Next Article in Journal
Understanding Reference-Dependent Behaviors in Determining Electricity Consumption of Korean Households: Empirical Evidence and Policy Implications
Next Article in Special Issue
A Wind Power Forecasting Method Based on Lightweight Representation Learning and Multivariate Feature Mixing
Previous Article in Journal
Reviews of Photovoltaic and Energy Storage Systems in Buildings for Sustainable Power Generation and Utilization from Perspectives of System Integration and Optimization
Previous Article in Special Issue
ANN-Based Maximum Power Tracking for a Grid-Synchronized Wind Turbine-Driven Doubly Fed Induction Generator Fed by Matrix Converter
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Safety Assessment of Loop Closing in Active Distribution Networks Based on Probabilistic Power Flow

by
Wenchao Cai
1,2,
Yuan Gao
3,
Xiping Zhang
3,
Qin Si
1,
Jiaoxin Jia
2 and
Bingzhen Li
2,*
1
Inner Mongolia Electric Power Research Institute, Inner Mongolia Electric Power (Group) Co., Ltd., Hohhot 010020, China
2
Hebei Key Laboratory of Distributed Energy Storage and Microgrid, North China Electric Power University, Baoding 071003, China
3
Ordos Power Supply Branch, Inner Mongolia Electric Power (Group) Co., Ltd., Ordos 017004, China
*
Author to whom correspondence should be addressed.
Energies 2025, 18(11), 2685; https://doi.org/10.3390/en18112685
Submission received: 4 April 2025 / Revised: 14 May 2025 / Accepted: 19 May 2025 / Published: 22 May 2025
(This article belongs to the Special Issue Trends and Challenges in Power System Stability and Control)

Abstract

To investigate the security issues of loop-closing operations in medium–low-voltage distribution networks under the influence of stochastic fluctuations from distributed generators (DGs) and loads, probabilistic power flow is introduced for analyzing loop-closing currents in active distribution networks. A novel method combining Latin Hypercube Sampling (LHS) and the Gram–Charlier (GC) series, termed the LHS-GC method, is proposed to calculate the probability distribution of loop-closing currents. By modeling DGs and loads as random variables, their cumulants are efficiently obtained through LHS. Based on a linearized formulation of loop-closing current equations, the cumulants of loop-closing currents are calculated, ultimately reconstructing the probability distribution function of loop-closing currents in active distribution networks. Subsequently, a security assessment framework for loop-closing operations is established using the probability distribution of loop-closing currents. This framework provides a quantitative evaluation from two dimensions: preliminary loop-closing success rate and the severity of current limit violations, offering data-driven decision support for loop-closing operations. Taking the IEEE 34-node distribution network as an example for feeder loop-closing current assessment, the proposed LHS-GC method achieves results with less than 4% deviation from simulation values in terms of cumulative probability distribution of loop-closing currents and safety assessment metrics. Under a sampling scale of 500 points, the computational time is 0.76 s, demonstrating its efficiency and reliability. These outcomes provide actionable references for decision-making support in loop-closing operations of active distribution networks.

1. Introduction

In recent years, ensuring power-supply reliability has emerged as a critical user demand, driven by economic development. As the power-supply network on the user side, the distribution network in China typically adopts the “closed-loop design, open-loop operation” mode, with a radial network structure [1,2]. When the distribution network lines need to handle faults or undergo maintenance, ring switching can achieve load transfer without interrupting the power supply [3,4]. However, the “connect-first, disconnect-later” strategy modifies the distribution network’s topology and power flow. During the loop connection process, steady-state circulating currents and inrush currents may occur, leading to protection malfunctions and line overloads, which in turn threaten the safe and stable operation of the power system [5,6]. Therefore, a method is needed to safely assess the current generated during the loop closure.
In the field of looped connections in classical distribution networks, relevant research has formed a relatively comprehensive theoretical system. In terms of equivalent models for looped connections in distribution networks, reference [7] established a 10 kV distribution network equivalent model considering the equivalent of the main network, feeder load, and transformer. Reference [8] studied the calculation methods for loop current using the superposition theorem and Thevenin’s theorem, and analyzed the key factors affecting the magnitude of the current. Reference [9] constructed a typical equivalent network for loops, and designed three methods for calculating loop current in medium- and low-voltage distribution networks, taking into account the complexity of calculation and accuracy.
However, with the widespread adoption of renewable energy, photovoltaic (PV) power generation systems are playing an increasingly critical role in distribution networks. The inherent randomness and volatility of their output power significantly complicate the calculation of loop-closing currents in distribution networks [10,11]. Currently, theoretical research on loop-closing current assessment techniques remains relatively limited, with many studies focusing on power flow optimization and computation in active distribution networks. For instance, reference [12] proposed a nonparametric quasi-Monte Carlo (MC) method based on uniform experimental design for efficient probabilistic power flow calculations. This approach introduces a hybrid discrepancy metric to enhance result accuracy while reducing computational burden. Reference [13] developed a high-order Markov chain-based modeling framework for PV output and load characterization in probabilistic power flow, improving computational efficiency and accuracy through spatio-temporally correlated scenarios generated via joint probability distributions and inverse transform strategies. These studies reveal the potential of integrating probabilistic power flow methodologies into loop-closing current assessment, offering valuable insights for decision support in loop-closing operations. The calculation methods for probabilistic power flow can be categorized into three main classes: the simulation method, analytical method, and approximate method [14]. The simulation method [15,16], represented by the MC method, simulates various uncertain factors through large-scale random sampling. When the sampling is sufficient, the accuracy is high and the applicability is broad, but the computational efficiency is low in complex systems. The approximation method [17,18] directly describes the probabilistic statistical properties of the output random variables, but the computational load increases significantly when solving models with a large number of nodes. The analysis method [19,20] linearizes uncertain variables to obtain the probability distribution of the output random variables, demonstrating excellent performance in computational efficiency.
The cumulant method (CM) is one of the mainstream analytical methods for probabilistic power flow. However, in medium- and low-voltage distribution networks, the probability distributions of many input variables are often unknown. Traditional CM inevitably introduces computational errors due to its reliance on approximated probability distribution functions [21]. To address this limitation, reference [22] improved CM by incorporating maximum entropy-based probability density function approximation, significantly enhancing its accuracy. Building on these advancements, this study focuses on refining the application of CM in loop-closing current assessment. By integrating simulation and analytical approaches, we propose a hybrid method for calculating the probability distribution of loop-closing currents. This approach reduces cumulant calculation errors while maintaining computational efficiency. The technical novelty of the method is summarized as follows:
(1)
The Integration of LHS and GC Series: The combination of LHS and GC series for loop-closing current probability distribution calculation in active distribution networks retains the flexibility of simulation methods and the efficiency of analytical approaches.
(2)
The Direct Processing of Discrete Sampling Points: LHS is employed to handle discrete sampling points with minimal dependence on predefined input variable distributions. Correlations between sampling points are mitigated through ranking strategies, yielding more accurate cumulants for input variables.
(3)
Linear Relationship Between Loop Currents and Nodal Power Injections: By establishing a linear relationship between loop-closing currents and nodal power injections, convolution operations are transformed into algebraic cumulant calculations, significantly enhancing computational efficiency.
(4)
Two-Dimensional Safety Assessment Framework: A quantitative framework incorporating preliminary loop-closing success rate and the severity of current limit violations is developed, providing comprehensive and precise quantitative support for operational decision making.
(5)
The proposed method is fundamentally based on linearized power flow equations, making it inherently applicable to distribution networks of diverse scales and configurations. Its effectiveness has been validated through a case study on the IEEE 34-node system, demonstrating robust potential for real-world engineering applications.

