An Unsupervised Neural Network Approach for Solving the Optimal
Power Flow Problem
Alexander Svensson Marcial
a
and Magnus Perninge
b
Department of Physics and Electrical Engineering, Linnaeus University, V
¨
axj
¨
o, Sweden
Keywords:
AC-OPF, Unsupervised Learning, Deep Learning.
Abstract:
Optimal Power Flow is a central tool for power system operation and planning. Given the substantial rise
in intermittent power and shorter time windows in electricity markets, there’s a need for fast and efficient
solutions to the Optimal Power Flow problem. With this in consideration, this paper propose an unsupervised
deep learning approach to approximate the optimal solution of Optimal Power Flow problems. Once trained,
deep learning models benefit from being several orders of magnitude faster during inference compared to
conventional non-linear solvers.
1 INTRODUCTION
Optimal power flow (OPF) is a fundamental tool used
in power system operation and planning. Since its
initial formulation in the 1960s (Carpentier, 1962)
OPF has been instrumental in numerous applications,
encompassing economic dispatch, securing reactive
power reserves for voltage stability (Capitanescu,
2011), generation and transmission expansion plan-
ning (Skolfield and Escobedo, 2022). Furthermore,
it serves as a foundational tool for market clearing in
deregulated electricity markets.
In general, OPF is a non-convex optimization
problem that classifies as NP-hard (Lavaei and Low,
2011). Its objective is to find an optimal dispatch
p
G
that minimises a specific performance metric
J(x, p
G
),where x denotes the state of the power sys-
tem. A feasible dispatch p
G
U ensures the power
system’s secure operation while aligning generation
with demand.
Efforts to develop efficient algorithms for OPF
have led to the introduction of several approximate
methods. Most notably is the DC-OPF formula-
tion which linearizes power flow constraints by ne-
glecting line resistances under the assumption that
line resistances are significantly lower than line re-
actances, combined with the approximation of a flat
voltage profile and small voltage angle deviations be-
tween nodes. While the objective function is often
a
https://orcid.org/0000-0002-2028-9847
b
https://orcid.org/0000-0003-3111-4820
a smooth convex function, which enables the DC-
OPF to be effectively solved with standard solvers,
this computational efficiency trades off against accu-
racy. Specifically, DC-OPF tends to under perform in
heavily loaded systems and often doesn’t produce fea-
sible solutions for the original OPF-problem (Baker,
2021).To counteract these infeasibilities, DC-OPF is
generally recomputed iteratively, incorporating con-
straint adjustments (Low, 2014).
The task faced by grid operators, ensuring that en-
ergy supply meets demand, has grown considerably
more difficult due to the rapid increase of intermittent
energy sources such as wind and photovoltaic sys-
tems. Unlike conventional power sources, intermit-
tent energy systems exhibit a stochastic energy pro-
duction, with an output dependent on weather condi-
tions as well as the time of the day. Consequently, to
adapt to the increased randomness of injected power,
stochastic optimization models have been proposed
in literature. Nonetheless, the underlying difficulties
in solving OPF remains in stochastic formulations.
This is why the proposed methods found in literature
employ the DC-OPF approximation or convex relax-
ations to reduce the computational burden,
This increase in complexity calls for more com-
putational efficient algorithms, capable of solving the
OPF in short timescales (Tang et al., 2017).. Rapid
results from OPF is not only essential for real-time
grid management but also for addressing economic
dispatch problems in electricity markets with short
market time units.
In recent years, the prowess of machine learning,
214
Marcial, A. and Perninge, M.
An Unsupervised Neural Network Approach for Solving the Optimal Power Flow Problem.
DOI: 10.5220/0012187400003543
In Proceedings of the 20th International Conference on Informatics in Control, Automation and Robotics (ICINCO 2023) - Volume 1, pages 214-220
ISBN: 978-989-758-670-5; ISSN: 2184-2809
Copyright © 2023 by SCITEPRESS Science and Technology Publications, Lda. Under CC license (CC BY-NC-ND 4.0)
and in particular deep neural networks (DNNs) have
found a wide range of applications from the typical
classification problem to applications in optimization
theory, aimed at curtailing the on-line computational
burden. Within the research aimed towards optimal
power flow, a number of recent papers have utilized
DNNs in an effort to solve the OPF problem.
