs with hybrid power sources, incorporating a multi-loop voltage controller and virtual impedance loop. Reference [23] addresses renewable energy integration challenges with an adaptive droop controller and particle swarm optimization, enhancing stability and minimizing fluctuations. Reference [24] proposes virtual inertia and master-slave control using a recurrent probabilistic wavelet fuzzy neural network. Reference [25] offers a distributed intelligent secondary control method for AC MGs with a brain emotional learning-based controller. Reference [26] develops an adaptive dynamic programming-based strategy, optimizing the power outputs of micro-turbine and battery energy storage station (BESS) with PV integration. It is shown that this strategy outperforms PI-derivative, linear quadratic regulator, and fuzzy logic controllers. Despite valuable insights, these intelligence-driven methods exhibit limitations such as insufficient exploration of replacing physical controllers in GFM or GFL control loops and inadequate adaptive learning mechanisms for dynamic adaptation. Addressing these gaps paves the way for advanced intelligence-driven methods in MG control.

A robustness-guided neural network (RNN) framework is proposed in this paper, and the advantages of the proposed RNN framework compared with recent GFM converter control methods for DERs are summarized in Table I.

TABLE I  Comparision of Proposed RNN Framework with Recent GFM Converter Control Methods for DERs
Ref.MethodControl architectureAdaptability to uncertaintySupportiveness for islanding modeSupportiveness for inertia-less MGHistorical data learningSimplification of control loopComputational burden
[8] Multi-layer interactive control PI-based GFM Limited × Low
[11] Dynamic modeling for PV-based MGs PI-based GFM Limited × × Moderate
[12] Settling-angle-based stability criterion PI-based GFL and GFM Moderate × × × Low
[20] Decentralized control for PV-based DC MGs Predictive control for GFM PV unit High × × Moderate
This paper Proposed RNN Intelligence-driven GFM Very high High

Note that the intelligence-driven method may require significant computational resources and deep expertise for designing and training the proposed RNN framework. However, since this process is performed offline, the computational burden is limited to the training phase. Once trained, the control or decision-making response of the proposed RNN framework is faster and more effective compared with traditional methods in the literature. To address the identified gaps in the literature, this paper presents the following contributions.

1) A new RNN framework for GFM converter control is introduced, enhancing the robustness in MG uncertainties and improving the damping performance during rapid changes of voltage and frequency in islanding MGs.

2) Physical PI controllers in the GFM converter control topology are replaced with two RNN schemes integrated into pulse width modulation (PWM), which simplifies the control architecture, eliminates additional control loops, and enhances adaptability and responsiveness to uncertainties and disturbances.

3) The intelligence-driven GFM converter control method is proposed, which involves injecting the input into the PWM, and reduces the reliance on intermediate control layers. This method improves the responsiveness and adaptiveness of DERs. Although this method shares structural similarities with traditional PI-based control loops, the novelty lies in the specific way it is implemented and the resulting performance, adaptability, and stability enhancements.

4) The proposed RNN framework continuously updates the RNN model, allowing it to learn from new data, adjust parameters, and effectively mitigate adverse effects during recurrent events.

II. Overview of GFM Converter Control in Islanding MGs

An overview of multi-MGs with DERs is shown in Fig. 1, where Nmg is the number of MG modules; NL,i and NDER,i are the total numbers of loads and DERs of the ith sub-MG, respectively; PCC denotes point of common coupling; and Nsmg is the total number of sub-MGs within an MG.

Fig. 1  Overview of multi-MGs with DERs.

Overall, the MG consists of two layers, i.e., physical layer and cyber layer. The physical layer incorporates DERs equipped with control systems and other infrastructures. Specifically, the cyber layer is constructed hierarchically, consisting of centralized, local, and distributed layers. Each of the cyber layer receives information from the lower layer such as voltage, frequency, active/reactive power, and device/component statuses.

The hierarchical GFM converter control method chooses voltage and frequency references, i.e., vreft and freft, from the GFM converter to form its MG/sub-MG. These selected references are then transmitted to other GFL converters to synchronize with the corresponding GFM converters.

In the grid-connected scenario, multi-MGs, each with sub-MGs, operate in grid-tied modes. vreft and freft are derived from PCC using a phase-locked-loop strategy, which can ensure a synchronization with the main grid through centralized control and account for variable delays when vreft=vPCCt and freft=fPCCt, where vPCCt and fPCCt are the voltage and frequency of the PCC at time t, respectively. In the islanding scenario, when the MG disconnects from the grid, the voltage and frequency references change. These references are set to be the values of the DER with the highest power output. Consequently, the voltage reference is vreft=vDERt, and the frequency reference is fref=fDERt, where vDERt and fDERt are the voltage and frequency of the DER at the highest power output at time t, respectively. Accordingly, the DER with the highest capacity operates in GFM mode, while others operate in GFL mode. The decentralized control is used with limited observable output signals and specific local communications.