2. Calculation of Loop Current in Active Distribution Networks

2.1. Calculation of Steady-State Current in a Closed Loop

Taking the typical active distribution network combined-loop operation shown in Figure 1 as an example, the calculation theory and method of combined-loop steady-state current are studied.
In the figure, the distribution network operates in an open loop. Two main transformers T1 and T2 are, respectively, configured at 110 kV busbar; the distributed power supplies DG1 and DG2 are, respectively, connected at 10 kV busbar; the outlet of the feeder is controlled by circuit breakers QF1 and QF2, respectively; and the feeder is electrically connected through connection switch QF3. In terms of load distribution, Saj and Sbj, respectively, represent the power load of the two feeders, and Saf and Sbf, respectively, represent the power load of the two sides of the connection switch.
The steady-state current analysis of the ring network is shown in Figure 2. Figure 2a is the network after the closure of the interconnection switch QF3, where I ˙ a and I ˙ b are the steady-state currents of the two feeders after the ring is closed, and I ˙ c is the circulating current generated by the closure of the ring; Figure 2b is the network before the ring is closed, where I ˙ a and I ˙ b are the initial currents of the two feeders before the ring is closed; Figure 2c is the equivalent circuit of the closed-ring network, simplified according to Thevenin’s theorem, where the equivalent voltage source U ˙ oc represents the voltage phasor difference across the interconnection switch before the ring closure, and the equivalent impedance Zeq is the sum of the impedances of the ring network.
According to the reference [7], the mesh current in a loop can be represented as:
I ˙ c = U ˙ oc 3 Z eq = U ˙ a U ˙ b 3 Z eq
According to the superposition theorem, when the tie switch QF3 is closed to form a looped network, the steady-state current in the feeders can be decomposed into two independent components: the initial current components I ˙ a and I ˙ b from the original open-loop operation of each feeder, and the circulating current component I ˙ c generated by the potential difference at the loop-closing point. By vectorially superimposing these two components, the steady-state current phasors of the feeders under looped operation can be determined as follows:
I ˙ a = I ˙ a + I ˙ c I ˙ b = I ˙ b I ˙ c

2.2. Calculation of Loop Impulse Current

The combined-loop operation may cause a large transient current shock, and the influence of the combined-loop shock current on the safety of the power grid needs to be considered. According to reference [10], the loop impulse current instantaneous values iM and the maximum rms values IM are:
i M = 2 1 + e 0.01 T a I c = 2 k m I c
I M = I c 1 + 2 ( k m 1 ) 2
where Ic is the rms value of the circulating current; Ta is the attenuation time constant of the resultant loop shock current; and Ta = Leq/Req.
As can be seen from Equations (1)–(4), the amplitude of the combined-loop steady-state current and the shock current is determined by the voltage difference on both sides of the break of the link switch, and decreases as the equivalent impedance of the combined-loop network increases. According to Equation (4), the effective values of the maximum shock currents of the feeders at both ends of the combined ring are calculated as follows:
I Ma = I a + I c 1 + 2 ( k m 1 ) 2 I Mb = I b + I c 1 + 2 ( k m 1 ) 2
Through power flow calculations at the reference node, the voltage magnitudes Ui and phase angles δi (i = 1, 2, …, n) of all nodes in the pre-loop-closing network can be obtained. Let U = [ U 1 , δ 1 , , U a , δ a , U b , δ b , , U n , δ n ] denote the state variables of the closed-loop network, and I = [ I a , I b , I Ma , I Mb ] represent the effective values of the steady-state currents and inrush currents on both sides of the feeders after loop closing. The loop-closing current equations can then be expressed as:
I = G ( U )
In engineering practice, the inrush coefficient km is typically selected within the range of 1.8–1.9 [23,24]. For conservative design purposes, km = 1.9 is adopted. Substituting this value into Equation (5), the expanded form of Equation (6) can be derived as:
I ˙ a = I ˙ a + I ˙ c = g 1 ( U ) I ˙ b = I ˙ a I ˙ c = g 2 ( U ) I Ma = I a + 1.62 I c = g 3 ( U ) I Mb = I b + 1.62 I c = g 4 ( U )