(Zhang and Zhang, 2022) does not directly pro-
vide predictions for the OPF, instead, they employ a
supervised learning approach to forecast an appropri-
ate warm-start for a conventional non-linear solver.
Similarly, (Fioretto et al., 2020) adopts a supervised
methodology, utilizing a lagrangian formulation to
enforce the constraints. Meanwhile, in (Huang et al.,
2021) while a supervised approach is presented no ef-
fort is done during training to enforce the constraints,
instead they develop a post-processing step to en-
hance the feasibility of the neural network’s predic-
tions.
This paper’s primary contribution is the introduc-
tion of an unsupervised methodology for predicting
the solutions to the OPF problem. Moreover, we de-
tail the utilization of a custom loss function, leverag-
ing the augmented lagrangian method to train the neu-
ral network towards optimal solutions. Consequently,
our approach removes the dependence of extensive
datasets of labeled data.
2 OPTIMAL POWER FLOW
In this section we describe the underlying model used
in Optimal Power Flow as well as presenting the gen-
eral AC-OPF formulation.
2.1 Power System Model
Power systems are modeled at the system level as
buses interconnected by transmission lines, with gen-
erators and loads connected to a subset of these buses.
A power system can thus be conveniently modelled as
a graph P = (N ,B), where N N represents the set
of edges connecting the nodes. We will adhere to the
naming conventions of power systems, referring to N
as the set of nodes and B the set of transmission lines.
In modelling the power system, two equivalent for-
mulations can be used (Low, 2014): one where P is
a directed graph and the other where P is an undi-
rected graph. In this paper, we have chosen the latter
formulation; thus if (k,l) B then (l, k) B. For an
n-bus system, we assume, without loss of generality,
that N = {0,...,n 1}. Additionally, there is a bus
r
B
N which we refer to as the reference bus.
Each transmission line connecting two buses has
the admittance y
i,k
= g
i,k
+ jb
i,k
, (i,k) B. There
may also be an impedance connecting a bus to ground,
denoted by y
s,i
for i N
The power system has a set of generators, indexed
by G N . For each i G, the associated genera-
tor can produce active power p
gi
R
+
and reactive
power q
gi
R. Let p
G
:= (p
gi
)
iG
R
|N |
+
represent
the vector of power generated in the power system.
Similarly, let q
G
:= (q
gi
)
iG
R
|N |
be the vector of
reactive power produced by the generators in the sys-
tem. Sometimes, it’s more practical to work with the
apparent power produced by the generators. In that
case, let s
G
:= (s
gi
)
iG
C
|N |
, where s
gi
= p
gi
+ jq
gi
.
Unlike the generators, each bus can be charac-
terized by the load s
di
= p
di
+ jq
di
, noting that for
some i N , s
di
= 0 + j0. To identify the non-
zero loads, we define D
p
:= {i N : p
di
̸= 0} and
D
q
:= {i N : q
di
̸= 0}. Further, let p
D
:= (p
di
)
iD
p
and q
D
:= (q
di
)
iD
q
.
In terms of bus-related variables, let V :=
(v
i
)
iN
C
|N |
denote the voltage at each bus, where
v
i
= V
i
e
jδ
i
. We further define the vectors of voltage
magnitude and angles as |V | := (V
1
,...,V
n
) R
|N |
and
V := (δ
1
,...,δ
n
) [π,π]
|N |
respectively. We as-
sume that δ
r
B
= 0 and that δ
i
, i N \{r
B
} is measured
with r
B
as reference.
Assuming that |N | = n, we define S :=
(s
1
,...,s
n
) C
|N |
as the vector of net apparent power
injection in each node of the power system. The
nodal apparent power injection is determined using
Kirchoffs law:
s
i
= v
i
k:(i,k)B
y
i,k
(v
i
v
k
) + |v
i
|
2
y
si
, i N (1)
The net apparent power injection in a node is
equivalent to the difference between the scheduled
generation s
gi
and the scheduled load s
di
;
s
i
=
s
gi
s
di
i G
s
di
i N \G.
