Next Article in Journal
VF-Mask-Net: A Visual Field Noise Reduction Method Using Neural Networks
Next Article in Special Issue
Harnessing the Radio Frequency Power Level of Cellular Terminals for Weather Parameter Sensing
Previous Article in Journal
AE-Qdrop: Towards Accurate and Efficient Low-Bit Post-Training Quantization for A Convolutional Neural Network
Previous Article in Special Issue
A Multifaceted Exploration of Atmospheric Turbulence and Its Impact on Optical Systems: Structure Constant Profiles and Astronomical Seeing
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Optimal Power Allocation in Optical GEO Satellite Downlinks Using Model-Free Deep Learning Algorithms

by
Theodore T. Kapsis
,
Nikolaos K. Lyras
and
Athanasios D. Panagopoulos
*
School of Electrical and Computer Engineering, National Technical University of Athens, GR15780 Athens, Greece
*
Author to whom correspondence should be addressed.
Electronics 2024, 13(3), 647; https://doi.org/10.3390/electronics13030647
Submission received: 21 December 2023 / Revised: 28 January 2024 / Accepted: 2 February 2024 / Published: 4 February 2024
(This article belongs to the Special Issue New Advances of Microwave and Optical Communication)

Abstract

:
Geostationary (GEO) satellites are employed in optical frequencies for a variety of satellite services providing wide coverage and connectivity. Multi-beam GEO high-throughput satellites offer Gbps broadband rates and, jointly with low-Earth-orbit mega-constellations, are anticipated to enable a large-scale free-space optical (FSO) network. In this paper, a power allocation methodology based on deep reinforcement learning (DRL) is proposed for optical satellite systems disregarding any channel statistics knowledge requirements. An all-FSO, multi-aperture GEO-to-ground system is considered and an ergodic capacity optimization problem for the downlink is formulated with transmitted power constraints. A power allocation algorithm was developed, aided by a deep neural network (DNN) which is fed channel state information (CSI) observations and trained in a parameterized on-policy manner through a stochastic policy gradient approach. The proposed method does not require the channels’ transition models or fading distributions. To validate and test the proposed allocation scheme, experimental measurements from the European Space Agency’s ARTEMIS optical satellite campaign were utilized. It is demonstrated that the predicted average capacity greatly exceeds other baseline heuristic algorithms while strongly converging to the supervised, unparameterized approach. The predicted average channel powers differ only by 0.1 W from the reference ones, while the baselines differ significantly more, about 0.1–0.5 W.