3. Establishment of System Probability Model

3.1. Power Probability Model of Distributed Power Supply

Distributed household photovoltaic has developed rapidly, and photovoltaic power generation systems will become an important form of renewable energy power generation in low- and medium-voltage distribution networks. The increasing penetration of distributed PV systems, characterized by their randomness and output power variability, has significantly intensified the challenges in performing power flow calculations prior to loop-closing operations. Therefore, not only should the load fluctuation be considered in the calculation of combined loop current, but a probability model for photovoltaic power generation should also be constructed. The illumination intensity r can be approximately regarded as a Beta distribution over a finite period of time, and its probability density function is [13]:
f ( r ) = Γ ( α + β ) Γ ( α ) Γ ( β ) ( r r max ) α 1 ( 1 r r max ) β 1
where rmax is the maximum illumination intensity, α and β are the two shape parameters of the Beta distribution, and Γ is the Gamma function.
In practice, the shape parameters of the Beta distribution can be determined by statistical analysis, and two shape parameters in the Beta distribution are calculated from the expectations μ and variance σ of the given time-series data [11]:
α = μ μ ( 1 μ ) σ 2 1 β = ( 1 μ ) μ ( 1 μ ) σ 2 1
The power probability characteristic of the photovoltaic power supply satisfies the following requirements:
P V ( r ) = r A η
where PV is the total output power; A is the total area of the photovoltaic panel, and η is the photoelectric conversion efficiency.
It can be seen from Equation (10) that the output power of the photovoltaic system has a positive correlation with the illumination. In combination with Equation (8), the probability density function of the light intensity and the photovoltaic output force can be derived:
f ( P V ) = Γ ( α + β ) Γ ( α ) Γ ( β ) ( P V max ) α 1 ( 1 P V max ) β 1
where PVmax is the maximum power of photovoltaic power generation, which meets PVmax = rmax.
In practice, through the regulation of the control system, photovoltaic power generation units can achieve stable control of the power factor, so they can be modeled as PQ nodes in the load flow calculation of the distribution network. Thus, under the known power factor angle of φ, the reactive output power of the photovoltaic array can be obtained as QV [14].
Q V = P V tan φ
After applying PQ node equivalent processing to the photovoltaic power source, it can be regarded as a load to derive the Thevenin equivalent impedance expression ZVeq:
Z veq = P V U V 2 P V 2 + Q V 2 j Q V U V 2 P V 2 + Q V 2

3.2. Load Power Probability Model

Using a normal distribution to approximate the uncertainty of reactive power load, assuming the corresponding parameters for active power and reactive power are μP, σP and μQ, σQ, the probability density functions of active power PL and reactive power QL can be expressed as [20]:
f ( P L ) = 1 2 π σ P exp P L μ P 2 σ P 2 f ( Q L ) = 1 2 π σ Q exp Q L μ Q 2 σ Q 2

4. Solution of Cumulants for Discrete Input Variables

4.1. Latin Hypercube Sampling

When calculating the joint probability of the load considering the randomness of photovoltaic power generation, it is necessary to establish a method based on stochastic power flow calculations. LHS can effectively reflect the overall distribution of random variables through sampling values. This method divides the range of each variable equally and samples uniformly within each interval, thereby improving the accuracy of the simulation. Therefore, compared to MC sampling, it can obtain more accurate sample mean and variance values with fewer sampling numbers.
LHS mainly has two steps of sampling and arrangement. The sampling step ensures that the sample distribution area can be sufficiently collected, assuming X1, X2, …, and Xp are p input random variables, where the cumulative probability distribution function of Xp is:
Y p = F p ( X p )
The sampling scale is N, the value-taking interval (0,1) of Yp is divided into N intervals, and the length of each interval is 1/N. By setting the sample point Yp at the center of each interval, the i-th sample value of the corresponding input random variable Xp can be obtained from the inverse function:
X pi = F p 1 ( i 0.5 N )
The sampling matrix of each input random variable is 1 × N, and all p input random variables are processed to form an initial sampling matrix of p × N, so that the sampling step is finished, and the next order is carried out. The sorting step randomly arranges the position of each element in each row, reduces the correlation between each random variable of the initial sampling matrix, and obtains the final sampling matrix. Calculations of the order of origin moments α p ( v ) and cumulants γ p ( v ) of the input random variable Xp are then performed based on the final sampling matrix.

4.2. Latin Hypercube Sampling for Discrete Data

The LHS requires the cumulative probability distribution function Yp = Fp(Xp) of the random variable Xp as input. However, in engineering applications, many of the operating parameters and historical data of variables are often discrete, making it impossible to directly obtain the cumulative probability distribution function of these discrete data. Therefore, a method is needed to determine semi-variates using LHS based on discrete data.
For a subsample set of the random variable Xp, when it contains N discrete observations, it can be represented as {xp1, xp2, xp3, …, xpN}. Sort the subsample set of each input random variable in ascending order and calculate the empirical distribution function of Xp:
H p ( X p ) = 0 ,   x p < x p 1 k n , x pk x p x p k + 1 , k = 1 , 2 , , N 1 1 , x p x p N + 1
According to the empirical distribution function Hp(Xp), a sample set of Xp  { x p 1 , x p 2 , x p 3 , , x pN } is obtained using LHS, and the moments of different orders of Xp are calculated, respectively:
α p ( v ) = 1 N i = 1 N x pk ( v ) , k = 1 , 2 , , n
Then, based on the relationship between cumulants and the moment about the origin, we obtain the cumulants of each order for Xp.

5. Probabilistic Calculation of Loop-Closing Currents and Line Losses

5.1. Linearization of the Loop Current Equation

Let W = [ P 1 , Q 1 , , P n , Q n ] be the injection power of each node in the ring network, and the matrix form of the node power equation is:
W = F ( U )
The state variable U of the closed-loop network system can be decomposed into the expected value U0 and the disturbance amount ΔU, which satisfies the relationship W0 = F(U0) with the expected value of the node injection power W0. Expanding Equation (19) according to the Taylor series and ignoring higher-order terms yields:
Δ W = J 0 Δ U
where J0 is the Jacobian matrix, defined as:
J 0 = F ( U ) U U = U 0
Similarly, the loop current equation can also be subjected to similar linearization processing, resulting in:
Δ I = G 0 Δ U = G 0 J 0 1 Δ W
where G0 is the coefficient matrix, defined as:
G 0 = G ( U ) U U = U 0
Equation (22) shows that the combined-loop current can be expressed as a linear combination of the variable ΔW, providing a theoretical basis for obtaining the probability distribution of the combined-loop current based on cumulants.