(2)
Note that, if i / D
p
D
q
then s
di
= 0 + j0
Generators have both lower and upper limits in
terms of active and reactive power output. For a gen-
erator located at bus i G, the range of its active
power output is constrained to p
gi
[P
min
gi
,P
max
gi
]. Sim-
ilarly the reactive power is constrained to the interval
q
gi
[Q
min
gi
,Q
max
gi
].
To ensure stable operation of the power system,
it’s imperative that the voltages across the buses re-
main within acceptable levels. Consequently, the
voltage magnitude for bus i N is confined to the
interval [V
min
i
,V
max
i
].
An Unsupervised Neural Network Approach for Solving the Optimal Power Flow Problem
215
We define the restrictions on the generators capa-
bilities and bus voltage limits to be the sets
A :=
iG
[P
min
gi
,P
max
gi
] (3)
R :=
iG
[Q
min
gi
,P
max
gi
] (4)
V :=
iN
[V
min
i
,V
max
i
] (5)
2.2 Optimization Model
In this paper, the OPF formulation we employ focuses
on economic dispatch. Specifically the objective is to
identify the operating point for generators that results
in the most cost efficient production. Given a cost
function C : A R, the OPF can be expressed as
min
V
C(p
G
) (6)
s.t. g(V, S) = 0 (7)
p
G
A (8)
q
G
R (9)
|V | V . (10)
While there exist various ways to formulate the cost
function given by equation (6), polynomial and piece-
wise linear approximations are frequently encoun-
tered in literature. For the purpose of this paper, we
opt for the polynomial approach. Consequently, the
cost function is expressed as:
C(p
G
) :=
iG
(C
i
2
p
2
gi
+C
i
1
p
gi
+C
i
0
) (11)
Note that, the specific formulation of the cost function
does not limit the applicability of this paper; the pro-
posed methodology can accommodate a diverse range
of formulations.
The power balance at each bus within the power
system is represented by equation (7) and is defined
as:
g
i
(V,S) = f
i
(V ) s
i
, i N (12)
where f
i
(V ) corresponds to the right-hand side of
(1)
3 DNN APPROXIMATION OF
OPTIMAL SOLUTION
3.1 Function Approximation
In this paper, we assume that the power system de-
scribed by P is static, meaning the network topol-
ogy remains unaltered. Therefore, given (p
D
,q
D
)
D
p
× D
q
, the objective is to determine the optimal
voltage vector V
(if it exists) that solves the OPF
instance defined by (p
D
,q
D
). With this context, the
OPF can be perceived as an operator that maps a given
instance defined by the loading into an optimal solu-
tion, (Zhou et al., 2020),(Falconer and Mones, 2022)
F : P
D
× Q
D
C
|N |
(13)
here, P
D
and Q
D
represent sets of active and reactive
loads, respectively.
For the development of the proposed approxima-
tion of the solution to the OPF, it’s advantageous to
modify the standard OPF formulation. Consider the
power flow equations in (1), they can be divided into
two sets of equations,
s
gi
= s
i
+ s
di
, i G (14)
0 = s
i
+ s
di
, i N \G. (15)
Combining (14) with (8)-(9) yields the inequalities
P
min
gi
Re(s
i
+ s
di
) P
max
gi
(16)
Q
min
gi
Im(s
i
+ s
di
) Q
max
gi
(17)
Further we define the set
U(p
D
,q
D
) := {V C
|N |
: (16),(17),(10) feasible}.
(18)
Then, given s
D
P
D
× Q
D
and V U(s
D
) an operat-
ing point (p
G
,q
G
) can be recovered from (7).
Evaluating F using conventional iterative solvers
is cumbersome. To reduce the computational burden
at inference time we therefore propose the use of a
neural network, denoted
ˆ
F(p
D
,q
D
;θ) to approximate
(13), where θ is the parameters that are learned during
training of the neural network.
A major problem is to ensure that
ˆ
F(p
D
,q
D
;θ) U(p
D
,q
D
). Lagrangian relaxation
is a common method in conventional constrained
optimization and it has been used in training of neural
networks for constrained optimization problems. In
(Zhang and Zhang, 2022) it is used to solve super-
vised AC-OPF using a two networks for predicting
the primal and dual variables respectively. In (Pan
et al., 2020), (Pan et al., 2022) a penalty function is
used as a regularization term in the loss function for
penalizing deviations from the constraints.
ICINCO 2023 - 20th International Conference on Informatics in Control, Automation and Robotics
216
Figure 1: Neural network architecture.
3.2 Unsupervised Learning
To train
ˆ
F(·;θ) using supervised learning, one re-
quires a substantial dataset of load samples, denoted
as D
x
= {(p
i
D
,q
i
D
)}
Γ
i=1
, where (p
i
D
,q
i
D
) P
D
× Q
D
.
Correspondingly, there is a set D
y
= {y
i
}
Γ
i=1
repre-
senting the ground truth, with y
i
= F(p
i
D
,q
i
D
). The
neural network is then trained using some loss func-
tion, e.g. the empiric risk, as shown in (19). Reg-
ularization terms can further be incorporated to pe-
nalize any constraint violations. Equation (19) will
be revisited in subsequent sections, particularly when
comparing the supervised and unsupervised learning
methodologies.
min
θ
1
Γ
Γ
i=1
ˆ
F(x
i
;θ) y
i
2
2
,
(19)
Given that y
i
U(s
i
D
), the network is trained to
imitate the true solution. Under an unsupervised
framework, the label set D
y
is eliminated. This means
the network need to find an optimal θ
such that
ˆ
F(p
D
,q
D
;θ
) F(p
D
,q
D
), without the guidance of
ground truth. In essence, for each sample in D
x
the
neural network need to predict a feasible voltage that
minimize the cost. The unsupervised training can
therefore be formulated as the constrained optimiza-
tion problem
argmin
θ
Γ
i=1
C(
ˆ
F(S
i
D
;θ))
s.t.
ˆ
F(p
i
D
,q
i
D
;θ) V (S
i
D
) (p
i
D
,q
i
D
) D
x
(20)
Similar to approaches used in conventional opti-
mization algorithms, (20) is in this paper relaxed to an
unconstrained optimization problem. Specifically, we
will use the Augmented Lagrangian Method (ALM)
(Bertsekas, 2016) which is a method that combines
the Lagrangian method and penalty methods. Pre-
vious works in approximating constrained optimiza-
tion problems using machine learning have reported
ALM successful. In particular, ALM has been used
in physics informed neural networks such as in (Djeu-
mou et al., 2022) (Lu et al., 2021) (Dener et al., 2020)
3.3 Neural Network Architecture
Consistent with the previous cited works of using ma-
chine learning for OPF, the neural network architec-
ture employed in this paper is a feed-forward architec-
ture, (Figure 1), as it has shown good performance in
supervised settings. For the unsupervised approach, a
single neural network is utilized. However, for the su-
pervised approach, two neural networks are deployed:
one to predict voltage magnitudes and another for
voltage angles.
As highlighted in section 3.1, the neural network’s
input comprises the load (p
D
,q
D
) R
|N |
× R
|N |
.
This configuration corresponds to an input layer in
the neural network with 2|N | neurons. However, it is
worth noting that since generally N \(D
p
D
q
) ̸=
/
0,
we can effectively reduce the neuron count in the in-
put layer to |D
p
| + |D
q
|.
Subsequent to the input layer, there is a sequence
of hidden layers, L
i
, with each layer comprising n
i
neurons. Both the depth (number of layer) and width
(neurons per layer) of this structure are considered hy-
perparameters, that are problem specific. An increas-
ing size of N benefits from an increasing number of
neurons per layer. In the specific architecture used
in this paper, all activation function within the hidden
layers are Rectifier Linear Units (ReLU).
The output layer is designed with 2|N | 1 neu-
rons. Out of these, |N | neurons are allocated for rep-
resenting the voltage magnitude V , and the remainder
serves as outputs for the voltage angle, recalling that
the bus angle of bus n
r
B
is assumed known.
One strategy to integrate a subset of the con-
straints into the architecture of the neural network
is to use appropriate activation functions on the out-
put layer of the neural network. Specifically, let
x = (x
v
,x
δ
) be the output vector of the neural net-
work, corresponding to the voltage magnitudes and
angles respectively. Then using e.g. the sigmoid
activation function we obtain the mapping σ(x) =
(σ
v
(x
v
),σ
δ
(x
δ
)) [0,1]
|N |
× [0,1]
|N |−1
. The actual
voltage magnitude and angles can then be obtained
by
σ
V
(x
v
) 7→ (V
min
i
)
iN
+
σ
v,i
(x
i
)(V
max
i
V
min
i
)
iN
(21)
and the corresponding angles are obtained as
An Unsupervised Neural Network Approach for Solving the Optimal Power Flow Problem
217
σ
δ
(x
δ
) 7→ (δ
min
i
)
iN \{n
R
}
+
σ
i
(x
i
)(δ
max
i
δ
min
i
)
iN \{n
r
B
}
(22)
Through this methodology, the neural network ef-
fectively enforces the voltage constraints.
3.4 Loss Function
To solve (20), ALM is used to relax the constrained
optimization problem into an unconstrained problem.
In general, ALM algorithm consists of solving a se-
quence of sub-optimization problems
x
k
= argmin
x
L
k
(x,λ
k
,µ
k
,ρ
k
)
(23)
with,
L
k
(x,λ
k
,µ
k
,ρ
k
) = f (x) + λ
T
k
g(x) + µ
T
k
h(x)+
+
ρ
k
2
h(x)
2
2
+
ρ
k
2
φ(h(x))
2
2
(24)
Where, λ
k
and µ
k
are Lagrange multipliers and the
norms serves as penalty terms, weighted by the
penalty parameters ρ
k
, and φ(h(x) = max{0,h(x)}.
Then, under some mild regularity conditions it can be
shown that lim
k
{x
k
} = x
, solves the original opti-
mization problem.
In practice, each sub-problem is finitely iterated
until a stopping criteria ||
x
L|| < ε
k
(Bertsekas,
2016), however in the training of the neural network
we have chosen a slight different approach were we
instead use stochastic gradient descent until each sub-
optimization problem stops improving, which is cov-
ered in the next section.
3.5 Training
The load samples set, D
x
is divided into two disjoint
sets D
T
and D
E
out of which the former is used for
training and the latter for evaluating the performance
of the trained network.
In the spirit of ALM, the training of
ˆ
F involves
solving a sequence of optimization problems. For
each outer iteration k {0,...,K 1}, the goal is to
determine the optimal parameter θ
k
corresponding to
sub-optimization k. Within each outer-iteration, the
neural network is trained using the Adam optimizer
(Kingma and Ba, 2017). The training process follows
the commonly used mini-batch approach, where the
training set is randomly segmented into mini-batches
B
i
D
T
. Each mini-batch is used to compute the gra-
dient and subsequent weight updates for the network.
The entire dataset is processed in this fashion up to
n
epoch
times.
In an effort to avoid over fitting and reduce train-
ing time, an early stopping mechanism with a preset
Figure 2: Neural network configuration for unsupervised
learning.
patience is employed. The loss function L
k
is evalu-
ated after each epoch. If no improvement is recorded
for a fixed number of consecutive epochs, the train-
ing of the k-th optimization problem is halted. Ad-
ditionally, adaptive learning rates are used, reducing
the learning rate if there is no observed improvement
in loss. The learning rate is reverted back to the ini-
tial learning rate in each new outer iteration. Once an
inner iteration is completed, the Lagrange multipliers
are updated in accordance with the ALM algorithm:
λ
k+1
= λ
k
+ ρ
k
g(x
k
)
µ
k+1
= max{0, ρ
k
h(x
k
)}
(25)
Data: Training dataset D
Result: optimal parameter θ
λ
0
0;
µ
0
0;
ρ
0
ρ
init
;
for k 0 to K 1 do
minimize L
k
(θ,λ
k
,µ
k
,ρ
k
) using Adam.;
λ
k+1
λ
k
+ ρ
k
g(
ˆ
F(S
D
;θ
k
);
µ
k+1
max{0, µ
k
+ ρ
k
h(
ˆ
F(S
D
;θ
k
)};
if h(x)
> ηh
then
h
h(x)
ρ
k+1
min{γρ
k
,ρ
max
}
end
end
Algorithm 1: Training the neural network.
4 NUMERICAL RESULTS
4.1 Hyperparameters
The proposed approach is conducted on the IEEE
300-bus system. The neural network is developed us-
ing the pyTorch library and trained on a MacBook pro
2 GHz Quad-Core Intel Core i5, 16 GB RAM.
For the supervised approach, both neural networks
feature an architecture with four hidden layers, with
a sequentially number of neurons: 1024,768,512 and
256. Meanwhile, for the unsupervised approach, the
hyperparameters with respect to hidden layers has a
network configuration as illustrated in Figure 2. This
configuration comprises hidden layers with neuron
counts of 2048,1024,768,512 and 256.
ICINCO 2023 - 20th International Conference on Informatics in Control, Automation and Robotics
218
For the training dataset, we utilize a total of |D| =
30000 samples of loads. The loads exhibit correla-
tions. The outer-loop is executed for 50 iterations,
while the inner iteration involves up to 50 epochs.
To facilitate learning, we employ an exponential de-
cay strategy for the learning rate with a decay factor
of γ = 0.9. The learning rate is decayed if the loss
does not improve for 5 consecutive epochs. We ini-
tialize the learning rate at 10
4
. Furthermore, we set
the early stopping patience to 10 epochs, allowing for
timely termination if necessary. Regarding the ALM
algorithm, we set the initial penalty parameter, ρ
0
to
500, while the Lagrange multipliers are initialized as
zeros.
4.2 Performance
Figure 3: Mean violation of the equality constraints.
Figure 4: Optimality gap of the objective function.
The dataset D was divided into a training set D
T
and
a test set D
E
with the proportion |D
T
= 30000 and
|D
E
| = 5000. The dataset used is labeled with the op-
timal solution retrieved from Matpower (Zimmerman
et al., 2011)(Wang et al., 2007). The unsupervised ap-
proach was trained using the Mean Squared Error as
loss function, without any regularization terms added.
The results obtained using the test set are comprehen-
sively displayed in Table 1. Histograms illustrating
the equality violation as and the optimality gap, when
using the unsupervised approach, can be seen in Fig-
ure 3-4 respectively.
Using the labeled data, the optimality gap is computed
as
O
gap
= 100
ˆ
F(s
D
;θ
) F(s
D
)
F(s
D
)
. (26)
The values reported in Table 1 are the average across
the entire test set, except the maximum equality con-
straint violation which shows the maximum across the
entire test set.
Table 1: Result of supervised and unsupervised training.
Unsupervised Supervised
Opt.gap [%] 0.316 0.216
Eq. 1.3·10
3
2.6·10
3
Ineq. 6.7·10
7
5.4·10
6
Max eq. 0.023 0.068
5 CONCLUSION
The need for fast and accurate solutions of the Opti-
mal Power Flow problem is becoming more important
as the power systems are experiencing more unpre-
dictability in the form of intermittent power. While
training neural networks do not offer a lighter com-
putational workload compared to conventional itera-
tive non-linear solvers, they have a distinct advantage
in that they shift the computational load off-line. A
trained neural network outputs the solution in negli-
gible time. This makes neural network a promising
candidate for time sensitive applications.
Although the computational burden of supervised
learning is offline, it comes with its own challenges. It
demands a considerable amount of labeled data. This
translates to the need for a repository of OPF solutions
coming from solver like the interior-point method.
This paper primarily delved into the application
of unsupervised learning for addressing the AC-OPF
problem, without considering post-processing. A log-
ical extension to this paper would be to explore other
neural network architectures. The feed-forward archi-
tecture, as employed in this paper, might encounter
challenges when confronted with significantly larger
power grid models. Hence it is therefore interesting
to investigate other architectures that scales well with
increasingly complex models.