We also consider the time delay in all input and output signals across different sub-MGs, which accounts for the impact of communication delays and reflects how these delays influence both local and centralized control processes. As a result, the time t can be represented as t-τDER,lt, where τDER,lt is the local communication delay specific to the corresponding lth DER within the smgth sub-MG, and with conditions τDER,1tτDER,2tτDER,3tτDER,NLt. This adjustment accounts for the effect of time delay. For example, it accounts for the delay of the signal by shifting the original signal by τDER,lt. This delay arises from the exchange of information between local controllers and their associated DERs. The value of τDER,lt is always less than τct, i.e., τDER,ltτct, where τct is the communication delay between local controllers and the centralized controller. Both τDER,lt and τct depend on time and vary over time. Accordingly, τDER,lt and τct are represented by variable time delays. The concept of variable time delays can be found in [

27], [28]. The distinction between these two delays lies in the fact that τDER,lt is associated with localized interactions within a sub-MG, whereas τct is the larger-scale communication delay across the entire system. Moreover, if sub-MGs are further isolated from the others, voltage and frequency references can be obtained using vreft=vsmg,DERt and fref=fsmg,DERt, where vsmg,DERt and fsmg,DERt are the voltage and frequency of the DER in the corresponding smgth sub-MG with the highest power outputs at time t, respectively. These references are based on the DER with the highest power outputs within each sub-MG at time t. This scenario uses distributed control to eliminate communication issues.

III. Proposed Intelligence-driven GFM Control Method for Replacing GFM Control Loops

A. Proposed RNN Schemes

An overview of a GFM converter is shown in Fig. 2. SVSC1-SVSC6 are the modulation signals of insulated gate bipolar transistor switches of a GFM converter. The dynamics of low-inertia MGs can be expressed using (1), which establish a connection between system outputs yt and underlying dynamics:

yt=fyxt,ut,dt (1)

Fig. 2  Overview of GFM converter.

where xt is the state variable; ut=-Kyt is the input vector from the intelligence-driven method and K denotes the RNN; dt is the disturbance vector originating from wind and PV generators; and fy is the nonlinear function incorporating these inputs to determine yt.

In the proposed RNN framework, ut takes on a more direct role by being seamlessly integrated into the PWM process. By directly injecting ut into the PWM, the physical PI controllers are replaced by the RNN within the GFM converter control loops. Note that the proposed RNN framework shares similarities with traditional PI-based control methods, which simplifies the control architecture and overcomes challenges of physical controllers, such as limited flexibility and reliance on accurate mathematical models. In RNN scheme 1, as shown in Fig. 3, the traditional PI controllers in the voltage and current control loops are replaced with intelligence-driven modules. These modules include RNNvc,d and RNNvc,q for voltage control, RNNic,d and RNNic,q for current control, and RNNθ for phase angle control. Here, the modules use RNNs, which provide enhanced performance and offer greater adaptability compared with traditional PI controllers.

Fig. 3  Proposed RNN scheme 1.

Accordingly, ut of the RNN scheme 1 can be written as:

ut=vdq*tidq*tθ*tT (2)

where T is the transpose operator; vdq* and idq* are the dq-axis commanded signal vectors from voltage and current control loops, respectively; and θ* is the commanded signal from phase angle control loop.

In (2), vd* and vq* are transformed to va*, vb*, and vc* using the inverse Park transformation, which are consequently used to command the PWM of the GFM converters. From Fig. 3, expressing (2) will yield:

vdq*t=RNNic,dqidq*t-32ωtLf iqdt (3)
idq*t=RNNvc,dqvdqreft-vdqt+idqt-32ωCfvqdt (4)
θ*t=2πRNNθtt+ΔtPt+kffreftdt (5)

where Δ denotes the small deviation; iqd and vqd are the qd-axis current and voltage, respectively; vdqref is the reference signal of vdq; P is the vector representing active power output; kf is the frequency control gain; ω is the angular velocity; fref is the vector representing frequency reference; and Cf and Lf are the capacitance and inductance, respectively.

Figure 4 shows RNN scheme 2, which employs a unified RNNvic module integrating voltage and current control loops.

Fig. 4  Proposed RNN scheme 2.

The total output number of RNN scheme 2 differs from that in RNN scheme 1. RNN scheme 2 forms a unified model. Additional variables are introduced as Δvdqt=vdqreft-vdqt and Δidqt=idqreft-idqt. Accordingly, idqreft is the reference signal of idqt. With these considerations, ut of RNN scheme 2 can be expressed as:

ut=vdq*tθ*tT (6)
vdq*t=RNNvicΔvdqt,Δiqdt (7)
θ*t=RNNvicPt,freft (8)

B. Robust Intelligent-driven Controller Design

To replace traditional PI controllers with the proposed method, specific expressions using (3)-(5), (7), and (8) are initially defined. K is introduced. Consequently, ut can be modified as:

ut=foKWh,n,bh,n,bo,t,yt+𝒫t (9)
ut=foh=1Nhn=1NnWh,ny˜h,nt+bh,n+bo+𝒫t (10)

where subscripts h=1,2,,Nh and n=1,2,,Nn are the layer and node indices, respectively; Nh and Nn are the total numbers of hidden layers and nodes, respectively; 𝒫 represents any known physically-guided equation, if applicable; fo is the nonlinear activation function for the output layer; y˜h,n is the transformed output vector y that passes through a node of a hidden layer (at an input layer h=1, y˜=y); Wh,n and bh,n are the weight and bias vectors, respectively; and bo is the bias vector of the output layer.