1. Introduction

The progressively demanding criteria of fifth-generation (5G) mobile communications sparked the deployment of high-throughput short-range links and mesh topologies to cater to the increased user capacity and to reduce energy consumption [1,2]. The sub-10 GHz radio frequency (RF) band is almost exhausted and strictly regulated. Consequently, scientific research is moving towards the incorporation of millimeter and nanometer wavelengths [1,2]. Optical wireless communications (OWC) operate similarly to fiber optics by modulating a coherent laser beam that propagates point to point and by line of sight [2,3,4,5,6]. Free-space optical (FSO) technology has excellent backhauling capabilities of ultra-fast transfer of traffic between antenna towers and small cells [2,3,4,5,6]. FSO also mitigates the last mile problem, i.e., congestion in the component linking the user to the Cloud/Internet [2,3,4,5,6].
Satellite communications (SatComs) cover a large portion of the Earth, including remote locations, and enable a vast variety of forecasting and broadcasting applications [6,7]. Yet, the cost of designing and manufacturing them is substantial, satellite debris causes “space pollution”, and a satellite constellation requires numerous satellites in addition to very quick handovers to guarantee visibility and continuity [6,7]. Especially for high-speed SatComs, FSO systems have exhibited great potential due to their easy installment (<30 min), their operation with low initial expense and maintenance, no licensing requirements like RF systems, and their secure connections due to their large antenna gains, allowing several FSO links to be deployed in parallel and in proximity [2,3,4,5,6]. Nevertheless, they demand advanced pointing, acquisition, and tracking components. The process of creating a private encryption key between two parties is known as quantum key distribution (QKD) [7]. Fundamentally, QKD is an optical technology that can supply encryption keys for any two locations connected by an optical link [7]. The application of QKD over optical fiber, on the other hand, is restricted by exponential fiber losses. In this setting, QKD over satellites is becoming more popular. The capacity to put up robust optical links and guarantee a minimal quantum bit error rate (QBER) by overcoming numerous transmission obstructions is critical to the success of QKD over satellites [7].
However, the FSO beam is susceptible to various problematic atmospheric phenomena, i.e., absorption, scattering, cirrus clouds, and turbulence [2,3,4,5,6]. Physical obstructions, geometric losses, and the blockage of the link caused by cloud occurrence are some of the difficulties [8,9]. For mitigating the cloud occlusion, site diversity can be used [8]. Received irradiance scintillation is created from rapid changes in wind speed, pressure, and temperature, which in turn induce changes in the refractive index [10,11,12]. The impact of scintillation depends primarily upon the time of day, the elevation angle of the link, and the altitude of the station [10,11,12]. In the daytime, at low elevation angles and low-altitude stations (denser atmosphere), turbulence is more extreme [10,11,12].
In [10,11,12], the power allocation (PA) problem is investigated for optical satellite downlinks under weak turbulence and solved using a Karush–Kuhn–Tucker (KKT), water-filling-inspired algorithm. However, the reported methodology depends on the knowledge of the system and channel model. In [13,14], a radio-on-FSO (RoFSO) wavelength multiplexing scenario is considered where the channel model is assumed unknown. The developed model-free primal–dual deep learning algorithm (PDDL) strongly converges to the precise solution derived from the model-based algorithm. In both studies, however, synthetic data were employed while the optical satellite-to-ground scenario was not explored. Likewise, in [15] a MIMO FSO system is studied and the PA problem is solved via reinforcement learning (RL) in a deep deterministic policy gradient (DDPG) approach. More power allocation problems regarding FSO, RF, and terrestrial and non-terrestrial networks exploiting ML and deep RL techniques can be found in [16,17,18,19,20,21,22].
In this paper, several channel-model-free methodologies and heuristics are explored for optimal PA, and then compared to the exact, model-based solution for an all-optical, multi-aperture satellite downlink between a geostationary (GEO) satellite and an optical ground station (OGS). Deep learning (DL) constitutes a powerful tool to handle data from complex and fading communication channels [23,24,25]. Therefore, the PA problem is formulated as a constrained learning optimization problem with peak and total expected power inequality constraints. Specifically, a deep reinforcement learning (DRL) module is proposed that assists the agents in producing actions via a stochastic policy gradient technique, given channel state information (CSI) observations from the environment. The policy is structured as a deep neural network (DNN) and trained according to the REINFORCE (REward Increment = Nonnegative Factor times Offset Reinforcement times Characteristic Eligibility) algorithm [26], but modified to include the power constraints, and the multi-agent optical environment where the agents act, independently accessing only their local observations but collaboratively trying to maximize the global reward. The proposed PA strategy is deemed appropriate for difficult optical conditions since it learns explicitly from observation. Whilst the optimization problem has been approached in the literature, it has never been presented in a MIMO optical satellite scenario along with available experimental data.
Summarizing the main contributions of this paper:
-
We propose a DRL-aided algorithm to optimally allocate power in an optical GEO-to-ground multi-aperture system. The proposed method accurately adjusts the expected power for each optical channel, without any knowledge of path losses and scintillation conditions. Only CSI samples are utilized. The total expected power and peak power are constrained. Although in an LEO/MEO scenario deep learning would be even more beneficial, the GEO optical CSI is still unstable due to variations in the refractive index structure parameter along the slant path, thin clouds that may attenuate or block the laser beam, and pointing and tracking errors. Thus, it is challenging to have a more accurate system model and fading distribution knowledge in the short and long term.
-
Instead of simulated data, experimental irradiance time series from the European Space Agency’s (ESA’s) ARTEMIS optical satellite sessions were employed to evaluate the performance of the proposed methodology [27,28].
-
The achieved ergodic system capacity from the application of the proposed algorithm greatly exceeds the performance of the model-free Equal Power, Random Power, and Deep Q-Learning Network (DQN) schemes [20,21,22] and approaches the model-based, unparameterized solution with very good agreement.
-
An investigation of the impact of the number of hidden layers and neurons, policy distribution, and hyperparameter selection and overfitting effects was carried out.
-
The proposed solution differs from other standard learning formulas because it applies to a multi-agent optical satellite PA problem based on the parameter sharing approach, allowing centralized learning under a single policy for faster convergence. The learning model is scalable as it has been tested in scenarios with a large number of optical satellite downlinks and a great amount of data and retained its performance. It is especially more scalable than the DQN algorithm because the Q-table is not scalable when there are large, high-dimensional, and continuous state–action pairs [20,21,22].
In Section 2, the optical carrier is described, the system model is presented, and the PA problem is formalized as a learning program. In Section 3, the proposed DRL-aided methodology is reported along with other heuristic, model-free strategies. In Section 4, the ARTEMIS mission is briefly discussed, and experimental measurements are employed to evaluate and compare the considered PA methodologies. Performance results are drawn and commented on. Section 5 concludes this work.