5.2. Loop-closing Current Probability Density Function

Apply the LHS method to process discrete data to obtain the cumulants Δ W P ( k ) and Δ W L ( k ) for photovoltaic power generation and load. Due to the additivity of cumulants, by summing the cumulants of each node’s load power Δ W L ( k ) and the cumulant of the photovoltaic output Δ W P ( k ) , we obtain the cumulative amount of node injection power Δ W ( k ) .
Δ W ( k ) = Δ W L ( k ) + Δ W P ( k )
Refer to the linear properties of the cumulant and combine with Equation (22) to obtain the cumulant of the combined ring current Δ I ( k ) .
Δ I ( k ) = G 0 J 0 1 k Δ W ( k )
After obtaining the cumulants of each order of the circular current, the cumulative probability distribution function F(x) can be constructed using the Gram–Charlier series expansion method, as follows [22]:
F ( x ) = Ψ ( x ) + c 1 1 ! Ψ ( x ) + c 2 2 ! Ψ ( x ) +
where Ψ ( x ) is the cumulative distribution function of the standard normal distribution; the coefficients ci can be obtained using the semi-invariants of each order.

5.3. Probability Function of Power Loss in Distribution Networks

Power loss is a critical comprehensive indicator for evaluating the economic and technical performance of power grid operation. The integration of DGs into distribution networks significantly complicates power loss calculations. To reduce computational complexity, power loss can be defined as the difference between injected power and output power. The cumulants of different orders for line losses in the distribution network can be calculated as follows:
Δ P loss ( k ) = P s ( k ) + i = 1 g P i _ DG ( k ) j = 1 h P j _ L ( k )
where Δ P loss ( k ) is the k-th order cumulant of distribution network power loss; P s ( k ) is the k-th order cumulant of the output power from the main power source; i = 1 g P i _ DG ( k ) is the sum of k-th order cumulants from g distributed generators; j = 1 h P j _ L ( k ) is the sum of k-th order cumulants from h loads.

6. Safety Assessment Indicators for Distribution Network Ring Connection

6.1. Initial Success Rate of Loop Closure

The conditions for the safe loop operation of the distribution network are (1) the steady-state current of each feeder after looping does not exceed its maximum carrying capacity; (2) the impact current of each feeder during the closing loop is below the fast trip action threshold of the protective device.
Let F1(x), F2(x), F3(x), and F4(x) be the cumulative distribution functions of the combined-loop currents I a , I b , IMa, and IMb, respectively. The current carrying limits of the two feeder lines are Imax.a and Imax.b, and the settings for the current instantaneous protection are IsetI;.a and IsetI.b. Then, the probability of the combined-loop current exceeding the safety threshold can be expressed as [25]:
P 1 = P ( I ˙ a I max . a ) = 1 F 1 ( I max . a ) P 2 = P ( I ˙ a I max . b ) = 1 F 2 ( I max . b ) P 3 = P ( I ˙ Ma I set . a ) = 1 F 3 ( I set . a ) P 4 = P ( I ˙ Mb I set . b ) = 1 F 4 ( I set . b )
The preliminary success rate of grid connection is quantitatively described to assess the impact of distributed power sources on exceeding the grid connection current limit, providing data support for grid connection decisions. The expression for the preliminary success rate P is as follows:
P = ( 1 P 1 ) ( 1 P 2 ) ( 1 P 3 ) ( 1 P 4 )
If P is greater than 95%, this operation is defined as a preliminary successful closure.

6.2. Excessive Ring Current Limit

Although the initial success rate of the loop closure can indicate a high probability of meeting the requirements for loop-closure current, certain loop-closure currents with lower probabilities may cause serious violation issues. Therefore, it is necessary to define the maximum violation rate of the loop-closure current ηM and the average violation rate ηA as follows [25]:
η M = I F I Lmax 1 × 100 %
η A = I Lmax I F x f ( x ) d x I Lmax I F f ( x ) d x 1 × 100 %
where IF is the current value corresponding to the cumulative probability distribution of the line current at 99.9%, ILmax is the maximum current-carrying capacity of the line, and f(x) is the probability density function of the line current.
Since the power system has a certain overload capacity, it is considered that the safe operation of the distribution network requires that the ηM does not exceed 10% and ηA does not exceed 5%. If a preliminary loop closing is unsuccessful but the exceeding current rate meets the requirements, then the part of the current that does not meet the preliminary loop-closing success rate is determined to be the safe current. In summary, using preliminary loop-closing success rate, maximum exceeding rate, and average exceeding rate as quantitative standards for assessing the safety of loop-closing operations provides data references for the operators’ loop-closing decisions.
In summary, the process of the LHS-GC method based on LHS and Gram–Charlier series is shown in Figure 3. The arrows in Figure 3 indicate the next step.

7. Example Verification

This paper uses the IEEE 34-node distribution network system as an example to verify the proposed method. The standard system is simplified to a single voltage level by removing transformers and voltage regulators from the lines. Due to the text length constraints in the main body, the line impedance data of IEEE 34 nodes and the normal distribution data of loads are presented in Appendix A and Appendix B. Based on the simplified distribution network, photovoltaic power sources are configured using solar radiation intensity samples from a region in South China as input data. The system topology and an example of the tie line are shown in Figure 4. In Figure 4, the black solid line represents the existing distribution line, and the blue dashed line represents the tie line that has not been closed.
Based on the sample data of solar radiation intensity within a day, the shape parameters of the Beta distribution are calculated to be α = 0.679 and β = 1.778, with a maximum solar radiation intensity of 1.134 kW/m2, as shown in Figure 5. Take the reference capacity SB as 1 MVA and the reference voltage UB as 24.9 kV.
Based on variations in PV integration nodes, loop-closing line locations, and load standard deviations, two cases are defined as shown in Table 1. These cases are selected to analyze the safety of loop-closing operations and power losses.