For the RNN scheme 1 shown in Fig. 3 with (3)-(5), the output signals of all individual RNNs are explicitly expressed in the form of (10) as:

idq*t=foh=1Nhn=1NnWh,n1v˜dq,h,nt-ηvdqreft+bh,n1+bo1+idqt-32ωtCfvqdt (11)
vdq*t=foh=1Nhn=1NnWh,n2i˜dq,h,n*t+bh,n2+bo2-32ωtLfiqdt (12)
θ*t=fo2πh=1Nhn=1NnWh,n3wθP˜h,nt,f˜h,nreft+bh,n3+bo3 (13)
  wθP˜h,nt,f˜h,nreft=tt+ΔtP˜h,nt+kff˜h,nreftdt (14)

where v˜dq,h,n, i˜dq,h,n*, P˜h,n, and f˜h,nref are the transformed output vectors of vdq, idq*, P, and fref, respectively. In (11), if h=1, η=1; otherwise, η=0.

For the RNN scheme 2 illustrated in Fig. 4, the output signals from the RNNvic are expressed in the form of (10) as:

vdq*t=foh=1Nhn=1NnWh,n1wvicv˜dq,h,nt,i˜dq,h,nt+bh,n1+bo (15)
θ*t=foh=1Nhn=1NnWh,n2wθP˜h,nt,f˜h,nreft+bh,n2+bo2 (16)
wvicv˜dq,h,nt,i˜dq,h,nt=αdqv˜dq,h,nt-ηvdqreft+βdqi˜dq,h,nt-ηidqreft (17)
wθP˜h,nt,f˜h,nreft=αθP˜h,nt+βθf˜h,nreft (18)

where αdq and βdq are the weighting vectors of the first layer (h=1) of v˜dq,h,n and i˜dq,h,n, and η=1, respectively. For other layers (h>1), αdq+βdq=1 and η=0, and consequently wvic=αdv˜d,h,nt+βdi˜d,h,nt+αqv˜q,h,nt+βqi˜q,h,nt. Here, the vectors Wh,n1, bh,n1, bo1, Wh,n2, bh,n2, bo2, Wh,n3, bh,n3, and bo3 are trained and optimized for the RNN scheme 1. The vectors Wh,n1, bh,n1, bo1, αdq, βdq, Wh,n2, bh,n2, bo2, αθ, and βθ are trained and optimized for the RNN scheme 2. In the training stage, these parameters are automatically optimized and updated by a stochastic gradient descent method.

C. Data-driven Training Strategy

Let t-τct and t-τlt be the time with variable delays affected centralized and local control platforms, respectively; and ymtyt be the measured time-varying output vector in a form of time-series data containing crucial MG dynamics behavior. Accordingly, the vectors ymt and dmt can be expressed as:

ymt=Λte-ξttcosωdtt+ϕt (19)
dmt=ΔP¯DERt+ΔP¯DLt (20)

where dmd and ymy; Λ is the signal amplitude; ξ is the damping coefficient; ωd is the angular frequency representing the frequency at which the poorly-damped oscillation would occur; ϕ is the vector of phase shifts; ΔP¯DER is the vector of average active power output deviations of DERs; and ΔP¯DL is the vector of average active power output deviations of distributed loads.

Here, ymt=ΔF¯t,ΔV¯tT, where ΔF¯ and ΔV¯ are the vectors of average values of bus frequency and voltage deviations of DERs, respectively. To get ymt and dmt from measurements, the 100-ms stamped time (denoted by Ts) is regarded. In any rth signal pattern within a moving window of size k, the vectors ymt in (19) and dmt in (20) can be represented by their discrete-time counterparts, ymr and dmr, as:

ymr=ymrTs     ymr+1Ts          ymr+kTs (21)
dmr=dmrTs     dmr+1Ts          dmr+kTs (22)

Both RNN scheme 1 and RNN scheme 2 are trained by prioritizing robustness and high damping as well as reducing the rates of changes of frequencies and voltages of all buses in the MG, minimizing the loss function L, as:

L=1Njj=1Njr=1Nr1M (23)
M=γrob,jr-1+γdamp,jr+γRoC,jr-12 (24)

where γrob,jr, γdamp,jr, and γRoC,jr are the indices representing the MG robustness, damping performance, and ability to reduce the rates of changes of frequency and voltage, respectively; Nj is the total number of training data sets; and Nr is the total number of signal patters in each data.

The term γrob,jr-1 serves the purpose of enhancing robustness in both frequency and voltage control loops, which is considered by:

γrob,jr=G˜F,jrG˜V,jrT (25)

where operator returns the -norm of its argument; and G˜F,jr and G˜V,jr denote the closed-loop systems including RNN, which are identified by ΔF¯, dmr and ΔV¯, dmr, using the sub-space state-space identification, respectively.

The overall damping performance is determined by:

γdamp,jr=os=1Nosξ˜j,osr=ξ˜j,1r+ξ˜j,2r++ξ˜j,Nosr (26)

where Nos is the total number of oscillation modes; and ξ˜j,osr is the estimated damping calculated from the identified matrices G˜F,jr and G˜V,jr.

The rates of changes of frequency and voltage can be evaluated by:

γRoC,jr=RoCoFjr+RoCoVjr (27)

where RoCoFjr and RoCoVjr are the rates of changes of average frequency and voltage, respectively.