2. Channel Model and Power Allocation Problem

The turbulent cells of the atmosphere, known as eddies, work as a prism that will enhance or degrade a propagating optical signal [3]. If an eddy’s diameter is almost the size of an incoming beam, it will result in received irradiance I r fluctuations called scintillation, which is the primary factor of deterioration in the FSO downlink [3]. The scintillation index (SI) σ I 2 constitutes the normalized ratio of the standard deviation of received irradiance fluctuations to the mean received irradiance [3,29]:
σ I 2 = I r 2 I r 2 I r 2
where I r represents the mean received optical irradiance in W / m 2 . In satellite downlinks with σ I 2 < 1 and elevation angle > 20°, the atmospheric turbulence is considered weak [10,11,12]. In weak turbulence conditions, the optical channel follows the Lognormal ( μ ,   σ I 2 ) with Probability Density Function (PDF) [5]:
f I r ( I r ) = 1 I r 2 π σ I exp { [ ln ( I r / I r ) + 0.5 σ I 2 ] 2 2 σ I 2 }
where μ = 0.5 σ I 2 .
The optical satellite downlink system under investigation is represented by the real-valued channel matrix H N × M where N is the number of ground receivers (Rx), and M is the number of on-board transmitters (Tx). It is assumed that the covariance matrix Σ = Ε [ ( H Ε [ H ] ) ( H Ε [ H ] ) T ] has zero non-diagonal elements, i.e., linearly uncorrelated channels. In practice, this is achieved by having a distance between the Rx elements larger than Fried’s parameter [1,29]. The CSI of the channels is denoted as h i j 1 i N ,   1 j M and the allocated power as P i j h i j . The channel capacity is C i j P i j h i j , h i j . The average (ergodic) channel capacity is E C i j P i j h i j , h i j .
For independent channels, and a proportionate number of Rx and Tx: N = M = L and i = j = l, the total average capacity maximization problem is mathematically formalized as follows:
( OP ) max { P l ( h l ) : l } l = 1 L Ε [ C l ( P l ( h l ) , h l ) ]
s . t . 0 P l ( h l ) P s , l i = 1 L E [ P l ( h l ) ] P a v
where P s is the maximum safety power allocated to a single channel, and P a v is the total available power constraint.

3. Proposed Methodology

3.1. The Exact Solution and the Supervised Approach

The Lagrangian function to maximize is [30]
L ( P l ( h l ) , υ ) 0 P l P s = l = 1 L Ε [ C l ( P l ( h l ) , h l ) ] υ ( l = 1 L E [ P l ( h l ) ] P a v )
where υ is the dual variable.
If the statistical models f h l h l are available, the average channel capacity is then [11,12]:
Ε [ C l ( P l ( h l ) , h l ) ] = 0 log 2 ( 1 + P l ( h l ) | h l | 2 N 0 ) f h l ( h l ) d h l
and the average transmitted power is then [11,12]:
E [ P l ( h l ) ] = 0 P l ( h l ) f h l ( h l ) d h l
The PA problem (3) and (4) is then convex [30]. The instantaneous optimal PA for the ith channel is given in [11,12]:
P l * ( h l ) = min { P s , max ( 1 υ * N 0 h l , 0 ) }
where υ * is the optimal dual multiplier, and N 0 is the optical noise variance.
The υ * is evaluated numerically from (4), (8) as follows:
l = 1 L 0 P l * ( h l ) f h l ( h l ) d h l = P a v
or via an iterative algorithm, e.g., sub-gradient method with decreasing step size η as below [30]:
υ k + 1 = [ υ k η k ( P a v Ε [ l = 1 L P l ( h l ) ] ) ] + η k + 1 = η k / ( k + 1 )
Note that the analytical expression (8) is valid as long as the N 0 is constant and not a function of P l h l . Also, (8) requires precise h l , and (9) demands precise channel PDF knowledge to calculate υ * . In practice, it is very hard to assume accurate model information, CSI values, and constant optical noise.
Supervised learning is a process of providing labeled input datasets to the learning model to make predictions (regression) or decisions (classification) based on the labeled samples. In a regression problem, the output node is continuous, ranging theoretically from negative infinity to positive infinity. The mean square error (MSE) is the most common loss function to employ. From (8) and (10), the exact solution is provided (no need for a brute-force or exhaustive search), which can be utilized as ground truth to train a DNN to predict the optimal power allocation. Non-linear regression fitting or a multi-layer perceptron model may be implemented. However, this supervised approach is model-based.

3.2. The Proposed DRL-Aided Algorithm

DNNs consist of several layers of interconnected neurons and enable the learning of complex non-linear functions. RL teaches an agent how to execute actions upon entering a new state to maximize the cumulative future returns [23,24,25]. A DRL methodology refines its policy by directly calculating the policy gradient to maximize the cumulative rewards. The referred policy is a DNN that uses a continuous state as input and produces a probability distribution as output. A continuous action is then sampled. The core idea is to parameterize the policy using a DNN’s weights parameter θ w where w = i = 1 Q w i w i + 1 assuming a DNN with Q layers, and each has a corresponding dimension w i [23,24,25].
Then, we denote the parameterized policy as the probability distribution of selecting power P t , l for the lth channel in time-step t, and from the state h t , l , θ [13,14]:
π ( h , θ ) = Pr [ P t , l | h t , l , θ ]
where h t , l represents recorded CSI data of lth channel, and P t , l 0 , P s is the allocated power as derived from π h , θ .
The agent’s reward-return r t + 1 at the next time-step t + 1 is the following expression [13,14]:
r t + 1 = L ( θ , υ ) 0 π ( h t , l , θ ) P s = l = 1 L C t , l ( π ( h t , l , θ ) , h t , l ) υ ( l = 1 L π ( h t , l , θ ) P a v )
Then, the discounted cumulative returns G t are defined as
G t = r t + 1 + γ r t + 2 + γ 2 r t + 3 + = t = t + 1 T γ t t 1 r t
where T is the duration of an episode, t = 0…T − 1, and γ ∈ [0, 1) is the discount factor [20,21,22]. The objective function is the long-term discounted rewards [13,14]:
J ( θ ) = Ε [ G t | π ( h t , θ ) ]
In general, the gradient for a whole episode with duration T is
θ J ( θ ) = T θ π θ ( T ) G ( T ) d T
If we apply the log-derivative trick θ π θ = π θ θ log π θ [13,14], then:
θ J ( θ ) = Ε [ θ log π θ ( T ) G ( T ) ]
Then, we obtain a Monte-Carlo approximation of E   . in (16) for all the t = 0, 1, …, T − 1 time-steps and l = 0, 1, …, L − 1 channels [31]:
θ J ( θ ) = 1 T t = 0 T 1 l = 0 L 1 [ θ log π ( h t , l , θ ) G t ]
To summarize, the proposed model-free, DRL methodology requires only h l samples that can be observed, C l values that can be experimentally measured, and a differentiable policy to be optimized. The proposed methodology is summarized in Algorithm 1.
Algorithm 1. DRL-aided Proposed Power Allocation Methodology
1:Initialize number of channels L, number of episodes Ep, episode duration T, learning rate a, discount factor γ, and weights θ of the policy network;
2: for   i   0 to Ep − 1 do
3: Set   θ J ( θ ) = 0 ;
4: for   t   0 to T − 1 do
5: for   l   0 to L − 1 do
6: Fetch   recorded   CSI   data { h t , l } for each tth time-step and lth optical channel;
7: Sample   the   allocated   power { π ( h t , l , θ ) } using the current policy distribution;
8: Calculate   the   returns { r t + 1 , l } from (12);
9: Calculate   G t , l = t = t + 1 T γ t t 1 r t , l from (13);
10: θ J ( θ ) θ J ( θ ) + θ log π ( h t , l , θ ) G t , l from (17);
11: end
12: end
13: θ θ + α θ J ( θ ) / T gradient ascent the policy parameter (DNN’s weights);
14: υ [ υ α t = 0 T 1 { l = 0 L 1 π ( h t , l , θ ) P a v } / T ] + update dual variable (total power constraint);
15: end

4. Simulation Results

In this section, the DRL-aided proposed PA methodology is applied to the experimental ARTEMIS time series data along with other heuristic, model-free strategies, i.e., DQN, Equal Power, and Random Power algorithms and the model-based supervised method.
The ESA’s data relay satellite mission ARTEMIS (Advanced Relay and Technology Mission Satellite) took place in 2003 when the spacecraft attained GEO orbit, and involved a variety of optical communication telescopes for satellite-to-ground and intersatellite bi-directional links [27,28]. The experimental data to be utilized are recorded time series from the downlink transactions from the laser terminal (OPALE) onboard ARTEMIS to the reflector telescope (LUCE) installed on the Teide Observatory in Tenerife, 2400 m above sea level [27,28]. The spacecraft’s location and communication features as well as LUCE’s and OGS’s essential characteristics are given in Table 1.
In Figure 1, the normalized PDFs of five experimental time series are illustrated vs. the normalized PDFs of synthesized data that are lognormally distributed [27,28]. It is obvious that the PDFs derived from the retrieved ARTEMIS data accurately fit the lognormal PDFs, implying weak turbulence conditions, and that the channel model in (2) can be applied to find the exact solution which will be used as labeled data for the supervised approach, elaborated in Section 3.1.
The channel samples are then employed to form a 5 × 5 diagonal Multi-Input Multi-Output (MIMO) optical satellite-to-ground system. In general, the channels are assumed cloud-free but suffer from propagation losses and turbulence-induced scintillation.
In Table 2, the experimental optical channels’ mean powers and SIs are included. The lower the P r and the higher the SI, the worse the slant path conditions. The channels’ correlation is negligible; hence, they are safely considered independent and will be denoted as ch.0, ch.1, ch.2, ch.3, and ch.4.
To develop the proposed DRL-aided power allocation algorithm, the policy’s DNN had to be accurately specified. In particular, the neural network’s layers, nodes, and probability distribution must be chosen in a way to avoid overfitting, overhead, and high inference time. In Figure 2, the results of an investigation of the optimal number of hidden layers and neurons are shown. From Figure 2a, it is seen that one or two hidden layers are sufficient as they achieve the maximum average capacity and loss function. From Figure 2b, it is concluded that 256 neurons yield the best performance.
Finally, the stochastic policy that will determine the agents’ actions was investigated. In Figure 3, three different probability distributions, i.e., truncated Normal ( μ , σ , 0 , P s ) , truncated Weibull ( k , λ , 0 , P s ) , and truncated Exponential ( λ , 0 , P s ) were tested and evaluated. Normal and Weibull are two-parameter distributions while Exponential is one-parameter. It is observed that the two-parameter distributions are the best choice and that truncated Normal performs slightly better.
The finalized policy structure consists of a four-layer DNN, as reported in Table 3. In addition, the truncated Normal (TN) distribution is selected as a stochastic power policy for the five optical channels.
The output nodes are used as parameters of the five TN distributions—means and standard deviations. The optimization algorithm involved in the training is RMSProp with step learning rate scheduling. Specifically, 5000 episodes with a duration of T = 100 time-steps are considered to have frequent weight updates. The initial learning rate is set to 0.001 and the discount factor to 0.5 because fading negatively affects the impact of an agent’s actions on his future expected returns [20]. Also, given the total power constraint, the agents’ behaviors affect others’ rewards due to the unpredictability of their neighbors’ actions. Higher γ is also undesirable because it decelerates the response to channel fluctuations [20]. The proposed algorithm was implemented in PyTorch.
In Figure 4, the predicted average capacity (a) and the constraint function (b) l = 0 L 1 E P l h l P a v are depicted for P a v = 2.0   W . In Figure 5, the average channel powers are illustrated for the proposed, the supervised, the DQN, the Equal Power, and the Random Power algorithms for P a v = 2.0   W . Likewise, in Figure 6, the predicted average capacity (a) and the constraint function (b) are depicted for P a v = 3.0   W , and in Figure 7, the average channel powers are illustrated for P a v = 3.0   W . Finally, in Figure 8, the predicted average capacity (a) and the constraint function (b) are depicted for P a v = 4.0   W , and in Figure 9, the average channel powers are illustrated for P a v = 4.0   W . The DQN is a model-free, off-policy method that tries to predict the Q-values (expected future rewards) for every state–action pair using the ϵ-greedy policy that alternates between exploration and exploitation [20,21,22]. DQN employs discretized actions: { 0 , P ¯ / A 1 , 2 P ¯ / A 1 , , P ¯ } where A is the number of actions [20,21,22]. Here, A = 21 was chosen.
The supervised algorithm was trained using the exact solution taken from the water-filling algorithm as if the channel statistics had been known to the transmitter. That scenario is expressed in Section 3.1 and incorporates (2), (8), and (10).
In Figure 4, Figure 6 and Figure 8, we observe that the proposed model-free DRL-aided solution outperforms the three baseline algorithms, showing a predicted average capacity advantage (18% better than Equal Power, 16% better than DQN, and 53% better than Random Power for P a v = 2.0   W ). Additionally, it approximates the supervised method with excellent accuracy (2% worse performance for P a v = 2.0   W ). The episode duration of T = 100 steps causes some waveform fluctuation, which can be resolved by using more time-steps at the cost of inference time. In Figure 5, Figure 7 and Figure 9, the average channel powers of the proposed algorithm match very well with the supervised ones, i.e., more power is assigned to the channels with better conditions, only differing by 0.1–0.2 W. The allocated channel powers of the other schemes differ significantly more, about 0.1–0.5 W. The inference time of an episode is 0.7–0.8 s including recorded data sampling, cumulative reward calculation, gradient computation, and back-propagation.
In Figure 10, the means and standard deviations of the TN distributions for ch.0 and ch.3 are depicted. Gradually, the values are stabilized according to their conditions in Table 2. The mean value of ch.0 (~1.2) is greater than that of ch.3 (~0.85), and the standard deviation of ch.0 (~0.2) is smaller than that of ch.3 (~0.5). Thus, ch.0 has a much higher probability of being allocated with P s = 1.0   W than the weaker ch.3.
Finally, in Figure 11 the loss function is plotted for a training (two-thirds of original dataset) and a validation set (one-third of original dataset), and for two fixed hyperparameter λ values. The λ can be selected to adjust the trade-off between the power constraint violation and the loss function. The training loss is on the same level as the validation loss; therefore, no overfitting is observed.