7.1. Probabilistic Calculation of Loop-Closing Currents and Power Losses

This case study uses loop-closing currents obtained from MATLAB/SIMULINK simulations as a reference to compare the results of three methods: MC method, traditional cumulant method, and the proposed LHS-GC method, with a sampling size of 500 for both MC and LHS. The cumulative probability distributions of loop-closing currents for Case 1 and Case 2 are illustrated in Figure 6 and Figure 7, respectively.
As shown in Figure 6 and Figure 7, compared to the MC method and traditional CM, the proposed LHS-GC method achieves higher accuracy, with its cumulative probability distribution curves aligning more closely with simulation results. Due to limitations in sampling scale, the traditional CM demonstrates higher precision than the MC method under the current configuration.
Figure 8 illustrates the deviations of the three methods from simulation values at the 90% cumulative probability level. Taking Case 1 as an example, the calculation errors of the LHS-GC method for steady-state currents and inrush currents in the two feeders are 3.49%, 6.33%, 3.4%, and 3.11% lower than those of the MC method, and 1.72%, 3.1%, 3.09%, and 1.8% lower than those of the CM, respectively. Furthermore, the current calculation errors of the LHS-GC method are consistently below 3%, confirming its superior performance in approximating the cumulative probability distribution function of feeder loop-closing currents. These results validate the reliability of the LHS-GC method for loop-closing safety assessments.
Table 2 compares the computation times of the methods. The LHS-GC method requires shorter single-cycle time and total duration, achieving a 68% speed improvement compared to the MC method. Under identical sampling scales, it exhibits significant speed advantages. Therefore, compared to the traditional CM, the LHS-GC method achieves higher accuracy with only a minor sacrifice in computational efficiency, making it a practical tool for real-time operator decision making.
Under both operating cases, the power losses of the entire network before and after loop closing are shown in Figure 9. For Case 1, the power losses showed no significant variation between pre- and post-loop-closing operations. In contrast, for Case 2, the power losses decreased after loop closing compared to pre-loop-closing levels, contributing to the economic operation of the distribution network. The proposed LHS-GC method accurately captures the variations in power losses, exhibiting errors within 8% compared to simulation values at the 90% cumulative probability level. This demonstrates its capability to provide a reliable reference for assessing the economic and technical performance of post-loop-closing distribution networks.

7.2. Loop-Closing Safety Assessment

Safety assessments for loop-closing operations are conducted using the cumulative probability distribution curves of feeder loop-closing currents calculated by the LHS-GC method. Based on the 3-sigma principle of load normal distribution, the maximum allowable current-carrying capacity of branches is determined. The instantaneous current protection settings are derived from the maximum short-circuit current at the end of the protected line during a fault. The calculated results for the maximum allowable current-carrying capacity and protection settings under Case 1 and Case 2 are summarized in Table 3.
The safety assessment results of loop-closing currents are presented in Table 4 and Table 5. In Case 1, the proposed LHS-GC method exhibits errors of 0.69% and 1.17% in the maximum violation rates for feeders 26–28 and 32–34 compared to simulation values, respectively. The errors in average exceedance rates are 0.36% and 0.95%, respectively. In Case 2, the errors of maximum violation rates and average exceedance rates calculated by LHS-GC for feeders 4–6 and 11–13 are 1.53%, 0.59%, and 0.86%, 0.67%, respectively, all within 2%. Additionally, the preliminary loop-closing success rates calculated by LHS-GC for both cases closely align with simulation values, exhibiting errors of 1.26% and 1.81%.
In summary, an analysis of these two cases demonstrates that the LHS-GC method can effectively assess the safety of loop-closing operations in active distribution networks, even with limited samples. The entire process—from calculating the cumulative probability distribution of loop-closing currents to deriving safety assessment metrics—requires only 0.76 s, meeting the timeliness requirements for engineering applications.

8. Conclusions

To address the challenges in the safety assessment of loop-closing operations for active distribution networks, this paper proposes an LHS-GC-based framework for probabilistic loop-closing current analysis and safety evaluation. Validation on the IEEE 34-node system yields the following conclusions:
(1)
Compared to the MC method, the LHS-GC method significantly improves both accuracy and computational speed under identical sampling scales. Compared to the traditional CM, it achieves higher accuracy while maintaining comparable speed. With a sampling size of 500, the method completes calculations in 0.76 s with probability distribution errors below 4%.
(2)
Under the two-dimensional safety assessment framework (preliminary success rate and severity of current violations), the LHS-GC method achieves errors within 3% for all metrics. The total computation time of 0.76 s fulfills real-time operational demands for active distribution networks.
(3)
The LHS-GC method accurately captures the probabilistic distribution of loop-closing currents and safety metrics across diverse scenarios, including varying PV integration nodes, loop-closing line locations, and load standard deviations. Rooted in linearized power flow equations, the method is inherently adaptable and holds potential for application in distribution networks with various configurations.

Author Contributions

Conceptualization, W.C.; methodology, Y.G.; formal analysis, X.Z.; data curation, Q.S.; writing—original draft preparation, B.L.; writing—review and editing, W.C.; funding acquisition, J.J. All authors have read and agreed to the published version of the manuscript.

Funding

Science and technology project funding of Inner Mongolia Power Group: 2024-4-47; Supported by the National Natural Science Foundation of China: 52207102.

Data Availability Statement

Data are unavailable due to privacy restrictions.

Conflicts of Interest

Authors Wenchao Cai, Qin Si, Yuan Gao and Xiping Zhang were employed by the company Inner Mongolia Electric Power (Group) Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

Abbreviations

The following abbreviations are used in this manuscript:
PVPhotovoltaic
MCMonte Carlo method
CMCumulant method
LHSLatin hypercube sampling
GCGram–Charlier
DGDistributed generation
QFCircuit breaker
Saj, Sbjthe power load of the two feeders
Saf, Sbfthe power load of the two sides of the connection switch
I ˙ a , I ˙ b the initial currents of the two feeders
I ˙ a , I ˙ b Current of the feeders after closing the loop
I ˙ c the circulating current
U ˙ oc Voltage difference across the contact switch
ZeqEquivalent impedance of the loop
iMInstantaneous values of the loop impulse current
IMMaximum effective value of loop impulse current
Icthe rms value of the circulating current
TaDecay time constant of the loop impulse current
ReqEquivalent reactance of the loop
LeqEquivalent resistance of the loop
IMa, IMbMaximum impulse current effective value of the two feeders
kmImpact coefficient
rIllumination intensity
α, βShape parameters for beta distribution
Γ Gamma function
μExpectations for data in a given time series
σVariance of data for a given time series
PV, QVPhotovoltaic array output power
ATotal area of photovoltaic panels
ηPhotoelectric conversion efficiency
PVmaxMaximum power for photovoltaic power generation
XpInput random variables
YpCumulative distribution function of Xp
α p ( v ) Moment matrix of each order
γ p ( v ) Partial invariants of each order
Hp(Xp)Empirical distribution function
WInjected power of each node in the ring network
J0Jacobian matrix
Δ W P ( k ) Cumulants of photovoltaic power at various stages
Δ W L ( k ) Cumulants of load at each level
Δ W ( k ) Cumulants of the node injection power of various orders
Δ I ( k ) Cumulants of the loop current
Ψ ( x ) the cumulative distribution function of the standard normal distribution
Imax.a, Imax.bthe current carrying limits of the two feeder lines
IsetI.a, IsetI.bthe settings for the current instantaneous protection
PThe probability of the closed-loop current exceeding the safety threshold
ηMthe maximum violation rate of loop-closure current
ηAthe average violation rate
SBBenchmark capacity of distribution network
UBReference voltage of the distribution network