Let γrob,jr-1,γdamp,jr,γRoC,jr-1χjr, where χjr can be either γrob,jr-1, γdamp,jr, or γRoC,jr-1, the values obtained from (25)-(27) are normalized (denoted by operator norm) into a range of 0 and 1 by:

normχjr=χjr-minχjmaxχj-minχj=χ¯jr (28)

where over-line “  ¯” means the normalized variable.

By applying (28), variables γ¯rob,jr-1, γ¯damp,jr, and γ¯RoC,jr-1 can be obtained. Consequently, these values are substituted into (23) to calculate L for r,j. Therefore, by employing the norm function in (28), no weighting factors are required to balance the terms γrob,jr-1, γdamp,jr, and γRoC,jr-1 in (23). Here, the dynamics of the MG are apparent in the measured signals ymr, as expressed in (19) and (21). In the training process of the RNN and the computation of the loss function (23), only ymr and dmr in the forms of (21) and (22) are used. Exact MG parameters and their updates are not required. These parameters can be uncertain or change due to variations in MG topology, disconnection/synchronization from/to the main grid, and islanding conditions. The complete process of training and data preparation is presented in Fig. 5.

Fig. 5  Complete process of training and data preparation.

IV. Performance Evaluation

A. Test MG Description

The RNN schemes have undergone the validation within the MG with DERs, as illustrated in Fig. 6. In MG1, there are three sub-MGs with a maximum load of 23 kVA and a maximum generation of 69 kVA. Sub-MG1 includes five loads and three wind turbines (WTs). Sub-MG2 comprises three loads. Sub-MG3 is composed of three loads, one PV, and one WT. If the MG is disconnected from the main grid, LOC1 controls all DERs with assistance from DTC1 (sub-MG1), DTC2 (sub-MG2), and DTC3 (sub-MG3). In MG2, two sub-MGs have a maximum load of 27 kVA and a maximum generation of 57.5 kVA. Sub-MG1 includes five loads and five PVs, while sub-MG2 comprises four loads and three WTs. If the MG is disconnected, LOC2 controls all DERs with assistance from DTC1 (sub-MG1) and DTC2 (sub-MG2). In MG3, there are four sub-MGs with a maximum load of 47 kVA and a maximum generation of 105 kVA. Sub-MG1 includes five loads and three WTs. Sub-MG2 consists of two loads and five PVs. Sub-MG3 comprises three loads and three WTs. Sub-MG4 includes three loads and five PVs. In islanding mode, LOC3 controls all DERs with assistance from DTC1 (sub-MG1), DTC2 (sub-MG2), DTC3 (sub-MG3), and DTC4 (sub-MG4).

Fig. 6  Test MG system with DERs and proposed hierarchical GFM converter control structure consisting of three MGs and nine sub-MGs.

Under a normal MG condition, all MGs are connected to the utility grid to assist in supporting a maximum load of 200 kVA. The DERs in these MGs are modeled as GFL converters. Additionally, a 150 kVA synchronous generator is connected to the utility grid. This situation falls under the classification of a grid-connected scenario, as detailed in Section II. Under these circumstances, the operations of DERs and loads are as follows.

1) At the utility grid, the synchronous generator at the PCC is operated at 80%-95% of its maximum capacity.

2) In MG1, all DERs are operated at 50%-65% of their maximum capacities, and all loads fluctuate at 75%-90% of their maximum capacities.

3) In MG2, all DERs are operated at 45%-55% of their maximum capacities, and all loads fluctuate at 60%-75% of their maximum capacities.

4) In MG3, all DERs are operated at 50%-65% of their maximum capacities, and all loads fluctuate at 80%-95% of their maximum capacities.

B. Setup RNN Parameters

First, the total number of possible scenarios for the test system shown in Fig. 6 is calculated. There is a total of 6+6+18=30 possible scenarios to be trained for the RNNs. Second, for each scenario, the 50 s time-series data in (19) and (20) are collected within 1250 random operating points of MG.

In an individual random operating point, all loads randomly fluctuate by ±50% from their normal operating points specified in Section IV-A. Additionally, potential random outages and out of services of CTC, LOCs, and/or DTCs are considered at an individual random operating point. By incorporating these uncertainties into the training processes, the trained RNNs can resiliently operate under various grid-tied and islanding conditions with high robustness against communication issues. Then, the measured data sets in (19) and (20) are transformed into their discrete counterparts using (21) and (22) with Ts=100  ms or 0.1 s. A moving window length of k=50  s is used, and L can be calculated as specified in (23). The maximum number of iterations is set to be 60, and the hyperbolic tangent function is used as the nonlinear activation function.