5. Conclusions

In this paper, several channel-model-free methodologies and heuristics are explored for optimal PA and then compared to the unparameterized solution for an all-FSO, multi-aperture satellite downlink between a GEO satellite and an OGS. A DRL-aided power allocation methodology is proposed for optical satellite systems disregarding any channel statistics knowledge requirements. Therefore, the PA problem is formulated as a constrained learning optimization problem with peak and total expected power inequality constraints. A power allocation algorithm was developed, aided by a DNN which is fed CSI observations and trained in a parameterized on-policy manner through a stochastic policy gradient approach. The proposed method does not require the channels’ transition models or fading distributions. To validate and test the proposed allocation scheme, experimental measurements from the ARTEMIS optical satellite campaign were utilized. The proposed scheme performs 18% better than Equal Power, 16% better than DQN, and 53% better than Random Power, and it approximates the supervised method, with only 2% less accuracy, for P a v = 2.0   W . The predicted average channel powers match very well with the supervised ones, only differing by 0.1–0.2 W, while the allocated channel powers of the other schemes differ significantly more, about 0.1–0.5 W. Two hidden layers, 256 neurons, and two-parameter distributions are the optimal choices for the policy DNN. Finally, the proposed PA strategy is deemed appropriate for difficult optical conditions since it learns explicitly from observation.

Author Contributions

Conceptualization, T.T.K., N.K.L. and A.D.P.; methodology, T.T.K. and N.K.L.; software, T.T.K.; validation, T.T.K.; formal analysis, T.T.K.; investigation, T.T.K.; resources, N.K.L. and A.D.P.; writing—original draft preparation, T.T.K. and N.K.L.; writing—review and editing, A.D.P.; visualization, T.T.K.; supervision, A.D.P. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Acknowledgments

The work presented in this paper was carried out under the ONSET-CCN ESA project.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Sharma, S.K.; Chatzinotas, S.; Arapoglou, P.D. Satellite Communications in the 5G Era; Institution of Engineering and Technology: London, UK, 2018; pp. 1–500. [Google Scholar] [CrossRef]
  2. Jahid, A.; Alsharif, M.H.; Hall, T.J. A contemporary survey on Free Space Optical Communication: Potentials, technical challenges, recent advances and research direction. J. Net. Comput. Appl. 2022, 200, 103311. [Google Scholar] [CrossRef]
  3. Kaushal, H.; Kaddoum, G. Optical Communication in Space: Challenges and Mitigation Techniques. IEEE Commun. Surv. Tutor. 2016, 19, 57–96. [Google Scholar] [CrossRef]
  4. Hemmati, H. Near-Earth Laser Communications; CRC Press: Boca Raton, FL, USA, 2009; pp. 1–418. [Google Scholar] [CrossRef]
  5. Andrews, L.C.; Phillips, R.L. Laser Beam Propagation through Random Media; SPIE Optical Engineering Press: Bellingham, WA, USA, 2005; pp. 1–820. [Google Scholar] [CrossRef]
  6. Majumdar, A.K. Laser Communication with Constellation Satellites, UAVs, HAPs and Balloons: Fundamentals and Systems Analysis for Global Connectivity; Springer International Publishing: Cham, Switzerland, 2022; pp. 1–270. [Google Scholar] [CrossRef]
  7. Kodheli, O.; Lagunas, E.; Maturo, N.; Sharma, S.K.; Shankar, B.; Montoya, J.F.M.; Duncan, J.C.M.; Spano, D.; Chatzinotas, S.; Kisseleff, S.; et al. Satellite Communications in the New Space Era: A Survey and Future Challenges. IEEE Commun. Surv. Tutor. 2021, 23, 70–109. [Google Scholar] [CrossRef]
  8. Lyras, N.K.; Kourogiorgas, C.I.; Kapsis, T.T.; Panagopoulos, A.D. On Optical Satellite Communication Systems Design: CFLOS Calculation and OGS Selection. In Proceedings of the 13th European Conference on Antennas and Propagation (EuCAP), Krakow, Poland, 31 March–5 April 2019; pp. 1–5. [Google Scholar]
  9. Lyras, N.K.; Kourogiorgas, C.I.; Panagopoulos, A.D. Cloud free line of sight prediction modeling for optical satellite communication networks. IEEE Commun. Lett. 2017, 21, 1537–1540. [Google Scholar] [CrossRef]
  10. Kapsis, T.T.; Lyras, N.K.; Kourogiorgas, C.I.; Panagopoulos, A.D. Time Series Irradiance Synthesizer for Optical GEO Satellite Downlinks in 5G Networks. Future Internet 2019, 11, 131. [Google Scholar] [CrossRef]
  11. Kapsis, T.T.; Panagopoulos, A.D. Optimum power allocation based on channel conditions in optical satellite downlinks. Wirel. Pers. Commun. 2021, 116, 2997–3013. [Google Scholar] [CrossRef]
  12. Kapsis, T.T.; Panagopoulos, A.D. Robust power allocation in optical satellite MIMO links with pointing jitter. IEEE Wirel. Commun. Lett. 2022, 11, 957–961. [Google Scholar] [CrossRef]
  13. Gao, Z.; Eisen, M.; Ribeiro, A. Resource Allocation via Model-Free Deep Learning in Free Space Optical Communications. IEEE Trans. Commun. 2022, 70, 920–934. [Google Scholar] [CrossRef]
  14. Eisen, M.; Zhang, C.; Chamon, L.F.; Lee, D.D.; Ribeiro, A. Learning optimal resource allocations in wireless systems. IEEE Trans. Signal Process 2019, 67, 2775–2790. [Google Scholar] [CrossRef]
  15. Li, Y.; Geng, T.; Tian, R.; Gao, S. Power allocation in a spatial multiplexing free-space optical system with reinforcement learning. Opt. Commun. 2021, 488, 126856. [Google Scholar] [CrossRef]
  16. Lionis, A.; Peppas, K.; Nistazakis, H.E.; Tsigopoulos, A.D.; Cohn, K. Experimental Performance Analysis of an Optical Communication Channel over Maritime Environment. Electronics 2020, 9, 1109. [Google Scholar] [CrossRef]
  17. Lionis, A.; Sklavounos, A.; Stassinakis, A.; Cohn, K.; Tsigopoulos, A.; Peppas, K.; Aidinis, K.; Nistazakis, H. Experimental Machine Learning Approach for Optical Turbulence and FSO Outage Performance Modeling. Electronics 2023, 12, 506. [Google Scholar] [CrossRef]
  18. Sharma, S.; Yoon, W. Energy Efficient Power Allocation in Massive MIMO Based on Parameterized Deep DQN. Electronics 2023, 12, 4517. [Google Scholar] [CrossRef]
  19. Ortiz-Gomez, F.G.; Lei, L.; Lagunas, E.; Martinez, R.; Tarchi, D.; Querol, J.; Salas-Natera, M.A.; Chatzinotas, S. Machine Learning for Radio Resource Management in Multibeam GEO Satellite Systems. Electronics 2022, 11, 992. [Google Scholar] [CrossRef]
  20. Nasir, Y.S.; Guo, D. Multi-agent deep reinforcement learning for dynamic power allocation in wireless networks. IEEE J. Sel. Areas Commun. 2019, 37, 2239–2250. [Google Scholar] [CrossRef]
  21. Cho, Y.; Yang, W.; Oh, D.; Jo, H.S. Multi-Agent Deep Reinforcement Learning for Interference-Aware Channel Allocation in Non-Terrestrial Networks. IEEE Commun. Lett. 2023, 27, 936–940. [Google Scholar] [CrossRef]
  22. Hu, X.; Liu, S.; Chen, R.; Wang, W.; Wang, C. A deep reinforcement learning-based framework for dynamic resource allocation in multibeam satellite systems. IEEE Commun. Lett. 2018, 22, 1612–1615. [Google Scholar] [CrossRef]
  23. O’Shea, T.; Hoydis, J. An introduction to deep learning for the physical layer. IEEE Trans. Cogn. Commun. Netw. 2017, 3, 563–575. [Google Scholar] [CrossRef]
  24. Arulkumaran, K.; Deisenroth, M.P.; Brundage, M.; Bharath, A.A. Deep reinforcement learning: A brief survey. IEEE Signal Process Mag. 2017, 34, 26–38. [Google Scholar] [CrossRef]
  25. Alom, M.Z.; Taha, T.M.; Yakopcic, C.; Westberg, S.; Sidike, P.; Nasrin, M.S.; Hasan, M.; Van Essen, B.C.; Awwal, A.A.S.; Asari, V.K. A State-of-the-Art Survey on Deep Learning Theory and Architectures. Electronics 2019, 8, 292. [Google Scholar] [CrossRef]
  26. Williams, R.J. Simple statistical gradient-following algorithms for connectionist reinforcement learning. Mach. Learn. 1992, 8, 229–256. [Google Scholar] [CrossRef]
  27. Reyes, M.; Comeron, A.; Alonso, A.; Rodriguez, A.; Rubio, J.A.; Dios, V.F.; Chueca, S.; Sodnik, Z. Ground-to-satellite bidirectional laser links for validation of atmospheric turbulence model. In Proceedings of the Free-Space Laser Communication and Active Laser Illumination III, San Diego, CA, USA, 3–8 August 2003; Volume 5160, pp. 44–55. [Google Scholar] [CrossRef]
  28. Alonso, A.; Reyes, M.; Sodnik, Z. Performance of satellite-to-ground communications link between ARTEMIS and the optical ground station. In Proceedings of the Optics in Atmospheric Propagation and Adaptive Systems VII, Canary Islands, Spain, 13–16 September 2004; Volume 5572, pp. 372–383. [Google Scholar] [CrossRef]
  29. Kolev, D.R.; Toyoshima, M. Satellite-to-ground optical communications using small optical transponder (SOTA)—Received-power fluctuations. Opt. Express 2017, 25, 28319–28329. [Google Scholar] [CrossRef]
  30. Bertsekas, D.P. Convex Optimization Theory; Athena Scientific Belmont: Belmont, MA, USA, 2009. [Google Scholar]
  31. Hastings, W.K.; Mardia, K.V. Monte Carlo Sampling Methods Using Markov Chains and Their Applications. Biometrika 1970, 57, 97–109. [Google Scholar] [CrossRef]
Figure 1. Normalized PDFs of experimental and synthesized data. (a) 10 September 2003 20:10–20:30. (b) 10 September 2003 00:30–00:50. (c) 12 September 2003 00:30–00:50. (d) 13 September 2003 23:30–23:50. (e) 16 September 2003 20:10–20:30.
Figure 1. Normalized PDFs of experimental and synthesized data. (a) 10 September 2003 20:10–20:30. (b) 10 September 2003 00:30–00:50. (c) 12 September 2003 00:30–00:50. (d) 13 September 2003 23:30–23:50. (e) 16 September 2003 20:10–20:30.
Electronics 13 00647 g001
Figure 2. The loss function and predicted average capacity in terms of the (a) number of hidden layers (200 nodes/layer) and (b) number of neurons (2 hidden layers). P s = 1.0   W ,   P a v = 3.0   W .
Figure 2. The loss function and predicted average capacity in terms of the (a) number of hidden layers (200 nodes/layer) and (b) number of neurons (2 hidden layers). P s = 1.0   W ,   P a v = 3.0   W .
Electronics 13 00647 g002
Figure 3. Performance results for five optical satellite downlinks using three different policy distributions: truncated Normal ( μ , σ , 0 , P s ) , truncated Weibull ( k , λ , 0 , P s ) , and truncated Exponential ( λ , 0 , P s ) . P s = 1.0   W ,   P a v = 2.5   W .
Figure 3. Performance results for five optical satellite downlinks using three different policy distributions: truncated Normal ( μ , σ , 0 , P s ) , truncated Weibull ( k , λ , 0 , P s ) , and truncated Exponential ( λ , 0 , P s ) . P s = 1.0   W ,   P a v = 2.5   W .
Electronics 13 00647 g003
Figure 4. Learning episodes of five different PA algorithms for five optical channels: (a) the average system capacity and (b) the constraint function. P s = 1.0   W ,   P a v = 2.0   W .
Figure 4. Learning episodes of five different PA algorithms for five optical channels: (a) the average system capacity and (b) the constraint function. P s = 1.0   W ,   P a v = 2.0   W .
Electronics 13 00647 g004
Figure 5. The average channel powers of five different PA algorithms for five optical channels during 5000 episodes. P s = 1.0   W ,   P a v = 2.0   W .
Figure 5. The average channel powers of five different PA algorithms for five optical channels during 5000 episodes. P s = 1.0   W ,   P a v = 2.0   W .
Electronics 13 00647 g005
Figure 6. Learning episodes of five different PA algorithms for five optical channels: (a) the average system capacity and (b) the constraint function. P s = 1.0   W ,   P a v = 3.0   W .
Figure 6. Learning episodes of five different PA algorithms for five optical channels: (a) the average system capacity and (b) the constraint function. P s = 1.0   W ,   P a v = 3.0   W .
Electronics 13 00647 g006
Figure 7. The average channel powers of five different PA algorithms for five optical channels during 5000 episodes. P s = 1.0   W ,   P a v = 3.0   W .
Figure 7. The average channel powers of five different PA algorithms for five optical channels during 5000 episodes. P s = 1.0   W ,   P a v = 3.0   W .
Electronics 13 00647 g007
Figure 8. Learning episodes of five different PA algorithms for five optical channels: (a) the average system capacity and (b) the constraint function. P s = 1.0   W ,   P a v = 4.0   W .
Figure 8. Learning episodes of five different PA algorithms for five optical channels: (a) the average system capacity and (b) the constraint function. P s = 1.0   W ,   P a v = 4.0   W .
Electronics 13 00647 g008
Figure 9. The average channel powers of five different PA algorithms for five optical channels during 5000 episodes. P s = 1.0   W ,   P a v = 4.0   W .
Figure 9. The average channel powers of five different PA algorithms for five optical channels during 5000 episodes. P s = 1.0   W ,   P a v = 4.0   W .
Electronics 13 00647 g009
Figure 10. The mean and standard deviation of the TN distributions for two allocated channel powers corresponding to ch.0 and ch.3.
Figure 10. The mean and standard deviation of the TN distributions for two allocated channel powers corresponding to ch.0 and ch.3.
Electronics 13 00647 g010
Figure 11. Loss function across the training and validation sets for two hyperparameter λ values related to the total power constraint function. P s = 1.0   W ,   P a v = 3.0   W . The relaxed λ = 0.2 allows the power constraint to be violated, resulting in higher loss values, while the stricter λ = 1.0 yields lower loss values. No overfitting is observed.
Figure 11. Loss function across the training and validation sets for two hyperparameter λ values related to the total power constraint function. P s = 1.0   W ,   P a v = 3.0   W . The relaxed λ = 0.2 allows the power constraint to be violated, resulting in higher loss values, while the stricter λ = 1.0 yields lower loss values. No overfitting is observed.
Electronics 13 00647 g011
Table 1. ARTEMIS, OPALE, LUCE, and OGS Characteristics.
Table 1. ARTEMIS, OPALE, LUCE, and OGS Characteristics.
NameCharacteristicValue
ARTEMISLongitude21.5° East
OPALELatitude0.0° ± 2.81° North
Altitude35,787 km
Elevation Angle37°
CoverageEurope, Africa, and the Middle East
Wavelength819 nm
Beam Diameter (1/e2)125 mm
Transmitted Power10 mW
Modulation SchemeIntensity Modulation—Direct Detection
Data Rate2 Mbps (downlink), 50 Mbps (uplink)
LUCEAperture Diameter26 cm
OGSAltitude2.4 km
Table 2. Optical Channels’ Parameters.
Table 2. Optical Channels’ Parameters.
ChannelExperimentParametersValues
010/09/2003 20:10 P r ,   σ I 2 −13.34 dBm, 0.0101
110/09/2003 00:30 P r ,   σ I 2 −15.79 dBm, 0.0142
212/09/2003 00:30 P r ,   σ I 2 −19.78 dBm, 0.0274
313/09/2003 23:30 P r ,   σ I 2 −20.94 dBm, 0.0485
416/09/2003 20:10 P r ,   σ I 2 −13.85 dBm, 0.0121
Table 3. Policy DNN Architecture.
Table 3. Policy DNN Architecture.
Channel# NodesActivation
Input5Linear
Hidden #1200ReLu
Hidden #2100ReLu
Output10Softplus
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

Kapsis, T.T.; Lyras, N.K.; Panagopoulos, A.D. Optimal Power Allocation in Optical GEO Satellite Downlinks Using Model-Free Deep Learning Algorithms. Electronics 2024, 13, 647. https://doi.org/10.3390/electronics13030647

AMA Style

Kapsis TT, Lyras NK, Panagopoulos AD. Optimal Power Allocation in Optical GEO Satellite Downlinks Using Model-Free Deep Learning Algorithms. Electronics. 2024; 13(3):647. https://doi.org/10.3390/electronics13030647

Chicago/Turabian Style

Kapsis, Theodore T., Nikolaos K. Lyras, and Athanasios D. Panagopoulos. 2024. "Optimal Power Allocation in Optical GEO Satellite Downlinks Using Model-Free Deep Learning Algorithms" Electronics 13, no. 3: 647. https://doi.org/10.3390/electronics13030647

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