Appendix A

Table A1. Original data table of IEEE 34-node line.
Table A1. Original data table of IEEE 34-node line.
Starting NodeLeaf NodeResistor (p.u.)Reactance (p.u.)Electricity (p.u.)
122.030825 × 10−38.953083 × 10−44.500000E × 10−7
231.361754 × 10−36.003423 × 10−43.000000 × 10−7
342.536956 × 10−21.118441 × 10−25.500000 × 10−6
454.596905 × 10−32.026589 × 10−31.000000 × 10−6
462.951780 × 10−21.301320 × 10−26.500000 × 10−6
672.340171 × 10−21.031687 × 10−25.000000 × 10−6
787.871413 × 10−43.470187 × 10−40.000000
892.440138 × 10−41.075758 × 10−40.000000
9101.346012 × 10−35.934020 × 10−42.500000 × 10−7
9118.036713 × 10−33.543061 × 10−31.500000 × 10−6
10123.790085 × 10−21.670895 × 10−28.000000 × 10−6
11136.611987 × 10−42.914957 × 10−41.000000 × 10−7
11142.385038 × 10−31.051467 × 10−35.000000 × 10−8
12151.081532 × 10−24.768037 × 10−32.350000 × 10−5
13161.608917 × 10−27.093062 × 10−33.550000 × 10−5
14174.093135 × 10−41.804497 × 10−40.000000
17182.899042 × 10−21.278070 × 10−26.000000 × 10−6
17191.836401 × 10−28.095946 × 10−34.045000 × 10−5
18207.871413 × 10−43.470187 × 10−40.000000
20213.064467 × 10−36.580539 × 10−33.000000 × 10−6
20223.856993 × 10−31.700392 × 10−35.000000 × 10−7
21238.312212 × 10−33.664518 × 10−31.500000E × 10−6
22244.589034 × 10−32.023119 × 10−31.000000 × 10−6
22251.275169 × 10−35.621703 × 10−42.500000 × 10−7
24261.590025 × 10−37.009778 × 10−45.000000 × 10−8
24272.203996 × 10−49.716525 × 10−55.000000 × 10−9
26282.109539 × 10−39.300102 × 10−45.000000 × 10−7
27291.062641 × 10−34.684753 × 10−42.000000 × 10−7
28302.203996 × 10−49.716525 × 10−55.000000 × 10−9
28316.769415 × 10−42.984361 × 10−41.500000 × 10−7
29322.865194 × 10−31.263148 × 10−35.000000 × 10−7
30332.558431 × 10−31.684948 × 10−35.000000 × 10−7
32344.171849 × 10−41.839199 × 10−45.000000 × 10−8

Appendix B

Table A2. Load data table of IEEE 34 nodes (for example, a standard deviation of 30%).
Table A2. Load data table of IEEE 34 nodes (for example, a standard deviation of 30%).
NodeActive Power LoadReactive Power Load
ExpectationVarianceExpectationVariance
21.910000 × 10−25.730000 × 10−39.870000 × 10−32.960000 × 10−3
45.290000 × 10−31.587000 × 10−32.740000 × 10−38.220000 × 10−4
91.300000 × 10−43.900000 × 10−57.000000 × 10−52.100000 × 10−5
104.596905 × 10−33.390000 × 10−35.840000 × 10−31.752000 × 10−3
111.490000 × 10−24.470000 × 10−37.710000 × 10−32.313000 × 10−3
121.184000 × 10−23.552000 × 10−32.336000 × 10−27.008000 × 10−3
132.060000 × 10−36.180000 × 10−41.070000 × 10−33.210000 × 10−4
171.240000 × 10−33.720000 × 10−46.400000 × 10−41.920000 × 10−4
204.370000 × 10−31.311000 × 10−32.260000 × 10−36.780000 × 10−4
221.000000 × 10−23.000000 × 10−35.170000 × 10−31.551000 × 10−3
232.700000 × 10−28.100000 × 10−32.162000 × 10−26.486000 × 10−3
245.000000 × 10−21.500000 × 10−20.0000000.000000
264.657000 × 10−21.397100 × 10−22.972000 × 10−28.916000 × 10−3
273.040000 × 10−39.120000 × 10−41.570000 × 10−34.710000 × 10−4
281.310000 × 10−23.930000 × 10−36.770000 × 10−32.031000 × 10−3
291.490500 × 10−14.471500 × 10−21.490000 × 10−24.470000 × 10−3
309.200000 × 10−32.760000 × 10−34.760000 × 10−31.428000 × 10−3
318.859999 × 10−32.658000 × 10−37.090000 × 10−32.127000 × 10−3
327.540000 × 10−32.262000 × 10−33.900000 × 10−31.170000 × 10−3
341.945000 × 10−25.835000 × 10−31.344300 × 10−14.032900 × 10−2