With these settings, the training processes take approximately 13 hours and 17 minutes for RNN scheme 1, and 15 hours and 52 minutes for RNN scheme 2. It is found that the training time of RNN scheme 1 is shorter than that of the RNN scheme 2 since RNN scheme 1 incorporates physically-guided terms 𝒫t in the developed model, such as idqt-32ωtCfvqdt in (11) and 32ωtLfiqdt in (12), whereas 𝒫t=0 for RNN scheme 2. With known 𝒫t, as a result, the training process of RNN scheme 1 converges more quickly to minimize the targeted loss function in (23). However, if certain parameters such as idqt, ωt, Cf, and/or vqdt are missing or cannot be measured at time t, this creates an obstacle. Such issues may arise due to communication problems or other related factors, which can hinder the design process of RNN scheme 1 and may lead to a degradation in its performance. To validate the proposed method, the trained RNN scheme 1 and RNN scheme 2 are compared with traditional GFM converter control loops using model-based PI controllers, while additional signals from the RNN are incorporated into the junction points at inner-current control and phase angle control loops. Here, the compared strategy is referred to as PI with RNN.

C. Performance of Proposed RNN Schemes in Critical Scenarios

1) Case Study 1

In case study 1, the assumed events include periodic fluctuations of active and reactive power occurring every 1-5 s in all DERs and loads in a range of ±10% from their current operating points. Prior to disconnection, all DERs take vreft and freft from the PCC. In this state, all DERs operate in the GFL modes. The disturbance dt applied to case study 1 consists of two main factors: the unplanned disconnection event, which leads to the separation of sub-MG3 from the grid, and the fluctuations in the outputs of intermittent DERs.

Initially, MG1, MG2, and MG3 are interconnected and also connected to the main grid. At t=4.1 s, an unplanned disconnection occurs, separating sub-MG3 of MG3 from the rest of the grid. This event results in the formation of an islanding sub-MG3. Case study 1 results in the transition from grid-tied mode to sub-islanding condition. WT2 of sub-MG3 provides the highest active power support to sub-MG3 compared with the others, which operates in GFM mode to generate references for them. After t=20 s, intermittent DERs increase in sub-MG3 within a range of ±40% from their current operating points.

Accordingly, Fig. 7 shows the time-domain simulation of case study 1, presenting transient responses of RoCoV and RoCoF during the disconnection. Considering sub-MG3, the results reveal three significant periods including: ① before the disconnection of sub-MG3; ② during the reconnection; and ③ after the disconnection with high intermittency. The PI with RNN exhibits responses over different frequencies, i.e., from 0.327 Hz to 1.422 Hz. Modal analysis confirms that these modes mainly involve the state variables idq and vdq of WTs. These variables interact across all areas, contributing over 72.87% of the participation factors. The PI with RNN cannot suppress high-frequency oscillations. As a result, the transient responses are dominated by these high-frequency modes. In contrast, RNN scheme 1 and RNN scheme 2 can suppress these modes effectively. While some oscillations still occur, their amplitudes are much smaller compared with the PI with RNN. As a result, the PI with RNN struggles to stabilize RoCoV and RoCoF in case study 1, while RNN scheme 1 and RNN scheme 2 show better performance. Specifically, before the disconnection, the compared strategy (PI with RNN) exhibits higher fluctuations. In this stage, the values of RoCoV and RoCoF oscillate severely in the case of compared strategy. After the disturbance, both RNN schemes demonstrate better abilities to damp the fluctuations.

Fig. 7  Time-domain simulation of case study 1.

2) Case Study 2

A time-domain analysis is conducted to explore various re-connection scenarios. Case study 2 replicates the conditions in case study 1, except for the failures in receiving input signals for LOCs across all DERs during the first period between t=0 and t=4 s. These failures obstruct the transmission of local input-output signals to the RNNs, resulting in all MGs being entirely controlled by the DTCs of their corresponding sub-MGs. dt applied in case study 2 is similar to that in case study 1. However, it includes the additional effect of a complete failure to receive signals from all local controllers at the beginning of the simulation.

Initially, from t=0 to 4 s, each MG operates independently in an islanding mode. In MG1, WT3 in sub-MG1, PV1 in sub-MG2, and WT1 in sub-MG3 operate in the GFM mode, producing the highest active power within their respective sub-MGs. Other DERs in each sub-MG operate in the GFL mode, deriving V/f references from the corresponding DERs in GFM modes. Similarly, in MG2, PV5 in sub-MG1 and WT2 in sub-MG2 operate in GFM modes. In MG3, WT3 in sub-MG1, PV5 in sub-MG2, WT3 in sub-MG3, and PV5 in sub-MG4 also operate in GFM modes. To ensure the synchronization and form the sub-MGs, these sources generate the highest active power and produce V/f references for the GFL control loops of other DERs within their respective sub-MGs.

At t=4.1 s, MG3 reconnects with MG2, establishing the islanding mode between MG2 and MG3. During this period, it is assumed that the LOCs are restored, and the information from these LOCs can be transmitted to the CTC, thereby allowing the interconnected islanding MG to be governed by the CTC. Throughout this period, the CTC detects that the WT1 in sub-MG2 of MG2 operates in GFM mode as it provides the highest power supported to the grid. As a result, other DERs in this interconnected islanding MG derive V/f references from WT1 in sub-MG2 of MG2 for their GFL active and reactive power control loops. Later, at t=15.6 s, MG1 reconnects to the others. The MG is controlled by the CTC. During this period, the CTC detects that WT1 in sub-MG2 of MG2 is generating the highest active power to the grid. Consequently, WT1 in sub-MG2 of MG2 operates in the GFM mode, supplying V/f references to the GFL converters of other DERs for their active and reactive power control.