An Unsupervised Neural Network Approach for Solving the Optimal Power Flow Problem
219
REFERENCES
Baker, K. (2021). Solutions of dc opf are never ac feasible.
In Proceedings of the Twelfth ACM International Con-
ference on Future Energy Systems, pages 264–268.
Bertsekas, D. (2016). Nonlinear Programming, volume 4.
Athena Scientific.
Capitanescu, F. (2011). Assessing reactive power re-
serves with respect to operating constraints and volt-
age stability. IEEE Transactions on Power Systems,
26(4):2224–2234.
Carpentier, J. (1962). Contribution to the economic dis-
patch problem. Bulletin de la Societe Francoise des
Electriciens, 3(8):431–447.
Dener, A., Miller, M. A., Churchill, R. M., Munson, T., and
Chang, C.-S. (2020). Training neural networks under
physical constraints using a stochastic augmented la-
grangian approach. arXiv preprint arXiv:2009.07330.
Djeumou, F., Neary, C., Goubault, E., Putot, S., and Topcu,
U. (2022). Neural networks with physics-informed
architectures and constraints for dynamical systems
modeling. In Learning for Dynamics and Control
Conference, pages 263–277. PMLR.
Falconer, T. and Mones, L. (2022). Leveraging power grid
topology in machine learning assisted optimal power
flow.
Fioretto, F., Mak, T. W., and Van Hentenryck, P. (2020).
Predicting ac optimal power flows: Combining deep
learning and lagrangian dual methods. In Proceedings
of the AAAI conference on artificial intelligence, vol-
ume 34, pages 630–637.
Huang, W., Pan, X., Chen, M., and Low, S. H. (2021).
Deepopf-v: Solving ac-opf problems efficiently. IEEE
Transactions on Power Systems, 37(1):800–803.
Kingma, D. P. and Ba, J. (2017). Adam: A method for
stochastic optimization.
Lavaei, J. and Low, S. H. (2011). Zero duality gap in
optimal power flow problem. IEEE Transactions on
Power Systems, 27(1):92–107.
Low, S. H. (2014). Convex relaxation of optimal power
flow—part i: Formulations and equivalence. IEEE
Transactions on Control of Network Systems, 1(1):15–
27.
Lu, L., Pestourie, R., Yao, W., Wang, Z., Verdugo, F.,
and Johnson, S. G. (2021). Physics-informed neu-
ral networks with hard constraints for inverse design.
SIAM Journal on Scientific Computing, 43(6):B1105–
B1132.
Pan, X., Chen, M., Zhao, T., and Low, S. H. (2022). Deep-
opf: A feasibility-optimized deep neural network ap-
proach for ac optimal power flow problems.
Pan, X., Zhao, T., and Chen, M. (2020). Deepopf: Deep
neural network for dc optimal power flow.
Skolfield, J. K. and Escobedo, A. R. (2022). Operations
research in optimal power flow: A guide to recent and
emerging methodologies and applications. European
Journal of Operational Research, 300(2):387–404.
Tang, Y., Dvijotham, K., and Low, S. (2017). Real-time op-
timal power flow. IEEE Transactions on Smart Grid,
8(6):2963–2973.
Wang, H., Murillo-Sanchez, C. E., Zimmerman, R. D., and
Thomas, R. J. (2007). On computational issues of
market-based optimal power flow. IEEE Transactions
on Power Systems, 22(3):1185–1193.
Zhang, L. and Zhang, B. (2022). Learning to solve the ac
optimal power flow via a lagrangian approach. In 2022
North American Power Symposium (NAPS), pages 1–
6. IEEE.
Zhou, F., Anderson, J., and Low, S. H. (2020). The optimal
power flow operator: Theory and computation.
Zimmerman, R. D., Murillo-S
´
anchez, C. E., and Thomas,
R. J. (2011). Matpower: Steady-state operations, plan-
ning, and analysis tools for power systems research
and education. IEEE Transactions on Power Systems,
26(1):12–19.
ICINCO 2023 - 20th International Conference on Informatics in Control, Automation and Robotics
220