References

  1. Guan, Z.; Tang, P.; Mao, C.; Wang, D.; Wang, L.; Liu, W.; Du, M.; Li, J.; Wang, X. Control Strategy and Implementation of Seamless Closed-Loop Load Transfer Mobile Prototype for 400 V Distribution Network. IEEE Access. 2024, 12, 12279–12294. [Google Scholar] [CrossRef]
  2. Li, J.; Wang, X.; Xie, C.; Han, J.; Mao, C.; Liu, X.; Wang, D.; Zhang, Y.; Guan, Z.; Zhang, X. A novel method for seamless closed-loop load transfer in radial distribution networks. IEEE Trans. Ind. Appl. 2023, 59, 2254–2265. [Google Scholar] [CrossRef]
  3. Liao, H.; Liu, D.; Huang, Y.; Zhang, Y. Load transfer capability analysis considering interconnection of distributed generation and energy storage system. Int. Trans. Elect. Energy Syst. 2014, 24, 166–177. [Google Scholar] [CrossRef]
  4. Gan, G.; Wang, Z.; Zhou, Y.; Li, L.; Li, X. Reliability and economy based closed-loop operation modes for medium voltage distribution networks. Autom. Electr. Power Syst. 2015, 39, 144–150. [Google Scholar]
  5. Fernandes, D.A.; Costa, F.F.; Martins, J.R.S.; Lock, A.S.; da Silva, E.R.C.; Vitorino, M.A. Sensitive load voltage compensation performed by a suitable control method. IEEE Trans. Ind. Appl. 2017, 53, 4877–4885. [Google Scholar] [CrossRef]
  6. Liao, S.; Xu, J.; Sun, Y.; Gao, W.; Ma, X.; Zhou, M.; Qu, Y.; Li, X.; Gu, J.; Dong, J. Load-damping characteristic control method in an isolated power system with industrial voltage-sensitive load. IEEE Trans. Power Syst. 2016, 31, 1118–1128. [Google Scholar] [CrossRef]
  7. Li, J.; Wang, Z.; Ren, J.; Xu, M.; Tian, F.; Zou, H. Feasibility analysis of loop closing operation in a distribution network based on the main network and feeder load equivalent model. Power Syst. Prot. Control 2023, 51, 107–116. [Google Scholar]
  8. Shi, A.; Yang, Z.; Pang, J.; Han, S. Analysis of key factors affecting closing current of distribution network of the medium and low voltage. In Proceedings of the 2022 Annual Meeting of CSEE Study Committee of HVDC and Power Electronics (HVDC 2022), Guangzhou, China, 18-21 December 2022; pp. 374–380. [Google Scholar]
  9. Zhang, K.; Feng, Y.; Yu, Y.; Guo, Q. Study of calculation schemes of currents induced by closing-loop operations in medium-voltage distribution grids. In Proceedings of the 2019 IEEE Sustainable Power and Energy Conference (iSPEC), Beijing, China, 21–23 November 2019; pp. 2488–2493. [Google Scholar]
  10. Yang, B.; Wei, Y.; Zeng, Q.; Xu, J.; Zhang, F.; Huang, M. Research on system-friendly optical storage DC microgrid sending-out control strategy via MMC. In Proceedings of the 2021 11th International Conference on Power and Energy Systems (ICPES), Shanghai, China, 18–20 December 2021; pp. 501–505. [Google Scholar]
  11. Abbasi, A.R.; Mohammadi, M. Probabilistic load flow in distribution networks: An updated and comprehensive review with a new classification proposal. Electr. Power Syst. Res. 2023, 222, 109497. [Google Scholar] [CrossRef]
  12. Krishna, A.B.; Abhyankar, A.R. Uniform experimental design-based nonparametric quasi-monte carlo for efficient probabilistic power flow. IEEE Trans. Power Syst. 2023, 38, 2318–2332. [Google Scholar] [CrossRef]
  13. Liu, M.; Xiong, Y.; Li, Q.; Murad, M.A.A.; Zhong, W. Higher-order markov chain-based probabilistic power flow calculation method considering spatio-temporal correlations. Energies 2025, 18, 1058. [Google Scholar] [CrossRef]
  14. Prusty, B.R.; Jena, D. A critical review on probabilistic load flow studies in uncertainty constrained power systems with photovoltaic generation and a new approach. Renew. Sustain. Energy Rev. 2017, 69, 1286–1302. [Google Scholar] [CrossRef]
  15. Hajian, M.; Rosehart, W.D.; Zareipour, H. Probabilistic power flow by monte carlo simulation with latin supercube sampling. IEEE Trans. Power Syst. 2012, 28, 1550–1559. [Google Scholar] [CrossRef]
  16. Luo, G.; Chen, J.; Cai, D.; Shi, D.; Duan, X. Probabilistic assessment of available transfer capability considering spatial correlation in wind power integrated system. IET Gener. Transm. Distrib. 2013, 7, 1527–1535. [Google Scholar]
  17. Gupta, N.; Daratha, N. Probabilistic three-phase load flow for unbalanced electrical systems with wind farms. Int. J. Electr. Power Energy Syst. 2017, 87, 154–165. [Google Scholar] [CrossRef]
  18. Wang, X.; Gong, Y.; Jiang, C. Regional carbon emission management based on probabilistic power flow with correlated stochastic variables. IEEE Trans. Power Syst. 2015, 30, 1094–1103. [Google Scholar] [CrossRef]
  19. Schellenberg, A.; Rosehart, W.; Aguado, J. Cumulant-based probabilistic optimal power flow (P-OPF) with Gaussian and gamma distributions. IEEE Trans. Power Syst. 2005, 20, 773–781. [Google Scholar] [CrossRef]
  20. Liu, H.; Tang, C.; Han, J.; Li, T.; Li, J.; Zhang, K. Probabilistic load flow analysis of active distribution network adopting improved sequence operation methodology. IET Gener. Transm. Distrib. 2017, 11, 2147–2153. [Google Scholar] [CrossRef]
  21. Moreira, E.; Chagas, E.; Rodrigues, A.; da Silva, M.D.G. Linear probabilistic power flow for islanded microgrids. In Proceedings of the 2020 International Conference on Probabilistic Methods Applied to Power Systems (PMAPS), Liege, Belgium, 18–21 August 2020; pp. 1–6. [Google Scholar]
  22. Singh, V.; Moger, T.; Jena, D. Maximum entropy based probabilistic load flow for assessing input uncertainties and line outages in wind-integrated power systems. Electr. Power Syst. Res. 2025, 244, 111528. [Google Scholar] [CrossRef]
  23. Xiao, H.; Pei, W.; Wu, L.; Li, M.; Ma, T.; Hua, W. A novel deep learning based probabilistic power flow method for Multi-Microgrids distribution system with incomplete network information. Appl. Energy 2023, 335, 120716. [Google Scholar] [CrossRef]