Following this, at t=26.1 s, the islanding MG (comprising MG1, MG2, and MG3 connected together) reconnects to the utility grid. During this transition from islanding condition to grid-connected condition, the synchronous generator is not available, and the load is approximately 87.5±5 kVA and no inertia is supported. During this period, akin to the preceding timeframe, WT1 in sub-MG2 of MG2 continues to operate in the GFM mode. Consequently, Fig. 8 shows time-domain simulation results of case study 2, illustrating the transient responses of RoCoV and RoCoF under these re-synchronization conditions. After reconnecting sub-MG3, the RNN schemes restore RoCoV and RoCoF within 2.18 s. They reduce oscillation amplitudes by 66.32% and 47.26% compared with the PI with RNN. When sub-MG1 reconnects to the others, the amplitudes of these signals are similarly reduced by 135.15% and 127.80%, respectively. Moreover, when all MGs reconnect to the main grid, the PI with RNN shows very high amplitudes, especially at the first peak. Eventually, the system becomes unstable. In contrast, the RNN schemes maintain the fluctuations, keeping them within acceptable ranges for RoCoV and RoCoF. The results clearly show that the proposed RNN schemes perform better in reducing both RoCoV and RoCoF during re-synchronization under all conditions.

Fig. 8  Time-domain simulation results of case study 2.

D. Probabilistic Analyses Across a Wide Range of MG Operating Conditions

In this subsection, the probability analyses are carried out across 1000 different scenarios. It is important to note that these scenarios are entirely different from those used to train the RNNs. To verify this, in each scenario, time-series data over a duration of 50 s are gathered with a sampling interval of Ts=0.1 s. At each time step, the values for RoCoV, RoCoF, and ξ are collected, totaling 3×50/0.1=1500 points for further analysis. After obtaining these values for all scenarios, the data sets consist of a total of 3×500×1000=1.5×106 data points (0.5×106 for each of RoCoV, RoCoF, and ξ). Subsequently, a probability analysis method is utilized to estimate the densities of each of these data sets.

First, the focus is on evaluating the performance of hierarchical GFM converter control in scenarios where CTC, LOC(s), and/or DTC(s) are unexpectedly unavailable. The conditions are kept consistent where DERs and loads randomly vary in a range of ±25% from their normal operating points. In each scenario, we introduce additional factors of random failure, which include failures of CTC, LOC(s), DTC(s), or unexpected islanding. In this scenario, the verification is conducted on the PI with RNN while employing four GFM converter controls, which are centralized GFM converter control, local GFM converter control, distributed GFM converter control, and proposed hierarchical GFM converter control.

For the centralized GFM converter control, the assumption is made that this control has the capability to monitor global signals from all MGs with a random delay τc100,700ms. However, in the event of a failure at the CTC, the GFM converter loses its ability to receive all inputs for controlling MGs. For the local GFM converter control, the delay is randomly in the range τl0,120ms for each LOC. Besides, if any failure occurs at one or more LOCs, DTCs remain available to transmit signals to GFM converter control loops. However, the LOCs may restrict the ability to observe all signals. For the distributed GFM converter control, unlike centralized GFM converter control, no global signals are employed for GFM converter control. This control performs effectively, particularly in islanding scenarios without inertia support from the main grid. It is clear that the proposed hierarchical GFM converter control effectively handles unexpected islanding scenarios in a multi-MG system.

In Fig. 9, probabilistic analysis results are presented after applying random scenarios, where active and reactive power outputs vary stochastically in a range of ±25%. The analysis details RoCoV, RoCoF, and ξ. Under centralized GFM converter control, the ranges of RoCoV, RoCoF, and ξ (including potential negative values of ξ) are much larger. Specifically, RoCoV and RoCoF range from -0.396,0.377p.u./s and -0.467,0.495Hz/s, respectively, which show that stabilizing these parameters is more challenging during CTC failures. Local, distributed, and the proposed hierarchical GFM converter controls exhibit similar values for RoCoV and RoCoF, with ranges of -0.056,0.084p.u./s and -0.185,0.152Hz/s, respectively, effectively mitigating fluctuations during islanding and controller failures. However, local and distributed GFM converter controls exhibit small negative values of ξ, with values of -0.015 and -0.008, respectively. In contrast, the proposed hierarchical GFM converter control ensures the stability for all dominant modes without negative ξ, but with ξ falling within the range 0.009,0.388, and the highest probability is 9.024%. As can be observed, the proposed hierarchical GFM converter control outperforms other controls.

Fig. 9  Probabilistic analysis under conditions of random delays and failures of CTC, LOC, and/or DTC. (a) RoCoV. (b) RoCoF. (c) ξ.

Following this, performances of the RNN scheme 1 and RNN scheme 2 are evaluated. Here, MG3 is assumed to be isolated from both MG1 and MG2, creating an islanding condition for sub-MG4.

Two separate probabilistic analyses are conducted within these islanding sub-MGs of MG3: analysis 1 considers sub-MG1, sub-MG2, and sub-MG3 together, while analysis 2 focuses solely on sub-MG4. Figure 10 shows the probabilistic analyses, where active and reactive power outputs vary randomly by ±50% from their normal operating points. The outcomes for analysis 1 are presented in Fig. 10(a), while the outcomes for analysis 2 are shown in Fig. 10(b).

Fig. 10  Probabilistic analyses. (a) Analysis 1: results of MG3 with sub-MG1, sub-MG2, and sub-MG3. (b) Analysis 2: results of MG3 with sub-MG4.

In analysis 1, the variations of RoCoV and RoCoF are similar across all strategies. For the proposed RNN schemes, RoCoV and RoCoF fall within the ranges -0.227, 0.198p.u./s and -0.211,0.195Hz/s, respectively.

For the PI with RNN, the ranges of RoCoV and RoCoF are -0.305,0.311p.u./s and -0.353,0.357Hz/s, respectively. However, the difference is observed in ξ. The proposed RNN schemes do not show negative damping, while the PI with RNN shows values of ξ in the range -0.031,0.0745. In analysis 2, PI with RNN and RNN scheme 1 exhibit poorer performances in sub-MG4 compared with the RNN scheme 2 in managing RoCoV variations. The results for RoCoV range from -0.399,0.409p.u./s in the case of the RNN scheme 1. In contrast, for RNN scheme 2, the range of RoCoV is only -0.304,0.212p.u./s. Considering ξ, the RNN scheme 1 shows a significant improvement in damping, with values in the range 0.009,0.302. The highest probability occurs at a damping value of 0.1927 for RNN scheme 2. Similar to analysis 1, the worst results are observed in the PI with RNN, where negative damping might occur and fall within the range -0.071,0.144.

V. Conclusion

This paper introduces an intelligence-driven GFM converter control method for islanding MGs with DERs. The proposed method can handle both grid-tied and islanding scenarios by adaptively adjusting references in real time to mitigate communication issues, ensuring reliable operation and optimal performance under varying conditions. The proposed RNN scheme 1 employs separate RNN modules for voltage, current, and phase angle control, while the proposed RNN scheme 2 integrates them into a single RNN module to minimize control interaction. The RNN design incorporates a convolutional neural network structure focused on robustness, damping, and minimizing voltage and frequency changes during training. The proposed method includes normalizing indices and calculating a loss function to manage uncertainties and parameter changes during critical MG operations. Time-domain simulations validate the effectiveness of the proposed RNN schemes in mode transitions and disturbances, showing superior performance over a comparative strategy. Probabilistic analysis demonstrates both RNN schemes reduce voltage and frequency fluctuations, with RNN scheme 2 particularly effective in minimizing control loop interaction.

References

1

R. Musca, A. Vasile, and G. Zizzo, “Grid-forming converters: a critical review of pilot projects and demonstrators,” Renewable and Sustainable Energy Reviews, vol. 165, p. 112551, Sept. 2022. [Baidu Scholar] 

2

H. Zhang, W. Xiang, W. Lin et al., “Grid forming converters in renewable energy sources dominated power grid: control strategy, stability, application, and challenges,” Journal of Modern Power Systems and Clean Energy, vol. 9, no. 6, pp. 1239-1256, Dec. 2021. [Baidu Scholar] 

3

M. H. Khooban, “An optimal non-integer model predictive virtual inertia control in inverter-based modern AC power grids-based V2G technology,” IEEE Transactions on Energy Conversion, vol. 36, no. 2, pp. 1336-1346, Jun. 2020. [Baidu Scholar] 

4

A. Vosughi, A. Tamimi, A. B. King et al., “Cyber-physical vulnerability and resiliency analysis for DER integration: a review, challenges and research needs,” Renewable and Sustainable Energy Reviews, vol. 168, p. 112794, Oct. 2022. [Baidu Scholar] 

5

L. Wang, B. Zhang, Q. Li et al., “Robust distributed optimization for energy dispatch of multi-stakeholder multiple microgrids under uncertainty,” Applied Energy, vol. 255, p. 113845, Dec. 2019. [Baidu Scholar] 

6

M. Zhang, P. I. Gómez, Q. Xu et al., “Review of online learning for control and diagnostics of power converters and drives: algorithms, implementations and applications,” Renewable and Sustainable Energy Reviews, vol. 186, p. 113627, Oct. 2023. [Baidu Scholar] 

7

A. Rafiee, Y. Batmani, A. Mehrizi-Sani et al., “Load frequency control in microgrids: a robust bi-objective virtual dynamics technique,” IEEE Transactions on Power Systems, vol. 39, no. 4, pp. 5981-5990, Jul. 2024. [Baidu Scholar] 

8

M. S. Toularoud, M. K. Rudposhti, S. Bagheri et al., “A hierarchical control approach to improve the voltage and frequency stability for hybrid microgrids-based distributed energy resources,” Energy Reports, vol. 10, pp. 2693-2709, Jun. 2023. [Baidu Scholar] 

9

S. E. Sati, A. Al-Durra, H. Zeineldin et al., “A novel virtual inertia-based damping stabilizer for frequency control enhancement for islanded microgrid,” International Journal of Electrical Power & Energy Systems, vol. 155, p. 109580, Jan. 2024. [Baidu Scholar] 

10

S. Harasis, “Controllable transient power sharing of inverter-based droop controlled microgrid,” International Journal of Electrical Power & Energy Systems, vol. 155, p. 109565, Jan. 2024. [Baidu Scholar] 

11

Z. Zhao, J. Xie, S. Gong et al., “Modeling, oscillation analysis and distributed stabilization control of autonomous PV-based microgrids,” CSEE Journal of Power and Energy Systems, vol. 9, no. 3, pp. 912-936, May 2022. [Baidu Scholar] 

12

S. Jiang, Y. Zhu, and G. Konstantinou, “Settling-angle-based stability analysis for multiple current-controlled converters,” IEEE Transactions on Power Electronics, vol. 37, no. 11, pp. 12 992-12 997, Nov. 2022. [Baidu Scholar] 

13

X. Gao, D. Zhou, A. Anvari-Moghaddam et al., “Stability analysis of grid-following and grid-forming converters based on state-space modelling,” IEEE Transactions on Industry Applications, vol. 60, no. 3, pp. 4910-4920, May 2024. [Baidu Scholar] 

14

I. Subotić and D. GroSS, “Universal dual-port grid-forming control: bridging the gap between grid-forming and grid-following control,” IEEE Transactions on Power Systems, vol. 39, no. 6, pp. 6861-6875, Nov. 2024. [Baidu Scholar] 

15

L. S. Araujo, J. M. S. Callegari, B. J. C. Filho et al., “Heterogeneous microgrids: centralized control strategy with distributed grid-forming converters,” International Journal of Electrical Power & Energy Systems, vol. 158, p. 109950, Jul. 2024. [Baidu Scholar] 

16

F. Sadeque, M. Gursoy, and B. Mirafzal, “Grid-forming inverters in a microgrid: maintaining power during an outage and restoring connection to the utility grid without communication,” IEEE Transactions on Industrial Electronics, vol. 71, no. 10, pp. 11796-11805, Oct. 2024. [Baidu Scholar] 

17

C. Shen, W. Gu, and X. Shen, “Grid-forming control for solid oxide fuel cells in an islanded microgrid using maximum power point estimation method,” IEEE Transactions on Sustainable Energy, vol. 15, no. 3, pp. 1703-1714, Feb. 2024. [Baidu Scholar] 

18

M. S. Alvarez-Alvarado, C. Apolo-Tinoco, M. J. Ramirez-Prado et al., “Cyber-physical power systems: a comprehensive review about technologies drivers, standards, and future perspectives,” Computers and Electrical Engineering, vol. 116, p. 109149, May 2024. [Baidu Scholar] 

19

J. Hou, C. Hu, S. Lei et al., “Cyber resilience of power electronics-enabled power systems: a review,” Renewable and Sustainable Energy Reviews, vol. 189, p. 114036, Jan. 2024. [Baidu Scholar] 

20

Z. Zhao, Z. Zhang, Y. Wang et al., “Decentralized grid-forming control strategy for PV-based DC microgrids using finite control set model predictive control,” IEEE Transactions on Smart Grid, vol. 15, no. 6, pp. 5269-5283, Nov. 2024. [Baidu Scholar] 

21

H. Karimi, M. T. Beheshti, A. Ramezani et al., “Intelligent control of islanded AC microgrids based on adaptive neuro-fuzzy inference system,” International Journal of Electrical Power & Energy Systems, vol. 133, p. 107161, Jun. 2021. [Baidu Scholar] 

22

A. Oshnoei, M. Azzouz, A. S. Awad et al., “Adaptive damping control to enhance small-signal stability of DC microgrids,” IEEE Journal of Emerging and Selected Topics in Power Electronics, vol. 11, no. 3, pp. 2963-2978, Jun. 2023. [Baidu Scholar] 

23

N. Ghasemi, M. Ghanbari, and R. Ebrahimi, “Intelligent and optimal energy management strategy to control the micro-grid voltage and frequency by considering the load dynamics and transient stability,” International Journal of Electrical Power & Energy Systems, vol. 145, p. 108618, Feb. 2023. [Baidu Scholar] 

24

K. Tan, F. Lin, and C.-M. Shih et al., “Intelligent control of microgrid with virtual inertia using recurrent probabilistic wavelet fuzzy neural network,” IEEE Transactions on Power Electronics, vol. 35, no. 7, pp. 7451-7464, Jul. 2020. [Baidu Scholar] 

25

M. S. O. Yeganeh, A. Oshnoei, N. Mijatovic et al., “Intelligent secondary control of islanded AC microgrids: a brain emotional learning-based approach,” IEEE Transactions on Industrial Electronics, vol. 70, no. 7, pp. 6711-6723, Jul. 2023. [Baidu Scholar] 

26

C. Mu, Y. Zhang, H. Jia et al., “Energy-storage-based intelligent frequency control of microgrid with stochastic model uncertainties,” IEEE Transactions on Smart Grid, vol. 11, no. 2, pp. 1748-1758, Mar. 2020. [Baidu Scholar] 

27

F. Milano and M. Anghel, “Impact of time delays on power system stability,” IEEE Transactions on Circuits and Systems I: Regular Papers, vol. 59, no. 4, pp. 889-900, Apr. 2012. [Baidu Scholar] 

28

T. Surinkaew and I. Ngamroo, “Inter-area oscillation damping control design considering impact of variable latencies,” IEEE Transactions on Power Systems, vol. 34, no. 1, pp. 481-493, Jan. 2019. [Baidu Scholar]