  24. Kim, S.; Hur, J. A probabilistic modeling based on Monte Carlo simulation of wind powered EV charging stations for steady states security analysis. Energies 2020, 13, 5260. [Google Scholar] [CrossRef]
  25. Huang, Y.; Rong, N.; Chen, S.; Guo, Y. Active distribution network feeder loop closing current assessment based on DK2-PL. Electr. Power Eng. Technol. 2024, 43, 133–141. [Google Scholar]
Figure 1. Schematic diagram of active distribution network ring operation.
Figure 1. Schematic diagram of active distribution network ring operation.
Energies 18 02685 g001
Figure 2. Closed-loop steady-state current analysis diagram. (a) Network after ring closure; (b) network before closing the loop; (c) equivalence network for closed loops.
Figure 2. Closed-loop steady-state current analysis diagram. (a) Network after ring closure; (b) network before closing the loop; (c) equivalence network for closed loops.
Energies 18 02685 g002
Figure 3. LHS-GC calculation process.
Figure 3. LHS-GC calculation process.
Energies 18 02685 g003
Figure 4. A schematic diagram of the IEEE 34-node distribution network system.
Figure 4. A schematic diagram of the IEEE 34-node distribution network system.
Energies 18 02685 g004
Figure 5. A photovoltaic sample curve for a region in South China. (a) Diurnal variation curve of light intensity; (b) light intensity distribution characteristic curve.
Figure 5. A photovoltaic sample curve for a region in South China. (a) Diurnal variation curve of light intensity; (b) light intensity distribution characteristic curve.
Energies 18 02685 g005
Figure 6. Cumulative probability distribution of loop-closing currents for Case 1. (a) Line 26–28 loop steady-state current; (b) line 32–34 loop steady-state current; (c) line 26–28 loop impulse current; (d) line 32–34 loop impulse current.
Figure 6. Cumulative probability distribution of loop-closing currents for Case 1. (a) Line 26–28 loop steady-state current; (b) line 32–34 loop steady-state current; (c) line 26–28 loop impulse current; (d) line 32–34 loop impulse current.
Energies 18 02685 g006
Figure 7. Cumulative probability distribution of loop-closing currents for Case 1. (a) Line 4–6 loop steady-state current; (b) line 11–13 loop steady-state current; (c) line 4–6 loop impulse current; (d) line 11–13 loop impulse current.
Figure 7. Cumulative probability distribution of loop-closing currents for Case 1. (a) Line 4–6 loop steady-state current; (b) line 11–13 loop steady-state current; (c) line 4–6 loop impulse current; (d) line 11–13 loop impulse current.
Energies 18 02685 g007
Figure 8. Feeder current error at 90% cumulative distribution probability. (a) Case 1; (b) Case 2.
Figure 8. Feeder current error at 90% cumulative distribution probability. (a) Case 1; (b) Case 2.
Energies 18 02685 g008
Figure 9. Power loss in distribution network before and after loop closure. (a) Case 1; (b) Case 2.
Figure 9. Power loss in distribution network before and after loop closure. (a) Case 1; (b) Case 2.
Energies 18 02685 g009
Table 1. Configuration of loop-closing operating conditions.
Table 1. Configuration of loop-closing operating conditions.
CasePV Integration NodesLoop-Closing Line LocationsLoad Standard Deviation
123, 2928–3430%
26, 236–1320%
Table 2. Comparison of computation time.
Table 2. Comparison of computation time.
Computational MethodsSingle Cycle Time/sTotal Duration/s
MC method0.00192.43
CM-0.15
LHS-GC method0.00010.76
Table 3. The current indicator values of Case 1 and Case 2.
Table 3. The current indicator values of Case 1 and Case 2.
FeederMaximum Allowable Current-Carrying Capacity/AInstantaneous Current Protection Settings/A
26–283.2413.21
32–342.6236.50
4–612.93143.94
11–1312.73100.68
Table 4. Assessment indices of loop-closing safety under Case 1.
Table 4. Assessment indices of loop-closing safety under Case 1.
Loop CurrentPreliminary Closure ProbabilityMaximum Overrun RateAverage Exceedance Rate
Simulated ValueLHS-GCSimulated ValueLHS-GCSimulated ValueLHS-GC
26–28 Steady-state current94.66%95.92%12.43%13.12%4.37%4.01%
32–34 Steady-state current13.29%14.46%4.89%3.94%
26–28 impact current
32–34 impact current
Table 5. Assessment indices of loop-closing safety under Case 2.
Table 5. Assessment indices of loop-closing safety under Case 2.
Loop CurrentPreliminary Closure ProbabilityMaximum Overrun RateAverage Exceedance Rate
Simulated ValueLHS-GCSimulated ValueLHS-GCSimulated ValueLHS-GC
4–6 Steady-state current91.35%93.16%15.33%16.86%5.11%5.97%
11–13 Steady-state current14.72%15.31%5.25%5.92%
4–6 impact current
11–13 impact current
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Cai, W.; Gao, Y.; Zhang, X.; Si, Q.; Jia, J.; Li, B. Safety Assessment of Loop Closing in Active Distribution Networks Based on Probabilistic Power Flow. Energies 2025, 18, 2685. https://doi.org/10.3390/en18112685

AMA Style

Cai W, Gao Y, Zhang X, Si Q, Jia J, Li B. Safety Assessment of Loop Closing in Active Distribution Networks Based on Probabilistic Power Flow. Energies. 2025; 18(11):2685. https://doi.org/10.3390/en18112685

Chicago/Turabian Style

Cai, Wenchao, Yuan Gao, Xiping Zhang, Qin Si, Jiaoxin Jia, and Bingzhen Li. 2025. "Safety Assessment of Loop Closing in Active Distribution Networks Based on Probabilistic Power Flow" Energies 18, no. 11: 2685. https://doi.org/10.3390/en18112685

APA Style

Cai, W., Gao, Y., Zhang, X., Si, Q., Jia, J., & Li, B. (2025). Safety Assessment of Loop Closing in Active Distribution Networks Based on Probabilistic Power Flow. Energies, 18(11), 2685. https://doi.org/10.3390/en18112685

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop