Content uploaded by Pablo Miralles
Author content
All content in this area was uploaded by Pablo Miralles on Nov 22, 2021
Content may be subject to copyright.
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
IAC-21-B.1.4.4
Machine Learning in Earth Observation Operations: A review
Pablo Miralles*, Antonio Fulvio Scannapieco*, Nitya Jagadam*, Prerna Baranwal*, Bhavin Faldu*, Ruchita
Abhang*, Sahil Bhatia*, Sebastien Bonnart*, Ishita Bhatnagar*, Beenish Batul*, Pallavi Prasad*, Héctor
Ortega-González*, Harrish Joseph*, Harshal More*, Sondes Morchedi*, Aman Kumar Panda*, Marco
Zaccaria Di Fraia*, Daniel Wischert*, Daria Stepanova*
(*) In respective order:
GTD International, GTD Group, 99 Route d’Espagne, 31100, Toulouse, France, pmirallesr@gmail.com
Space Generation Advisory Council (SGAC), c/o European Space Policy Institute, Schwarzenbergplatz 6, 1030
Vienna, Austria, antonio.scannapieco@spacegeneration.org
Space Generation Advisory Council (SGAC), c/o European Space Policy Institute, Schwarzenbergplatz 6, 1030
Vienna, Austria, jagadamnitya@gmail.com
Department of Electrical & Electronics/Instrumentation & Department of Mathematics, Birla Institute of Technology
and Science (BITS), Pilani, Rajasthan - 333031, India, f2016568@pilani.bits-pilani.ac.in
Space Generation Advisory Council (SGAC), c/o European Space Policy Institute, Schwarzenbergplatz 6, 1030
Vienna, Austria, bhavinfaldu6474@gmail.com
Department of Computer Science Engineering, Savitribai Phule Pune University, Pune, Maharashtra - 411052,
India, rabhang09@gmail.com
Department of Aerospace Engineering, University of Petroleum and Energy Studies, Energy Acres, Bidholi via
Premnagar, Dehradun, Uttarakhand - 248007, India, sahil2112.b@gmail.com
Space Generation Advisory Council (SGAC), c/o European Space Policy Institute, Schwarzenbergplatz 6, 1030
Vienna, Austria, sgac@sbonnart.fr
Department of Chemical Engineering & Department of Physics, Birla Institute of Technology and Science (BITS),
Pilani, Rajasthan - 333031, India, f20180306@pilani.bits-pilani.ac.in
Space Generation Advisory Council (SGAC), c/o European Space Policy Institute, Schwarzenbergplatz 6, 1030
Vienna, Austria, 23.prasad@gmail.com
Space Generation Advisory Council (SGAC), c/o European Space Policy Institute, Schwarzenbergplatz 6, 1030
Vienna, Austria, ortegagonzalez.hector@gmail.com
Sapienza University of Rome, Via Eudossiana 18, 00184, Rome, Italy, harrishjagan@gmail.com
Sapienza University of Rome, Via Eudossiana 18, 00184, Rome, Italy, harshal.more@spacegeneration.org
Space Generation Advisory Council (SGAC), c/o European Space Policy Institute, Schwarzenbergplatz 6, 1030
Vienna, Austria, sondesmorchedi@gmail.com
ISAE-École Nationale Supérieure de Mécanique et d’Aérotechnique, Téléport 2 – 1 avenue Clément Ader BP 40109
86961 Futuroscope Chasseneuil Cedex, Poitiers, France. aman.panda143@gmail.com
Centre for Electronic Warfare Information and Cyber, Cranfield University, The Defence Academy Of The UK,
Shrivenham, SN6 8LA, UK, marco.di-fraia@cranfield.ac.uk
Space Generation Advisory Council (SGAC), c/o European Space Policy Institute, Schwarzenbergplatz 6, 1030
Vienna, Austria, daniel.wischert@spacegeneration.org
German Orbital Systems GmbH, Manfred-Fuchs-Platz 2-4, 28359 Bremen, Germany,
daria.stepanova@skolkovotech.ru
Abstract
Analysis of downlinked satellite imagery has undeniably benefited greatly from the ongoing Machine Learning
(ML) revolution. Other aspects of the Earth Observation industry, despite being less prone to an extensive application
of ML, are also following this trend. This work aims at presenting - in the form of a review of Machine Learning
applied to Earth Observation Operations - such applications, the existing use cases, potential opportunities and
pitfalls, and perceived gaps in research. A wide range of topics are discussed including mission planning, diagnosis,
prognosis, and repair of faults, optimization of telecommunications, enhanced GNC, on-board image processing, and
usage of Machine Learning models within platforms with limited compute and power capabilities. The review
tackles all on-board and off-board applications of machine learning to Earth Observation with one notable exception:
it omits all post-processing of payload data on the ground, a topic that has been studied extensively by past authors.
IAC-21-B.1.4.4 Page 1 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
This research was produced by a team of volunteers from the Small Satellite Project Group of the Space Generation
Advisory Council.
Keywords: Machine Learning; Artificial Intelligence; Earth Observation; Remote Sensing; Neural Networks;
Computing
Acronyms / Abbreviations
Artificial Intelligence (AI); Machine Learning (ML);
Deep Learning (DL); Fault Detection, Isolation and
Recovery (FDIR); Guidance, Navigation and Control
(GNC); Neural Network (NN); Convolutional Neural
Network (CNN); Deep Neural Network (DNN);
Artificial Neural Network (ANN); Binarized Neural
Network (BNN); Bayesian Network (BN); Dynamic
Bayesian Network (DBN); National Aeronautics and
Space Administration (NASA); European Space
Agency (ESA); On-Board Computer (OBC); Earth
Observation (EO); Random Decision Forest (RDF);
Bayesian Thresholding (BT); Support Vector Machine
(SVM); Commercial off-the-shelf (COTS); Size,
Weight and Power (SWaP); Light Detection and
Ranging (LIDAR); System on a Chip (SoC); False
Positives (FP); Consultative Committee for Space Data
Systems (CCSDS); Global Navigation Satellite System
(GNSS); Global Positioning System (GPS);
Proportional - Integral (PI); Proportional - Integral -
Derivative (PID); Attitude Orbital Determination
System (AODS); Reinforcement Learning (RL);
Extended Kalman Filter (EKF); Random Forest (RF);
Attitude and Orbit Control System (AOCS); k-Nearest
Neighbour (k-NN); Self-Organizing Map (SOM); On
Orbit Servicing (OOS); Anomaly Resolution and
Prognostic Health Management for Autonomy
(ARPHA); Density-Based Spatial Clustering of
Applications with Noise (DBSCAN); Electrical Power
System (EPS); Software and Sensor Health
Management (SSHM); Regularized Discriminant
Analysis (RDA); Adaptive Regularization of Weight
Vector (AROW); Soft Confidence-Weighted (SCW);
Centre national d'études spatiales (The National Centre
for Space Studies) (CNES); European Space Operations
Centre (ESOC); Radial Basis Function (RBF);
Deutsches Zentrum für Luft- und Raumfahrt (German
Aerospace Center) (DLR); One-Class Support Vector
Machine (OC-SVM); Normal Gaussian Herding
(NHERD); Thermal EMission Imaging System
(THEMIS); Intelligent Payload EXperiment (IPEX);
Hyperspectral Infrared Imager (HyspIRI);
Moderate-Resolution Imaging Spectroradiometer
(MODIS); Peak Signal to Noise Ratio (PSNR);
Structural Similarity Index (SSIM); Field
Programmable Gate Array (FPGA); Context-Based,
Adaptive, Lossless Image Codec (CALIC); Integral
Wavelet Transform (IWT); Peano-Hilbert (PH);
Learning Vector Quantization (LVQ); TensorFlow (TF);
Synthetic Aperture Radar (SAR); Ratio of Exponential
Weighted Average (ROEWA); Joint Photographic
Experts Group (JPEG); Space Test
Program-Houston-5-Cubesat Service protocol
(STP-H5-CSP); Neural Architecture Structure (NAS);
Central Processing Unit (CPU); Graphics Processing
Unit (GPU); Visual Processing Unit (VPU); Time
Processing Unit (TPU); Trained Ternary Quantization
(TTQ); Radiation Tolerant (RT); Mobile Neural
Architecture Search (MNAS); Knowledge Transfer
(KT); Knowledge Distillation (KD); Cubesat Service
Protocol (CSP); SpaceBorne Computer (SBC);
Modified National Institute of Standards and
Technology database (MNIST); National Information
Security Standardization Technical Committee
(NISSTC); Deutsches Institut für Normung (German
institute of Standardization) (DIN); Deutschen
Kommission Elektrotechnik Elektronik
Informationstechnik (German Commission for
Electrical, Electronic and Information Technologies)
(DKE); European Union (EU); European commission
(EC); Small and Medium Enterprises (SMEs);
DEpendable and Expandable Learning (DEEL); Earth
Observation Systems’ Data Information Systems
(EOSDIS); Space Generation Advisory Council
(SGAC); Small Satellite Project Group (SSPG); Deep
Reinforcement Learning (DRL); Geographic
Information Systems (GIS); Explainable AI (XAI).
Figures
Fig 1) On-board cloud segmentation from the CIAR
project.
Table
Table 1: Hardware Accelerators.
1. Introduction
Earth Observation (EO) satellites have allowed us to
look at our planet at a scale previously unattainable to
humankind. From the vantage point of space, it
becomes easier to monitor everything about our lives,
from the very large scale such as our impact on the
planet’s ecology [1] or the dynamics of the shifting
IAC-21-B.1.4.4 Page 2 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
Earth atmosphere[2] to the very small like the nature
and extent of specific facilities all around the world [3].
This capability has been and continues to be invaluable
to understanding the world around us and enforcing
regulations vital to the wellbeing of people all over the
globe.
However, as access to space becomes ever more
affordable, EO assets multiply at an increasingly faster
pace [4]. Moreover EO Operations - the sequence of
activities that take place in managing an EO spacecraft
from its launch to its demise - keep growing in number
and complexity as new assets are put into orbit. These
trends could soon lead to a situation where available
work-power becomes a limiting factor in the
deployment of EO systems. Orchestrating these
operations is, at its core, a control and data processing
problem - from taking in and analyzing large volumes
of telemetry from all EO platforms, to taking into
account their complex dynamics and evolving mission
profiles when utilizing them.
Machine Learning (ML) has taken the data
processing world by storm with one success story after
another. From object detection and classification [5] to
natural language processing [6] and nonlinear control
[7], the capacity of these algorithms to solve different
types of problems has been nothing short of
awe-inspiring.
For the purposes of the present review, we define
ML algorithms as those whose performance critically
depends and generally improves with exposure to
real-world data of the problem to be solved.
So far, these techniques have concentrated mostly on
analysis of downlinked imagery, due to the larger
availability of computing power and ease of deployment
relative to on-board applications, as well as its
relatedness to computer vision, one of the traditional
strong suits of ML.
But applications to other aspects of operations are
now starting to surface. The present review explores the
contexts for which these applications have been
proposed or in which they have been applied, exposes
the possibilities that they open up and risks that must be
avoided, and illustrates gaps in research that we believe
should be addressed by the Earth Observation
community.
A wide range of topics is discussed. We tackle
mission planning, a field in which the application of
sophisticated optimization algorithms has a long history
and that has seen some interesting progress recently;
Diagnosis, prognosis, and repair of faults, in which
either Bayesian Networks (BNs) or sequential neural
models seem poised to unlock great gains; enhanced
GNC, particularly through real-time guidance for
low-thrust trajectories and better processing of visual
data from star trackers and Sun/Earth sensors; and
on-board image processing, and how it enables
autonomous retasking and unlocks a tradeoff between
downlink capabilities and onboard compute power.
We also tackle some practical aspects of deploying
ML models in EO Operations. We survey how operators
can maximize the use of limited onboard capabilities by
appropriately optimizing the use of resources of ML
models, outlining a number of software and hardware
techniques devoted to that goal. We also review ongoing
efforts across the space industry to guide and
standardize the deployment of ML models, and the sort
of issues that a designer must keep in mind to avoid
common pitfalls of this technology.
Our goal is to provide EO Operators with a
comprehensive, if not exhaustive, an overview of the
current state of affairs of ML applications in their
domain of activity. The review aims to inspire
discussion and encourage further application proposals
and demonstrations, by connecting EO operators and
proponents of ML algorithms for EO Operations
problems.
The review tackles all on-board and off-board
applications of ML to EO with one notable exception: it
omits all post-processing of payload data on the ground,
a topic that has been studied extensively by past
authors.
One other topic that we do not dive into is the
optimization of tracking, telemetry, and command.
Although originally this was our intent, we found an
excellent and up-to-date review by Fourati and Alouini
[8]. Rather than needlessly repeat their work, we
encourage interested readers to check out the
outstanding article.
This research was produced by a team of 16
volunteers from the Small Satellite Project Group
(SSPG) of SGAC. SGAC is a non-profit
non-governmental organisation focused on the peaceful
uses of outer space with over 16,000 members
worldwide. It hosts 11 project groups, including the
SSPG, with more than 100 active volunteers. The SSPG
focuses on the uses of small satellites across the space
industry and how they can unlock the potential of space
for all of humankind.
2. Machine Learning in Earth Observation
Mission Planning
There are many constraints to mission planning.
Some relate to the target area: It needs to be under the
satellite and illuminated by the sun at capture time (orbit
and time-dependent); clouds are to be avoided (weather
dependent); the requester may set a deadline and/or a
priority. Others relate to the satellite such as limited
memory capacity; limited transmission capability;
reduced communication opportunities with the ground
antennas; multiple sensors to choose from; and limited
IAC-21-B.1.4.4 Page 3 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
manoeuvrability to skew the observation angle and
reach areas not directly flown over.
All of these parameters make optimal scheduling of
observations a highly combinational problem for a
mission that supports multiple independent requests and
it is even more complex when they are accomplished by
a constellation of satellites. Computational theory
literature does not provide deterministic algorithms that
can solve this problem in a timely and cost-effective
way. ML proposes a series of algorithms that may find
better solutions more efficiently.
There are many different formulations of the
observation scheduling problem, taking into account
different subsets of the constraints presented in the
previous section, adapted for different types of missions
and ground segments. The CCSDS’s mission planning
and scheduling [9] standardized definitions for the space
industry and provides an overview of the application of
the CCSDS’ Mission Operation service framework.
2.1. Current approaches
Non-ML algorithms to the satellite scheduling
problems can be classified into two categories: Exact
methods, and heuristic methods [10].
Exact methods typically consist of a combination of
branch and bound methods and mixed-integer linear
programming. These methods are computationally
costly and can become intractable for moderately sized
constellations.
Heuristic methods use an approximated rule to guide
the construction of a solution. Greedy algorithms
construct a solution by gradually choosing the best
action at every decision step according to some metric,
without regard as to how the overall sequence of
decisions plays out. Other heuristic methods include
backtracking through constraint programming and
search algorithms. Other forms of search include
hill-climbing or squeaky-wheel optimization, where the
geometry of the optimization functions is exploited to
accelerate the search process. Globus et al. [11]
compare multiple algorithms such as genetic, simulated
annealing, squeaky wheel and hill-climbing on a
problem with one or two satellites.
Evolutionary or genetic algorithms simulate
processes akin to biological evolution to optimize
candidate solutions according to a hand-crafted fitness
function. Mansour et al. [12] study the performances of
a genetic algorithm for a single satellite with limited
memory and multiple instruments and imaging modes.
Li et al. [13] explore genetic algorithms in order to
provide scheduling in real-time, optimizing the
transmission path towards the user.
Simulated annealing imitates the annealing
processes found in metals exposed to high temperatures,
and it forms the basis for another branch of heuristic
algorithms. The simulated annealing seems to provide
better results which are confirmed by Globus et al. [14]
in a very complete multi-satellite formulation of the
problem including satellite agility and priorities.
Lastly, multi-agent systems simulate interactions
between simple agents representing part of the systems
to determine an optimal policy. Bonnet et al. [15] use a
self-adaptive multi-agent system for real-time and
robust adaptation of a multi-satellite problem including
request priorities.
2.2. ML-based approaches
ML-based approaches can exploit the statistical
distribution of typical problem settings to accelerate the
finding of good solutions to the mission planning
problem.
Wang et al. [10] present a comprehensive review of
publications on the agile observation scheduling
problem, including ML and non ML approaches. The
authors classify approaches along multiple axes such as
time continuous and discrete-time model, type of
solving method and also other features such as
autonomy, and multi-objective profit function.
Neural networks (NNs) are explored by Wang et al.
[16] in order to provide immediate results for a
multi-satellite mission using deep reinforcement
learning (DRL).
Peng et al. [17] apply recursive NNs in a sequential
decision-making process in order to achieve low
scheduling computation time and low loss of profit
when compared to a deterministic resolution. Recursive
NNs allow the model to condition current decisions on
past inputs, instead of depending exclusively on the
present inputs to the system, providing the model with a
sort of memory.
We have not found any applications of Transformers
to this problem, a new sequence modelling technique
from the deep learning research field that has shown
excellent results in other sequential tasks like language
modelling and even in image processing tasks.
Neuroevolutionary techniques combine the
advantages of neural models and evolutionary
algorithms. Du et al. [18] leverage a prediction model
trained by a Cooperative Neuro-Evolution of
Augmenting Topologies algorithm in order to filter
tasks to be scheduled according to the probability to be
fulfilled before scheduling using genetic algorithms.
DRL uses NNs as function approximators to
approximate hard to determine functions in dynamic
programming. This has enabled groundbreaking
achievements in other control and scheduling problems
like playing Go or automated driving. Despite its
potential, it has not been extensively applied to this
problem set.
IAC-21-B.1.4.4 Page 4 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
Liu [19] applies Proximal Policy Optimization, a
method of the DRL literature, to mission planning for a
single satellite. Unfortunately, they do not compare
performance to other methods or extend it to a
multi-satellite setting.
Hadj-Salah et al. [20], [21] explore the application
of A2C, a DRL algorithm, to the mission planning
problem. They compare it to random planning and a
planning heuristic that compromises between greedy
and long-term planning. Their models are trained on a
simulated mission planning environment, then executed
on a real test scenario. Their long-term version of A2C
shows better performance than the heuristic algorithm,
albeit by a slim margin. In their later publication, they
augment the training process with techniques from the
domain randomization and transfer learning literature,
meant to increase robustness to the gap experienced
when passing from the simulated training scenario to
the real validation scenario. Their model beats the
heuristic approach by a slim margin.
2.3. Recommendations
Mission planning is a very rich problem that has
been explored for many years using machine learning
amongst other solutions.
Comparing the performances of algorithms
presented in different papers is not a suitable path
because each presents its own definition of the problem,
with a unique set of constraints, different mission
characteristics, variable satellite capabilities and
potentially incompatible metrics. For instance, a lot of
schedulers take into account satellite memory, limiting
the number of observations until a ground station is
visible, but few of them also make sure the ground
station is available for communication with the satellite
and not busy communicating with another one of the
constellations. Song et al. [22] introduce a framework in
order to facilitate future comparisons but additional
work on model standardization is needed before results
from different studies can be compared.
We observe a shifting trend in algorithms applied to
this problem over the years from genetic or annealing
to ML approaches such as NNs. Unfortunately, we
found no sources comparing the performances of
genetic and ML-based schedulers on a single problem
formulation.
Standardizing project formulations, constraints sets,
and optimization metrics seems to be a necessary step
for sustainable collaborative research in this field.
Relying on CCSDS’s published standards and models
could be a first step in the direction of a unified
approach.
3. Machine Learning in Earth Observation
Guidance, Navigation, and Control
Guidance, Navigation, and Control, often shortened
to GNC, describe the set of operations needed to move a
satellite platform or any other vehicle. Guidance relates
to the planning of paths from a current state to the
desired state. Navigation is the determination of the
present state. Control is the correct use of spacecraft
actuators, such as an engine, to execute the desired plan.
3.1. Current approaches
Two main tasks need to be achieved by a GNC
system: determination of the current state, which is an
estimation task, and use of the spacecraft’s actuators to
go from the current state to the desired state, which is a
control task.
Spacecraft control is typically subdivided into at
least two different granularity levels, where guidance is
the high-level control of the spacecraft from a current
dynamic state to a future one. A guidance module may
output the sequence of feasible dynamic states
necessary to achieve a new orbit from the required orbit.
EO Satellite maneuvers are often planned and optimized
on the ground, and the onboard guidance modules are
minimal. For the control task, a number of control
schemes are used, most notably controllers from the
robust control literature such as H∞ controllers.
As for navigation, spacecraft state is typically
determined via variants of the Kalman filter, such as the
Extended or the Unscented Kalman filters. These
methods are model-based, that is, they depend on an
explicit model of spacecraft dynamics for their
calculations.
Fuzzy controllers have been proposed as a possible
improvement to the classical approach. The literature
contains several works where GNC and AOCS
controllers based on fuzzy logic are compared to their
traditional counterparts. For instance, Wu et al. [23]
studied fitting the no longer in use X-38 re-entry vehicle
with fuzzy logic controllers. ESA also investigated the
usage of fuzzy logic controllers to carry out GEO-orbit
rendezvous autonomously [24] to aid in in-orbit
manufacturing. As another example, in [25], a
simulation of ROCSAT-1 / FORMOSAT-1’s attitude
controller is carried out, where the classical setup of a
PI pitch axis controller and PID roll/yaw axis controller
is replaced with two fuzzy controllers initially, and a
single consolidated fuzzy controller afterwards, yielding
considerable improvements against interference as well
as a lower steady-state error.
Nevertheless, despite the body of research backing
up their effectiveness, there is no widespread use of
fuzzy logic GNC controllers for space missions.
IAC-21-B.1.4.4 Page 5 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
3.2. ML-based approaches
Izzo et al. [26] present a survey of Artificial
Intelligence applied to GNC which, while not focused
on EO applications, can nonetheless be useful to
practitioners. The survey contains a section focusing on
ML approaches, on top of other AI approaches such as
evolutionary algorithms.
In another publication [27], Izzo and Öztürk
leveraged DRL to plan near-optimal real-time
computation of low-thrust transfers. They also suggest a
new method to generate training data for such problem
settings. Although originally designed for Earth-Venus
transfers, their solution is applicable to all low-thrust
transfers, but the data generation algorithm and
optimality comparisons are problem-dependent.
ML excels in problems where no structured
pre-existing model can be exploited. That is not the case
for the GNC problem, where the general form of the
dynamics governing spacecraft are well known and
solvable. It is however the case for visual-based GNC,
as no model exists for relating camera inputs to dynamic
state or control actions. For this reason, much research
on ML-powered GNC has focused on visual-based
GNC [28]–[30] for autonomous rendezvous. This,
however, is not directly relevant to the EO Operations
community, who are unlikely to engage in autonomous
docking as providers.
Hovell and Ulrich [31] propose a guidance module
for autonomous docking based on DRL that takes as
input the spacecraft’s state vector and distance to a
target spacecraft and outputs the desired velocity. A
classical controller takes over and transforms the
velocity output into control commands for the
spacecraft’s actuators. The algorithm, called Deep
Guidance, is trained for tracking and docking, and its
architecture is not easily convertible to other tasks,
which subtracts from its interest for the EO Operations
community. Nonetheless, it provides an example of
deep learning being used for spacecraft guidance.
An interesting streak of research looks into
applications of ML to processing visual navigation
sensors, particularly Earth and Sun sensors and star
trackers. Koizumi et al. [32] present a DL-powered
Earth sensor capable of determining the attitude of the
spacecraft by processing the images captured by a
COTS camera. It runs a real-time image processing
algorithm to extract features into the images separating
them into distinct feature sets using DL techniques. The
features sets are then compared to the preloaded data
sets to determine the position of the spacecraft relative
to Earth in the 3D plane. The primary advantage of the
system is the use of a COTS component and a single
board computer.
Another research thread explores the combination of
ML techniques and fuzzy controllers [33]. Classical
fuzzy controllers rely on manually set parameters that
define behaviour. This research thread attempts to
leverage ML techniques to learn the optimal value for
these parameters from a training dataset. These have the
advantage of interpretability - their reliance on
explicitly (if fuzzily enforced) rules means that they
remain grounded on human-interpretable system
models. Joghataie’ PhD. thesis [34] suggests the
development of a neuro-fuzzy controller, wherein the
tuning of the fuzzy logic is performed automatically by
using neural networks in a hybrid approach. Azarbad
[35] suggests a model applied to GPS systems that
outperforms the classical fuzzy controller. A simulation
study on MATLAB was done by Baranwal et al. in [36],
comparing the performance of a PID controller and a
fuzzy PID controller for a student satellite team. The
EKF based fuzzy controller outperformed the classical
controller. The study was done on a 3U CubeSat.
Further research can be done comparing these
controllers with ML-based approaches, paving the way
for future studies.
We have been unable to find a comparison between
the three types of controllers, i.e. neuro-based controller,
fuzzy controller and a hybrid model, as implementation
details in different studies differ, complicating their
comparison.
Wang et al. [37] have developed a DL framework
that stabilizes the spacecraft using a real-time torque
control. It is initially trained in a simulation
environment which enables it to learn the required
torque output and extrapolates it for unknown
disturbances. It performs better than a conventional PID
controller, as it can correct the attitude after unknown
disturbance rather than repeatable corrections.
A similar system is proposed by Yadava et al. [38].
They propose an AODS system that determines the
position of the spacecraft, taking inputs from the
magnetometers (magnetic vectors) and sun sensor (sun
vector) along with GPS data (position and velocity
vector), and determines the ideal attitude depending on
the position using a neural network. The required torque
calculations are made and sent to the RL based
controller to make the required adjustments. The system
performs better than classical PID controllers as it
consumes less computation power for subsequent cycles
as the algorithm learns.
3.3. Recommendations
Most ML for GNC applications in the space sector
seem to have been explored in the context of space
logistics and space exploration, rather than Earth
Observation. Although guidance and control for EO
platforms are simple compared to these applications, we
IAC-21-B.1.4.4 Page 6 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
believe there is a potential to adopt some of these
technologies.
Attitude determination is an area where EO
operations have high requirements. We believe that
vision processing applied to this area is just getting
started, and that use of more refined neural architectures
could enable improvements in performance or resource
consumption compared to current approaches.
4. Machine Learning in Earth Observation Fault
Detection, Isolation, and Recovery
Satellites performing EO tasks have stringent
requirements in terms of accuracy, continuity and
stability of payload operations. To this end, Fault
Detection, Isolation and Recovery (FDIR) is focused on
the development and improvement of tools to guarantee
and maintain reliable spacecraft operations. FDIR
describes a set of engineering disciplines focused on
safeguarding the spacecraft and maintaining it in
nominal operating conditions. The target of these
disciplines is represented by faults, irregular
occurrences and processes with the potential to disrupt
the mission, up to the point of failure.
ML can be an extremely powerful tool for FDIR.
Indeed, the core capability provided by ML is pattern
detection. Therefore, ML can be used both to detect
anomalies in the telemetry or outputs from any
subsystem (diagnosis), and identify signs indicating an
incipient fault (prognosis). This section presents
relevant ML literature for four significant sub-topics:
fault detection, fault diagnosis, recovery, and fault
avoidance
4.1. Fault Detection
Failure detection deals with identifying the presence
of faults and their rates of occurrence.
4.1.1. Current approaches
In classical approaches, the recognition of failures is
mainly based on constant thresholds and fixed logic
diagrams, defined during the design process. One of the
key issues with classical fault detection is model
brittleness. As fault detection schemes are based on
hardcoded thresholds, these models are easily disrupted
by noise and deviations from theoretical assumptions.
4.1.2. ML-based approaches
An example of an ML-based solution to the issue of
model brittleness can be found in a paper by Jaekel et al.
[39]. This work uses Self-Organizing Maps (SOM), an
unsupervised variant of Artificial Neural Networks
(ANNs), for the detection of failures in dexterous
manipulators for on-orbit servicing (OOS). SOMs
manage to adapt to the idiosyncrasies of incoming data
in a simulated environment and thus show increased
robustness to input variations with respect to traditional
methods. They can also deal with uncertainties and
noise in values. A client satellite is captured by a
servicing satellite with a dexterous manipulator having
7 degrees of freedom. They inject sensor failures,
including sensor outage and drift, during arm operations
and the results show that SOMs are a robust approach as
temporary fluctuations in the sensor, outliers and peaks
do not unnecessarily stop the current operation. But the
computational load is relatively high and needs to be
optimized to reduce system reaction time. The authors
suggest improving the precision and speed of the
method by adding more information from redundant
sensors.
Fuertes et al. [40] discuss ML-based fault detection
using NOSTRADAMUS, an algorithm developed by
the Centre National des Études Spatiales (CNES).
NOSTRADAMUS uses a One-Class - SVM
(OC-SVM), a common unsupervised learning algorithm
used to detect outliers, to detect the presence of an
anomaly in telemetry data. NOSTRADAMUS runs on
the ground segment, analyzing telemetry as it is
downlinked from the satellite. The performance of
NOSTRADAMUS is compared to algorithms inspired
by Novelty Detection (ESOC), Project Sybil [41], and
ATHMoS (DLR) [42]. If missing an anomaly is
regarded as unacceptable, NOSTRADAMUS is highly
preferred as it reaches a 100% detection rate with the
lowest false alarm rate (5%). For false alarm reduction,
the Novelty-inspired algorithm performs better with
85% of correct detections and less than 1% of false
alarms.
CNES is working on an on-board version of this
algorithm, as well as on extensions to the ground-based
variant for processing of multiple telemetry variables
based on dictionary learning approaches [43]. In
conversations during their collaboration with this
project, CNES teams signalled that explainability was a
crucial aspect of any technique. Being able to
understand the features of input data that signal a fault
lets the operational teams understand the context of
their satellite and know which actions must be taken to
remedy the situation - this is comparable in value to
being able to detect the anomaly in the first place.
Project Sybil is a joint initiative between the
Columbus Flight Control team by DLR and the
Advanced Mission Concept Section of ESA, and the
Ludwig Maximilians Universitat to apply an outlier
detection algorithm to the Columbus telemetry database.
It utilises the Density-Based Spatial Clustering of
Applications with Noise (DBSCAN) algorithm for
preprocessing the data after data segmentation and the
computation of its respective features. It is an
IAC-21-B.1.4.4 Page 7 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
unsupervised clustering algorithm that groups the data
according to their relative distances into a variable
number of clusters. Once this clustering is done, the
clusters containing less than 5 percent of the population
data are eliminated under the hypothesis that they may
represent non-nominal working modes in the learning
dataset. Project Sybill allows increasing mission success
rates by reducing the downtime due to failures of the
onboard systems.
4.2. Fault Diagnosis
While fault detection aims to identify the presence
of faults and performance degradations, fault diagnosis
aims to identify the root causes of these occurrences.
4.2.1. Current approaches
Though traditionally, fault diagnosis has been
achieved by human operators at the ground through
comparison with hardcoded hand-tuned thresholds. It is
difficult to deal with large amounts of data using this
approach. Iverson et al. [44] point out that for efficient
utilization of the data, there is a need for an autonomous
approach that eliminates the necessity of human experts
for diagnosis.
4.2.2. ML-based approaches
Ricks et al. [45] examine fault detection and
identification for a satellite Electrical Power System
(EPS) testbed using BNs compiled to arithmetic
circuits. BNs can be used to model partial knowledge
and uncertainty by identifying the system state based on
probabilistic relationships between a set of system
variables at a certain instance in time Meß et al. [46].
The proposed methods work for complex systems
exhibiting both continuous and discrete behaviour. The
discussed techniques can handle abrupt continuous
faults particularly well, which often pose problems. For
example, a nominal value region is not enough to detect
offset faults if they are small enough - the paper uses
cumulative sums to deal with these. Additionally,
“stuck” faults may be difficult to detect in low-noise
conditions since fluctuations might be infrequent. The
authors employ a tunable time interval which will mark
the sensor as working abnormally after it expires
without the readings having made any change.
Different types of nodes, modelling different
behaviours, are grouped to defined sensors and
components, which in turn are assembled to create the
entire EPS functional FDIR structure.
BNs have also been used by Schumann et al. [47] to
detect onboard failures and perform diagnoses. A
Software and Sensor Health Management (SSHM)
system is created for a simple GNC structure of a small
satellite using BNs that get information from hardware
sensors, software status signals, software quality
signals, and data from the Operating System to compute
if any failures are present, what are the most likely
causes behind them, and also give a statistically-sound
quality measure of the diagnoses. The developed SSHM
system requires no modification to the satellite
subsystems for which it performs FDIR - it just uses the
sensor data outputs. That way, model-level and
code-level Verification & Validation can be performed
independently on the SSHM system to certify that the
rate of false positives and false negatives is below a
selected threshold. This SSHM, applied to a simple
GNC system, was able to detect and diagnose both
hardware and software problems successfully.
Nevertheless, it remains a simplistic case and more
research into hierarchical SSHM systems is required in
order to apply them to large-scale BNs. The approach
can further be extended to failures that are not modelled
and unexpected and due to arising behaviour.
Although not specifically related to space systems,
Liu et al. [48] reviews the existing techniques for
ML-based fault diagnosis in rotating machinery. In
general, it presents useful research and conclusions
which we consider can be applied to reaction wheels in
the AOCS subsystem of spacecraft. K-Nearest
Neighbour (k-NN) is the simplest method reviewed,
which exhibits ease of implementation but necessitates
careful fine-tuning and large computation and storage
space. The authors cite BNs’ strong prior assumptions
as the biggest shortcoming of this family of algorithms
while mentioning as main advantages the fact that it
possesses a clear physical explanation of how it detects
faults as well as its reduced storage space requirement.
SVM is also reviewed and its high-dimension accuracy
is highlighted, even if the physical meaning is obscured,
unlike with the previous two techniques. Finally, DL
techniques have the potential to learn from data up to a
degree of complexity much higher than any of the other
techniques without the need for a manually crafted
feature extractor. However, the main drawback of this
approach is the need for large samples in order to train
the network, which is difficult to obtain unless the
spacecraft is a new iteration of previously flown models
for which data already exists. If the satellite is a one-off,
this can only be obtained in an approximate manner by
creating a simulation environment. The authors
underline that future ML-based fault diagnosis methods
should not be purely data-driven, but should take into
account possible failure mechanisms, system models
and prior knowledge in general, in order to increase
diagnostic performance.
Voss et al. [49] explore the use of DL for fault
detection and isolation in a simulation environment. A
NN is developed, trained offline and tested to detect and
IAC-21-B.1.4.4 Page 8 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
isolate single faults in the reaction wheels, GPS, star
tracker and magnetometer subsystems, as well as two
simultaneous faults. Additionally, a case study with
PROBA mission parameters is performed, for the
AOCS subsystem only. The implemented system
yielded mixed results: while there is a near-perfect
performance for some subsystems, the network fared
poorly regarding others, namely misalignment faults.
Also, fault isolation was much more reliable than fault
detection. On top of that, a large dataset is required for
this system to work, so creating a simulation
environment is mandatory, especially for one-off
spacecraft, in order to acquire enough data for adequate
training of the network. This study also assumes there is
enough electrical and computing power available on the
satellite to run this deep-learning-based solution. We
overview techniques for reducing resource consumption
of deep neural models and other techniques in section 6.
4.3. Recovery
Recovery in FDIR is a reconfiguration of the faulty
element and/or overall spacecraft in order to achieve
nominal system behaviour [39].
4.3.1. Current approaches
Traditional FDIR is able to respond to predefined
events by selecting a recovery path from the available
set of options. However, the status of the system and its
environment can exhibit various kinds of uncertain
behaviour due to their dependence on the internal
subsystem, component reliability factors, external
environment factors (e.g., illumination conditions,
thermal, radiation) and on system-environment
interactions (e.g., resource utilization profiles, stress
factors, degradation profiles) [46]. Due to these
uncertainties, the system and its environment cannot be
completely observed by traditional FDIR concepts that
pose limitations to autonomous isolation and recovery
[46]. One important example is Mars Express whose six
months of operation time was lost due to non-resolvable
memory failure that repeatedly transitioned it into safe
mode [39].
4.3.2. ML-based approaches
Raiteri et al. [50] discuss the use of Dynamic
Bayesian Networks (DBNs) to address issues like
partial observability, uncertain system evolution and
system-environment interaction, as well as the
prediction and mitigation of imminent failures. The BNs
do not model the relationship between variables at
previous points in time. DBNs are an extension to BNs
that refer to past values of certain variables to express
dynamic aspects of the system over discrete time [46].
The approach is applied by Raiteri et al. [50] onto the
power subsystem of a simulated ExoMars rover, by
simulating different failure scenarios. The DBNs can
infer whether the system is currently in a normal,
anomalous or failed state. On detection of a failure, a
suitable recovery plan is suggested. A preventive
recovery plan may be proposed in case an anomaly is
inferred. The FDIR presented in this paper also has the
capability of performing a prognostic state estimation
that can also be used for preventive recovery. The
proposed approach has been implemented in an
on-board software architecture called Anomaly
Resolution and Prognostic Health Management for
Autonomy (ARPHA). The results show that DBNs are
suitable for failure situations that require autonomous
(preventive as well as reactive) recovery.
AIKO Technologies [51] have developed a software
library, MiRAGE, that can enable the spacecraft to
make autonomous decisions for processing telemetry
and payload. The library is meant to be installed on the
satellites to enable functionalities such as event
detection, predictive maintenance and autonomous
re-planning.
4.4. Fault Avoidance
Fault avoidance methods are concerned with
preventing the occurrence of faults.
4.4.1. Current approaches
FDIR in past missions worked under the notion that
a fault is detected and then the algorithm will react,
according to predefined scenarios. In other words, the
classical FDIR approach responds only after failure
detection - failure avoidance cannot be achieved by such
a method. Regarding ML-based models, one of the
bottlenecks to having an on-board failure avoidance
system is that the models are trained on the ground with
limited data that does not represent actual behaviour in
space. This gives rise to the requirement of real-time
access to the data, which can be used to represent
multiple onboard scenarios, and closely represents
spacecraft behaviour during the mission.
4.4.2. ML-based approaches
Especially notable in the context of ML-enabled
fault avoidance is the work of Labrèche et al. [52]
discussing the OrbitAI experiment onboard the
OPS-SAT spacecraft. OPS-SAT is a special ESA
satellite deployed with the scope of being a testbench
for novel software technologies in orbit. OrbitAI uses
ML techniques to obtain intelligent FDIR algorithms
enabling the onboard camera to avoid direct exposure to
sunlight. Interestingly the ML model used is trained
IAC-21-B.1.4.4 Page 9 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
on-board, rather than offline. The model is trained with
five training algorithms tested of those natively
provided in the MochiMochi library [53] for online ML
training: Adam, RDA, AROW, SCW, and NHERD.
When using the figure of merit of balanced accuracy,
only one model appears to achieve values significantly
different from 0.5: the AROW algorithm in
three-dimensional input space.
4.5. Recommendations
FDIR innovation has been applied mainly to deep
space missions, which need a higher degree of
autonomy due to their long communication delays
inherent to the long distances travelled. However, the
analysis of the literature suggests that ML in EO FDIR
has promising prospects. The approach can be extended
to diagnose failures that are not modelled, unexpected
and due to arising behaviour, which offers a great
advantage in overcoming the model brittleness issues of
traditional FDIR. ML-based solutions for fault detection
and diagnosis can be integrated alongside the traditional
FDIR of the satellite. But ML-based recovery is
virtually unexplored and much research is needed in this
domain. The majority of the work in this field concerns
BNs, while other research avenues remain largely
unexplored, such as ANNs and DL. As it would be
shown and discussed in Section 6, power and
computational resources remain a big concern for
ML-based FDIR, especially for small satellites. The
benefits offered by ML-based FDIR can be further
researched to be implemented in future EO satellites to
perform FDIR on the AOCS subsystem, GNC,
On-Board Data Handling, Power subsystem as well as
detection of faulty sensors.
5. Machine Learning in Earth Observation
On-board Image Processing
5.1. Cloud and novelty detection
Clouds cover 66% of the Earth’s surface and are an
obstacle when observing the Earth’s surface in certain
wavelengths such as visible light. Removal of clouds
from satellite images is an important preprocessing
phase for most of the applications in remote sensing.
Researchers have explored various forms of Cloud
detection like “Cloud / No cloud”, “Snow / Cloud”, and
“Thin Cloud / Thick Cloud”, using various approaches
of ML and classical algorithms [54]. Cloud
detection/filtering can be used alongside novelty
detection. Novelty detection is to detect unexpected
features and it is especially important while looking into
new environments.
Good cloud detection algorithms are necessary to
optimize bandwidth and memory usage in EO missions
[55] and before the implementation of segmentation and
object detection methods. Convolutional neural
networks (CNN) have demonstrated excellent
performance in various visual recognition problems
such as image classification, and enable accurate
onboard cloud detection in small satellites.
With the increase in EO missions coupled with
high-resolution modern sensors, there is an increase in
bandwidth requirement that leads to the need to utilize
new techniques to efficiently manage the bandwidth
resources.
5.1.1. Current approaches
In the majority of missions, all images taken are
transmitted to the ground, which requires a significant
amount of bandwidth. Traditionally, data collection is
done by specifying in advance where and when to take
the measurements. Based on the content of the data,
there is no mechanism to tailor what is downlinked.
Other common approaches include novelty detection
based on spectral contrast, radiance spatial or temporal
contrast. But these methods are better used for dark
grounds like vegetation or deserts as clouds contrast in
colour compared to them. Furthermore, these methods
rely on manually chosen thresholds, which are
time-consuming to find and sometimes brittle.
Whereas spatial coherence is a better method of cloud
detection in areas with little contrast with the clouds (ice
sheets). NNs have also been shown to have greater
flexibility with classifying indistinct classes like clouds
on snow.
5.1.2. ML-based approaches to Cloud
Detection
For cloud detection, Zhang et al. [55] propose a
lightweight DNN based on U-Net. For performance
estimation of the proposed method, training and testing
of the red, green, blue and infrared waveband images
from Landsat-8 were used. The lightweight DNN is
based on U-Net and obtained better overall accuracy
while reaching the state-of-art inference speed by
applying the LeGall-5/3 wavelet transform on the
dataset which compresses the dataset and accelerates the
network for on-board use. Zhang et al. experimental
results illustrate that the proposed model maintains high
accuracy after four-level compression [55]. They reduce
processing time from 5.408s per million pixels to 0.12s
per million pixels, and average memory cost by around
30%. They observed that increasing batch size reduces
the time cost and increases peak memory. The largest
batch size experimented in the work, 1000, led to a time
cost reduced to 0.01s against around 20MB on peak
IAC-21-B.1.4.4 Page 10 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
memory cost. On the contrary, a smaller batch size, 100,
shows a time cost of 0.05s and a memory peak cost of
around 17MB. By taking advantage of mature image
compression systems in small satellites, the proposed
method provides a good possibility of onboard cloud
detection based on DL and hence helps in improving the
downlink data transmission efficiency and reducing the
memory cost. U-Net provides better accuracy on
compressed datasets. Also, the U-Net framework
showed great potential for pixel-wise classification of
remote sensing datasets [55].
Hinz et al. [56] also work on the detection of clouds,
in the framework of the H2020 EO-Alert project.
However, the EO-Alert project aims at keeping images
of clouds and enriching them with alert profiles in case
of severe storms for weather broadcasting. The
algorithm used is ML-based Gradient Boosted Decision
Trees and is embedded in a modular image processing
pipeline. Currently, tests of the pipeline are performed
in Matlab and are ported on hardware to be flown to
space.
Srivastava et al. [57] suggested using Kernel
methods for better onboard discovery computation of
cloud detection over snow and ice. This paper proposes
a Kernel method that can be used for both clustering
and classification of images on board any satellite. The
paper discusses a novel variant of the Probabilistic
Kernel (P-Kernels) with a mixture of Gaussian and
spherical covariance structures. It is very sensitive to
even the smallest of changes as it assumes all
observations to be independent. The results showed
great promise with clouds being differentiated much
better from Greenland ice sheets when compared to the
Gaussian and a Gaussian mixture model.
Giuffrida et al. [58] discuss a CNN deployed on the
PhiSat-1 reconfigurable nanosatellite to analyze
imagery from its Hyperscout-2 payload and select
images eligible for transmission to the ground. It is
implemented onboard the ESA Phisat-I mission to
classify cloud-covered images and clear ones. Only
images with less than 70% cloudiness are transmitted to
the ground. The network is trained and tested against an
extracted dataset from the Sentinel-2 mission, which
was appropriately pre-processed to emulate the
Hyperscout-2 hyperspectral sensor. On the test set, 92%
of accuracy is achieved with 1% of False Positives (FP).
The results showed a power consumption of 1.8 W,
requiring memory of 2.1 MB, keeping within the power
and the memory constraints.
Other solutions that have not flown yet and are in
the concept phase have been developed. Maskey et al.
[59] proposed an ultralight CNN algorithm, called
CubeSatNet, that prioritizes quality data over quantity
without changing the constraints of size, power, volume,
downlink and pointing requirements imposed by a 1U
CubeSat. The algorithm is trained over 48000
augmented images from CubeSats and validated against
12000 augmented images from CubeSats to classify
images as “bad” when cloudy, sunburnt, facing space or
saturated. Images are classified as “good” in all other
cases. If in orbit, the algorithm would select only
“good” images to be downlinked and discard images
that are covered in clouds or too bright or dark. Trained
on BIRDS3 satellite images, the algorithm reportedly
has an accuracy of 90% and can cut operation time by
about 2/3 while significantly improving the quality of
images received.
Murray [60] proposed a concept of on-board
processing with two cameras: the nadir-looking camera
performs the standard observation, whereas a
forward-looking camera observes if clouds are coming
in the trajectory of the satellite. A neural net
classification grid is used to identify clouds and an
algorithm then decides when to capture images with the
nadir-looking. This approach would be oriented towards
CubeSats.
Castaño et al. [61] trained an SVM for estimating
the opacity of atmospheric dust and water ice on Mars
on data from the THEMIS camera mounted on board
the Odyssey mission. The authors use both a regular
SVM and a reduced-set SVM. The reduced-set SVM is
trained on a reduced synthetic dataset maximizing the
similarity of the reduced-set SVM to the regular SVM.
The reduced amount of support vectors decreases
compute requirements. They then test both the full-size
SVM and reduced-set SVM on flight software, showing
the capability of such software to run the proposed
algorithms.
The authors mention two challenges related to the
analysis accuracy of onboard THEMIS data. Firstly, as
the onboard data is not calibrated, the deployed models
must be robust to significant noise. Secondly, due to
temperature fluctuations, the response function of the
camera can gradually increase or decrease its values,
even when there is no change in actual value. The
authors suggest characterizing the operation of the
algorithms in an environment as close as possible to that
of the spacecraft.
Lastly, the CIAR project from IRT Saint Exupéry
demonstrated cloud segmentation on board the
operational test-bed satellite OPS-SAT in 2021 [62].
Figure 1) showcases a visualization of their results.
IAC-21-B.1.4.4 Page 11 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
Figure 1) On-board cloud segmentation from the
CIAR project
5.1.3. ML-based approaches to Novelty
Detection
Wagstaff et al. [63] show the benefits of reduced
downlink data when performing cloud detection and
filtering for EO missions. Cloud detection is
demonstrated using Random Decision Forests (RDFs)
and Bayesian Thresholding (BT), while a third
saliency-based algorithm is used for novelty detection
onboard EO-1. For classifying the pixels, the RDF
method analyzes a window of values around the pixel.
In contrast, the BT independently performs the
classification of each pixel. BT uses the difference in
particular wavelengths between dark surface materials
and bright cloudy regions. The novelty detection
algorithm identifies such regions within an image that
may contain new features. EO-1’s primary science
instrument is Hyperion. It’s an Imaging Spectrometer
capable of data collection with high Spatial and Spectral
resolution. Using data from previous mission phases,
both cloud detection algorithms were trained to drop
useless images from the telemetry downstream. The
performance of the algorithms has been evaluated
onboard over a five-month period from November 2016
through March 2017. In comparison to ground testing,
the on-board performance showed similar or better
results on a diverse collection of targets. Both RDFs and
BT reached an accuracy of more than 90%. However,
in real-time, the RDFs were faster. The novelty
detection was able to detect new features in remote
locations such as small lakes and buildings and hence
such images could then be given priority for the
downlink. Such methods must be able to successfully
operate on board with limited resources while posing a
minimum risk to the overall spacecraft. With the
advancement in computing capabilities, more complex
models offering better accuracy can be used onboard
future EO missions.
Chien et al. [64] provide results of the IPEX, which
originated from a CubeSat that flew from December
2013 through January 2015 and validated autonomous
operations for the processing and the generation of
product onboard the platform hosting the Intelligent
Payload Module of the Hyperspectral Infrared Imager
(HyspIRI) mission concept. IPEX worked as a testbench
for on-board image classification, achieved through the
use of ML-based random decision forest algorithms.
Compared to other missions, the solution was enhanced
with the use of an ensemble of multiple trees to increase
the classifier’s reliability via statistical regularization
without the need for explicit pruning of the trees.
Additionally, the algorithm analyzes spatial
neighbourhoods within each image to incorporate local
morphology and texture, rather than single pixels.
Attention was also paid to reduce runtime by classifying
every 10th pixel along with the vertical and horizontal
directions and filling in the remainder with
nearest-neighbour interpolation. An extremely
interesting point is that the IPEX classifier was trained
before launch using just four hand-labelled images from
a high-altitude balloon flight that employed the same
type of camera used by IPEX. According to the authors
of that research, it was the first time an ML system has
been trained on a suborbital flight and then successfully
used in orbit.
IPEX also tested an unsupervised method to identify
images with potentially interesting content, to be paired
with supervised learning. In particular, a computer
vision visual salience software was employed to extract
interesting regions for downlink in acquired imagery.
The algorithm used a simple pixel-based measure of
visual salience for grayscale images that incorporate
local context to be compliant with limited resources
onboard the CubeSat. The algorithm is applied to a
downsampled version of the image using a 32 × 32
pixel window to identify the five most salient regions
within the image. The pipeline is completed with
thumbnails of salient regions and their salience scores,
which are being saved and made available for downlink
and examination on the ground. Images at full
resolution can be furthermore downlinked to ground
stations, in case needed.
5.1.4. Recommendations
With the strict limitation on bandwidth, onboard
filtering of useless data enables sending data to the
ground with minimum compromise on image quality
and the need for human intervention for decision
making. The results of ML algorithms can be improved
in terms of accuracy and precision, with the availability
of newly generated data.
5.2. Object/Image Classification
Image classification is a task of extracting
information on the basis of objects in the images instead
of individual pixels, where “objects” are referred to as
IAC-21-B.1.4.4 Page 12 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
meaningful scene components that distinguish an image
[65].
5.2.1. Current approaches
While current methods do extensively apply ML
algorithms to great success, image classification is done
on the ground instead of onboard a satellite.
5.2.2. ML-Based approaches
Arechiga et al. [66] give an example of an on-board
processing application where a CNN architecture is
used for object classification and trained using satellite
imagery of Planet’s Open California dataset. Nvidia
Jetson TX2 is used for implementing this application.
The authors suggest that more research can be done so
that the application can be enhanced to classify more
objects.
Machine intelligence is used to perform onboard
analysis of EO tasks such as hazard analysis (e.g. fire
and flood detection), target detection, area monitoring,
and weather forecasting [67]. Researchers at NASA
Goddard used ML to detect wildfires on MODIS
(Moderate-resolution imaging spectroradiometer) data.
In practice, the use of CNN’s is composed of two
main tasks: training and inference. Training is the
process of “learning” the optimal set of weights that
maximize the accuracy of the desired task (e.g. image
classification, object detection, semantic segmentation).
It is highly compute-intensive and often is accelerated
by GPUs. The inference is the process of using a trained
model (where parameters are no longer modified) to
make decisions on novel data. The inference is a less
computer-intensive process than training and has been
performed on CPUs, GPUs, and FPGAs.
5.2.3. Recommendations
Similarly to the case of onboard cloud detection,
moving object classification and detection onboard
satellite platforms allows operators to reduce the load of
ground-satellite communications links. EO Operators
can leverage the huge and quickly expanding research
field of computer vision.
The high-level information gained by using object
classification can then be used for other tasks, like
dynamic mission replanning.
5.3. On-board image compression
New, complex onboard sensors can rapidly saturate
both the downlink bandwidth of communication
transceivers and onboard data storage capability. More
efficient image-compression codecs are an emerging
requirement for spacecraft and can significantly reduce
transmitted or stored data volume. However, in the
tradeoff mission design, it is also necessary to consider
whether these are computationally intensive and require
rapid processing to sustain sensor data rates.
5.3.1. Current approaches
Systems to compress data in spaceborne operations
employed a wide variety of both lossless and lossy
compression schemes. Lossy compression is usually
used when the system bandwidth is too low and cannot
support lossless compression, when the science value is
not compromised by the distortion that would be
introduced by lossy compression, or when other sensors
that do not play a role in primary data products are
included. An example of this last case can be
scene-context cameras.
5.3.2. ML-Based approaches
Goodwill et al. [68] proposed an ML-based solution
to achieve good reconstruction fidelity after lossy
compression. The algorithm, CNN-JPEG, makes use of
a hybrid approach combining CNNs and JPEG
Compression. In the encoder, the image is fed to a
3-layer CNN to obtain a compact image representation,
which is then encoded with JPEG. The encoder, based
on previous work, is denoted by ComCNN and learns a
compact image representation that is half the size of the
original image. In the decoder, the resulting image is
upsampled to the original size and decoded with a
deeper 20-layer CNN, which reconstructs the original
image by learning a residual image and adding it to the
upsampled image.
Experimental results for CNN-JPEG show a 23.5%
and 33.5% increase in PSNR and SSIM over standard
JPEG, respectively, on an image dataset collected from
STP-H5-CSP compressed to the same file size. On the
same dataset, CNN-JPEG performed a 1.74 times
increase in average compression ratio at fixed PSNR. It
is interesting to also note that, according to the research,
the encoding segment of CNN-JPEG in TF Lite,
running on the Cortex-A9 cores of the Zynq-7020,
yielded an average execution time of 16.75s using a
single thread. Using the TF Lite interpreter to parallelize
operations was reportedly far from ideal linear speedup.
Authors also showed that leveraging the Zynq-7020
FPGA resources through SDSoC for hardware
acceleration helped in decreasing the average execution
time of the CNN-JPEG encoder to 2.293 s, with a 7.30
times speedup over the single-threaded TF Lite solution
and 6.87 times speedup over the single-threaded TF Lite
solution.
Vladimirova et al. [69] discuss the development of a
lossless compression method without the drawbacks of
low compression ratios using predictive NNs, coupled
IAC-21-B.1.4.4 Page 13 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
with integral wavelet transforms and the Peano-Hilbert
(PH) Scan algorithm. This is then benchmarked against
the CALIC Method using various image datasets. The
image is first sent through the IWT (Integral Wavelet
Transform) to produce a de-correlated image, which is
mapped and a PH scan is performed after which the NN
(a two-layer, 4x106x 1) scans and allocates a probability
distribution for the next incoming value. On the tested
data sets, using only the NN method achieved an
average compression ratio of 2.530, compared to the
CALIC method which achieved a ratio of 1.806.
Introducing the PH scan brought an 8.5% improvement
compared to the CALIC method at 2.747. The
IWT+PH+NN method overall achieved an improvement
of 13.1% compression ratio over the CALIC method.
The paper proposes potential applications of the
algorithm in previewing a satellite image before a full
image is transferred to assess features of the image and
hence would prevent bad images from being sent, such
as those affected by clouds or images suffering from
other distortions.
Cai et al. [70] proposed a novel method for LIDAR
image data compression. The method is called feature
indexing where specific features are assigned to a data
index system generated by DNNs. The whole program
is then uploaded to onboard hardware and it stores it as
a dictionary for reference. The OBC runs a feature
isolation program and identifies features and creates a
resultant dataset of pure indices based on the directory.
This data set is then transmitted with the location data
and then is decoded on the ground. Achieves a
compression level of 99.17% and works far better than
standard wavelet compression methods. The method
was tested against the LIDAR data of the Space Shuttle
program and achieved the above-mentioned results.
5.3.3. Recommendations
The exploitation of lossy compression to ease
downlink clearly represents a path to be explored. The
work by Goodwill et al. [68] also emphasizes the
importance of advancement in the field of hardware
acceleration and SoC FPGAs. Indeed, on-board
inference of CNNs is computationally expensive for
space platforms. Further advancements can possibly
support the application of more complex algorithms
even in constrained environments.
6. Machine Learning in resource-constrained
Earth Observation platforms
This section addresses the topic of ML in
resource-constrained spacecraft performing EO tasks.
These methods represent a powerful set of enabling
technologies, relevant both for the emerging interest in
small satellites and to preserve the operativity of large
platforms experiencing failures or operating with shared
resources. Moreover, the consistent technological lag of
space hardware makes considerations about reduced
available SWaP almost always necessary when
redeploying architectures developed for Earth-based
applications into orbit.
Within the scope of this work, the constraint on
resource availability will be limited to on-the-edge
computational and sensing capabilities, and not
extended to the data. It is also out of the scope of the
section to address scheduling approaches, which
optimize the availability of resources to multiple
subsystems or users. This variability, however, can be
also seen as a source of constraint over the available
budgets.
We investigate two ways in which this adaptation to
technological limitations can be implemented:
optimization of the AI architecture itself, and
optimization of the interplay between the model and the
hardware this operates on. In general,
resource-constrained platforms it is necessary to
maintain a holistic view of the architecture of the
software, the hardware, and the data at play.
It is worth noting that another emerging
technological field presenting similar constraints to the
space sector is represented by Internet-of-Things [71],
where the target platforms for AI are small, low-power
devices.
6.1. AI Architecture Optimization
6.1.1. Pruning
Pruning is the operation of removing or zeroing
parameters of a NN model thus reducing the network’s
size [72]. This process is generally performed by
associating scores with the network’s elements during
training in order to select the ones to prune. The lighter
model is then further trained and can be iteratively
re-pruned several times. Multiple pruning strategies
exist such as varying the number and nature of items
pruned, the number of iterations performed or changing
the scoring criteria [73]. There are also other emerging
pruning paradigms that do not rely on an iterative
process [74] [75].
Pruning’s main trade-off is to increase
computational efficiency at the cost of quality/accuracy
and increased training complexity. The objective is to
leverage compression rates of 4, 8 or even 32 while
costing at worst only a few percent of accuracy [73].
Performing pruning along this objective remains a
delicate task as literature demonstrates that keeping
good performances is dependent on the pruning method.
The main challenge of implementing pruning is thus to
determine and test which pruning methods to use in
IAC-21-B.1.4.4 Page 14 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
order to achieve the required compression while
keeping acceptable performances for a representative
type of datasets.
Although the lack of standards in evaluation
impedes the comparison of the multiple existing studies,
they all advertise significant compressing at low
accuracy cost, including several algorithms confirmed
by multiple papers [73]. Pruning has been successfully
applied in many image processing use cases but has also
been proven on voice processing [76], credit
classification [77], and multiple other types of datasets
[78].
Additional engineering and more complex training
on the ground in order to significantly reduce the
onboard execution constraints make pruning an
attractive trade-off and a strong technological enabler of
NN implementation in space.
Pruning is now developed enough to have
documented implementation and examples in ML
frameworks such as TF [79].
So far, pruning has been used as part of complex NN
applications for space but only on the ground with
applications such as image classification [80]–[83].
There are some applications aiming towards on-board
implementations like remote sensing image
classification [84], [85], vehicle detection in satellite
image [86] and image anomaly detection [87].
Unfortunately, the authors were unable to find
documented evidence of a pruned NN that flew on a
space mission.
6.1.2. Filter compression and matrix
factorization
In its section concerning “convolutional filter
compression and matrix factorization,” the paper by
Goel et al. [88] presents methods to adapt neural
networks to low-power platforms by operating at a
layer’s level. The distinction operated between the two
distinguishes between the types of network elements
that are being optimized.
Neural Networks can be algebraically represented as
n-dimensional matrices known as tensors. Matrix
factorization approaches reduce the complexity of these
underlying tensorial structures, to obtain compressed
networks without significant loss of accuracy. Filter
compression methods, on the other hand, reduce the
number of parameters in the network architecture by
acting on the structure of filters in the so-called
convolutional layers.
In particular, Goel et al. observe that filter
compression methods are capable of achieving a state of
the art accuracy in computer vision, albeit at times at a
high computational cost. As computer vision tasks are
essential in EO operations, this class of methods appears
to be the most significant within the scope of this paper.
Two architectures emerging as relevant for filter
compression are SqueezeNet [89] and MobileNets [90].
Both these architectures have found applications in the
EO community. For example, modified SqueezeNets
have been used by Alhichri [91], Alswayed et al. [92]
and Alhichri et al. [93] for the classification of remote
sensing images (both in drone and satellite images). In
particular, Alswayed et al. report results comparable to
or outperforming the state of the art at the time of
publication.
Poortinga et al. have used a MobileNet-based
architecture to map sugarcanes in satellite data of
Thailand [94], obtaining significant accuracy for the
task. Zhang et al. [95] also have used an architecture
capitalising on MobileNet, reporting results
outperforming the state of the art at the time. Similarly,
Yu et al. [96] present a MobileNet-based method to
classify remote sensing imagery and report
outperforming many state-of-the-art models while
requiring a smaller amount of training data. In their
report paper, Hoeser et al. [97] note that: “It is
important to note the small group of six items which use
MobileNets(...), of which five were published in 2019.
They describe an onset of interest in parameter efficient
models with high accuracy and they prove that such
models can compete in Earth observation studies.“
6.1.3. Architecture search
Neural Architecture Search (NAS) refers to a set of
tools and processes for the automatic generation of
optimal architectures for an ANN. NAS is a specific
instance of automated machine learning (AutoML), the
process of automating the overall ML construction
process [98]. As shown by Chan et. al [99], this process
can be specialized to address a constraint on available
resources.
Seminal developments in NAS emerged in late
2016, from the work of Zoph and Le [100] and Baker et
al. [101]. In a survey on the subject, Elsken et al. [102]
report three key parameters to operate a classification of
NAS processes. These are:
●Search space;
●Search strategy;
●Performance estimation strategy.
Being an approach to adapt the heavy computational
cost of NN to resource-constrained platforms, NAS has
naturally found application in many space-related use
cases. EO, there has been quite a research on
hyperspectral images classification using NAS, with
development performed by Liang et al. [103] have
employed NAS (and pruning) to detect aircraft in
remote sensing images. NAS Mobile neural architecture
IAC-21-B.1.4.4 Page 15 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
search (MNAS) [104] is a probable candidate in
implementing NAS to EO satellite inference on the edge
application.
6.1.4. Knowledge Transfer and distillation
In knowledge transfer (KT) and knowledge
distillation (KD) a small, lightweight network is trained
to reproduce the behaviour of a large, computationally
intensive network without having to fully duplicate the
architecture of the latter. This leads to small networks
both providing results comparable to those of large
networks and deployable on resource-constrained
platforms.
According to the paper of Goel et al. [88], in KT the
smaller network is trained using data labelled by the
larger network (defined as “synthetically labelled data”
by Ba and Caruana [105]), while in KD a small network
(student) is trained by a large network (teacher) to
replicate the latter’s output. Within the scope of this
section, it appears also relevant to discuss transfer
learning, which has attracted considerable interest from
the space community.
De Vieilleville et al. [106] proposed a distillation
method to perform DNN-mediated segmentation of EO
images on board of cubesats. In this work, they show
that a 10 to 30 fold reduction of the free parameters of
the network mediated through distillation leads to
weakly worse performance (+5/-10% accuracy).
Similarly, [107] provides a detailed distillation
implementation and results showing a strong reduction
of the NN execution load while keeping a steady
accuracy in remote sensing scene classification. [108]
applied distillation for mapping irrigated areas using
remote sensing data.
Since 2019, self-distillating networks are emerging
[109] with one successful implementation for cloud
detection in remote sensing by [110] achieving 200-fold
compression. Industrialization is not as developed as
pruning as there are only a few open access examples of
implementations but no widely developed library.
Unfortunately, the authors were unable to find
documented evidence of a distilled NN ever flown and
used on a space mission.
6.2. Hardware Acceleration
Computing limitations are demanding to ML-based
applications because of the significant amount of data to
be processed for DL. Many NN models require
high-end GPU devices to run in inference, and even
more so during training. In deploying ML to an EO
satellite it is widely acceptable to consider the
inferencing step of the process due to the
resource-constrained nature of a satellite system in orbit
in terms of volume, power, and mass especially under
the CubeSat standards. Progress in commercially
available off-the-shelf hardware in mobile edge
computing has a progressive effect in finding their way
to CubeSats in implementing DL algorithms for space
applications [111].
With CPUs considered to be general-purpose
computers, AI-specific hardware such as GPU’s, FPGA,
and ASIC takes the center stage which is designed to
accelerate the computation of linear algebra and
specializes in performing fast and matrix multiplications
with higher performance-per-watt ratios. Furthermore,
advanced next-generation architecture for onboard
computing which heavily depends on artificial
intelligence is developed like AI-OBC (Artificial
Intelligence-OnBoard Computing) [112] based on
distributed on-board architecture consisting of CPU,
Visual Processing Unit (VPU) emerging AI accelerator
class of microprocessor for running machine-learning
applications to train DNN and FPGA connected through
Cubesat Service Protocol (CSP) through which ML and
training are carried out in real-time with commercial
off-the-shelf COTS components to reduce cost and
development time.
Table 1. Hardware Accelerators
Name
Company
Description
Intel Movidius
Myriad 2
Vision
Processing
Unit (VPU)
Intel
Implemented with
DNN in Phisat-1 [113]
Myriad X
(VPU)
Intel
Active testing
[114]
Jetson Nano
(GPU)
Nvidia
Space Edge Zero
(2021) by Spiral blue
[115]
Tegra TX1
and TX2
(SoC)
Nvidia
Demonstrated AI
Image processing
capability [116], [117]
Coral TPU
Google
Used with SC-LEARN
Architecture for
Hyperspectral models
[118]
Apache 5
Almotive
In development
Neuromorphic
chip
Innatera
In development
IAC-21-B.1.4.4 Page 16 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
Spaceborne
Computer-2
(SBC-2)
Based on
Intel Xeon
Onboard ISS
Ultrascale
Radiation
Tolerant (RT)
Kintex FPGA
Xilinx
Prototype available
Xilinx
Zynq-7020
(ARM
Cortex-A9 +
FPGA)
Xilinx
Space Test Program
Houston 5/ CSP
(2017)
6.3. Quantization / BNNs
In quantized networks, the number of bits used to
represent numbers defining a model is reduced. This
provides a significant decrease in compute. memory,
and power requirements, for a comparatively low
decrease in performance. Quantization may be applied
to weights, activation functions or gradients of a
network, either during or after training. [119]–[121].
Quantization has been explored in research for remote
sensing image segmentation and processing but appears
to never have been flown in space.
Perhaps the most common established quantization
technique is reducing the bit-width of weights after
training. However, very low bit widths, typically of four
or less, usually incur heavy losses. This can be
mitigated by performing model training under the
reduced bit-width quantization. Good results have been
achieved with quantization, even going all the way to a
single bit.
Accuracy on par with full-precision NNs was
achieved for standard datasets in publications such as
Binary-Connect, XNOR-Net, and TTQ [122]–[124].
Quantization of already existing NNs such as AlexNet
[5] and VGGNet [125] applied to the ImageNet dataset
has been carried out without any accuracy loss while
reducing their sizes up to 50 times [126]. Quantization
both before and after model training is provided today
either as part of mainstream DL libraries [127], [128] or
third-party libraries such as Larq [129] respectively.
Although there exists no consensus on why
quantization works, a candidate explanation argues that
large amounts of pathway redundancy in NNs make the
expressivity loss a minor concern. Theoretical analysis
in that regard is still limited. Anderson and Berg [130]
found that statistical properties of the computation are
kept even when a network is binarized. Molchanov et al.
[131] indicate that nearly 99% of weights can be pruned
in certain NNs and achieved a 68-times sized reduction
on VGG-like networks without loss of accuracy.
Broadly speaking, quantization techniques can be
divided into two main categories: Deterministic and
Stochastic. Guo classifies deterministic quantization
methods [119] into:
●Rounding: Floating-point values are assigned
their nearest fixed-point representation.
●Vector Quantization: Weights are clustered
into groups, with the centroid of each group
replacing the real weights.
●Quantization as an optimization: Here, the
quantization is treated as an optimization
problem, which involves minimizing an error
function taking into account real and quantized
weight values.
Regarding stochastic quantization techniques, they
separate them into:
●Random Rounding: The quantized value is
obtained by sampling a discrete distribution
parameterized by the real values themselves.
●Probabilistic Quantization: Weights are
assumed to be discretely distributed, with the
methods trying to estimate which distribution
function it is.
Deterministic quantization has seen extensive
success, with rounding being the most commonly
successfully employed type of quantization, such as
Rastegari et al. [123] and [132], where a general
rounding function was introduced. In particular,
Binary-Connect Courbariaux et al. [122] used binary
rounding, achieving 98.8% accuracy on the MNIST
dataset. Also noteworthy is the use of vector
quantization in Gong et al. [133], where a network
compression ratio of 24 was obtained, losing only 1% of
accuracy on the ImageNet dataset. However, Stochastic
quantization has not experienced such a resounding
success, perhaps due to an over-reliance on statistical
assumptions which are not guaranteed to hold.
Quantization approaches may quantify several or all
of the following:
●Weights: The action of quantizing weights
yields a smaller network size and can
accelerate both the training as well as the
inference process. This comes at a price,
however: NNs will have a harder time
converging when training with quantized
weights, and a smaller learning rate is required.
Additionally, the gradient cannot
back-propagate through discrete neurons,
leading to the use of straight-through
estimators in order to estimate them, usually
with a high variance.
●Activations: The goal of quantized activations
is replacing inner products with binary
operations, reducing memory constraints since
the operation precision is reduced, all while
accelerating network training. In fact,
IAC-21-B.1.4.4 Page 17 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
activations may fill more memory than weights
[134]. Note that quantized activation will cause
what is called a “gradient mismatch”, where
the gradient of the activation function is
different from the one obtained from the
straight-through estimator used.
●Gradients: Quantizing the gradients is still a
relatively new avenue of research in NN
quantization. The main objective here is not
reducing the model size, but aiding in
distributed network training, where several
computing nodes need to share information of
the gradient values between them. The smaller
the size of the data the nodes need to share, the
faster parallel training can be performed.
Quantized gradients need to be carried out with
care since unsuitable implementations run the
risk of causing the gradient descent algorithm
not to converge.
7. Machine Learning standardization and issues in
Earth Observation Operations
Interest in AI and ML has increased in the past
years. Many organisations across different industries
and geographical locations are currently working on
creating guidelines, best practices, and standards, to
guide the correct implementation of these systems.
However, the process is far from being complete, and
we have found only one such example from the space
industry.
Standards, guidelines, and other documents
discussed in this section blur the line between
definitions of AI and ML. While we find this fact
misleading, we have kept the original usage from the
sources in order not to alter their message.
These bodies of work aim at aiding ML system
developers to avoid common pitfalls and problems
associated with these systems. We provide in this
section a cursory overview of what these problems are
in order to raise awareness amongst EO platform
operators.
7.1. Guidelines and Roadmaps
International Standards by ISO/IEC JTC 1/SC 42
committee [135] are currently available or under
development. These standards and projects represent the
united efforts of experts and entities in providing
guidance and focus on the standardization of Artificial
Intelligence, with currently more than 20 under
development and 6 already published. We found
ISO/IEC TR 24030:2021 to be particularly interesting
as it covers 132 use cases, as well as the projects under
development concerning Functional Safety and AI, data
quality and AI explainability.
The International Standardization Organization is
not alone in working on AI standardization, though.
The Chinese Big Data Security Standards Special
Working Group of the National Information Security
Standardization Technical Committee (NISSTC) wrote
the Artificial Intelligence Security Standardization
White Paper [136]. The focus of this White Paper
ranges from the security of AI to main security threats,
risks, and challenges. Seven recommendations have
been made on the importance of improving a system of
AI security standards, the need to speed up the
development of standards in key areas, promoting the
application of AI security standards, strengthening the
training of AI security standardization talent,
participating in international AI security
standardization, establishing an AI high-security risk
early warning mechanism, and improving AI security
supervision support capabilities.
Germany developed an Artificial Intelligence
Standardization Roadmap [137], continuously updated,
as a joint effort between DIN and DKE. The roadmap
strongly supports the idea that standardization would
help in getting the explainability and reliability of AI,
thus favouring its application. In the roadmap, five
recommendations are drafted. More specifically, they
deal with data reference models for the interoperability
of AI systems, development of an AI basic security
standard, practice-oriented initial criticality checking of
AI systems, national implementation program “Trusted
AI'' to strengthen the European quality infrastructure,
and analysis and evaluation of use cases for
standardization needs.
In addition, the work provides extensive analysis on
the definition of AI as well as classification schemes to
evaluate AI-based systems.
The work is particularly interesting also for
spotlighting issues as the risk-based assessment of
applications, trustworthiness, ethical approach and AI
application lifecycle. In addition, in each section of the
roadmap, specific needs in the direction of
standardization are pinpointed.
The European Commission (EC) shaped a white
paper [138] setting out policies to achieve the uptake of
AI in the European Union (EU) and to address risks
associated with the use of AI technology. Along the
sections of the document, which refrains from
addressing the development and use of AI for military
purposes, particular attention is given to the opportunity
to create an ecosystem of excellence. Six actions have
been highlighted, among which: focusing on SMEs and
ensuring that each member state has a digital hub highly
specialized in AI; strengthening public-private
partnerships in AI, data and robotics; and promoting the
use of AI in the public sector. An overview of the most
IAC-21-B.1.4.4 Page 18 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
significant risks is also provided, with more emphasis
on ethical and trustworthy AI.
The National Science and Technology Council from
the USA’s Executive Office developed an AI Research
Development Plan in 2016, later updated in 2019 [139].
The Plan does not define specific research agendas for
Federal agency investments but highlights strategies to
reach a given portfolio. While it must be noted that the
utmost focus of the strategies is not on the
standardization, strategy 4 "Ensure the Safety and
Security of AI Systems'' and Strategy 6 "Measure and
Evaluate AI Technologies through Standards and
Benchmarks" are covering aspects strictly related to
standards and certifiability. It is worth mentioning great
attention to the development of shared public datasets
and open source libraries, as means to keep the
technological lead.
Although slightly different in scope, as more
oriented towards certification rather than
standardization, it is worth mentioning the DEEL White
Paper [140]. The document aims at “sharing knowledge,
identifying challenges for the certification of systems
using ML, and fostering the research effort”. A
thorough discussion on the features that an ML-based
system should possess to be certified is carried on,
leading to the identification of seven challenges to
tackle: probabilistic assessment, resilience,
specifiability, data quality, explainability, robustness,
and verifiability.
7.2. Issues and Techniques
In this section, we offer a brief discussion of the
potential unique issues one may encounter when
developing and operating a system that incorporates
ML. Whenever possible, we discuss some current
approaches to bridge these issues. This discussion is
meant to be illustrative to the reader and an
encouragement to explore the topics in further detail,
but it attempts to be comprehensive on neither scope nor
depth. Furthermore, the topic is under active research
and is likely to expand in the coming years.
7.2.1. Explainability
ML models, and particularly large models with lots
of free parameters such as large decision trees or NNs,
can act as black boxes. The process by which they
arrive at the final output can be too complex to be
directly interpreted, thus becoming as inscrutable as if
the model’s internals had been inaccessible in the first
place.
However, transparency, explainability, and
interpretability are very important for any technical
system with a moderate or large impact, be it in terms of
dollars or human lives. Therefore model explainability
is very important in fields such as aerospace, medicine,
insurance, banking, and more.
Explainability is a hard problem because of several
reasons. Firstly, it is user-dependent: the type of
explanation expected for example by an average user
will differ from that expected by a regulator or an
engineer. This leads to the question “How detailed must
the explanation be, and what must it cover?”. Secondly,
the expected outcome of transmitting an explanation can
be hard to define: the questions to answers are “Must
the receiver become more able to predict model output
after receiving explanations? Must the explanation point
univocally to the features of the input data that had the
largest impact on the produced results, and is this
limited to input data or does it also include training
data? Perhaps it should illustrate a counterfactual -
«What would need to change for the decision to have
been different?»? Or perhaps something else entirely?
And are the previous goals mutually exclusive?”
There are a huge number of techniques to answer
some of these and related questions. The field of
explainable AI or XAI for short is huge and expanding
rapidly. Providing an overview of this field is not within
the scope of the current publication, but we recommend
our readers to consult the Interpretable Machine
Learning book [141] or one of the numerous reviews on
the topic to learn more [142], [143].
7.2.2. Robustness and reliability
Reliability is the rate of failure of a system when
operating in nominal conditions (e.g. 10-9 catastrophic
failures per flight hour [144]). Since a rate of system
error can be extremely challenging to calculate without
operating the system, heuristic development rules like
no single point-of-failure are accepted as valid ways to
achieve the goal. This acceptance stems from either a
competent authority, which implicitly accepts the risk of
not properly achieving the desired reliability level or
historical data when available. Neither is a possibility
for current ML-based systems, due to an absence of
historically validated, robust, and widely accepted
heuristic design rules.
For Machine Learning systems, reliability comes
from two distinct factors: accuracy and robustness. An
ML classifier with higher accuracy is less likely to
misclassify an input, hence is more reliable..
Performance does not usually come into play for
classical software system’s reliability as accuracy for a
valid set of inputs and execution path is 100%. This
section does not concern itself with increasing model
accuracy, a topic that is the main focus of each
application-specific research field mentioned so far.
Accuracy for an ML model is calculated over the
data points in the test dataset and only those. While this
IAC-21-B.1.4.4 Page 19 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
is also true for classical software testing, the notion of
input equivalence classes provides assurance that the
software system will continue to perform acceptably for
inputs outside the test set. Correctness equivalence
classes for ML models do not currently exist. A similar
notion of robustness can be used instead. A robust
model has bounded accuracy loss for inputs that are
within a bounded distance of the input distribution. This
fact can be used to construct arguments for the
reliability of an ML model.
Although classical software systems are also tested
on a discrete batch of test cases, equivalence classes are
used to demonstrate correctness against the entire range
of operating conditions. Equivalence classes are sets of
input data in which verification of a property
(correctness in this case) on one point implies (or is
assumed to imply) verification of said property on the
whole set. If a string processing system does not fail on
a generic string, it will not fail on other strings, save for
edge cases, which are also tested. That notion of input
equivalence does not currently exist for input to ML
systems. Equivalence class discovery for random forest
models is a topic under active research [136], [145].
When demonstrating model robustness, several
problems arise:
Firstly, how does one quantify the distance between
input data? Although several measures exist, they are
often hard to relate to the implicit notions of input
distance humans have. It is easier to qualitatively say to
what degree an image does not depict a cat than it is to
quantify it in a single measure. This only becomes
harder for more abstract forms of input such as satellite
telemetry data. Thus, relating system-level
specifications to notions of input distance is sometimes
complex. For a given distance definition, formal
verification methods attempt to formally prove certain
properties of DL models, including robustness
[146]–[149]. They allow a user to build a model tolerant
to a certain distance between inputs. Equivalent research
exists for other ML models, such as random forests
(RFs) [136], but the literature is significantly less
developed. Note that these approaches allow a designer
to fight adversarial examples, a specific and concerning
failure mode for ML models [150], [151]. Nonetheless,
the literature on the generation and defeat of adversarial
examples is highly active and ever-evolving, as
measures and counter-measures and
counter-counter-measures get deployed. It is out of
scope for this review to delve any deeper into that.
Secondly, there is the well-known issue of
generalization. A model may offer very good
performance on a dataset and very poor performance on
the actual population, in the phenomenon known as
overfitting. The PAC-Bayes approach offers
generalization bounds that specify a minimum number
of samples from distribution for a desired performance
and training process reliability levels within that
distribution. These bounds, however, are often
extremely conservative, and improving them is another
active field of research [152]. Since it is hard to
appropriately quantify these bounds, the only recourse
for organisations to ensure performance is to collect
massive amounts of data, which is prohibitively
expensive or downright impossible in many cases.Since
generalization and robustness shortcomings are highly
model-specific, one approach to tackle them focuses on
applying mixtures of models working in tandem, known
as ensemble models, and selecting an output based on
the collective response of the ensemble [153], [154].
Thirdly, and also related to the second issue, there is
the phenomenon of domain drift [155]. Models do not
just overfit to a given dataset but also to the current
population. And, as time goes by, systems change. An
FDIR system monitoring battery health will see its
voltage decrease over time as the battery ages, and the
statistical distribution of deviations around the nominal
value is likely to change too. The performance of the
ML model will thus decrease over time as the world
changes around it. Fine-tuning on new data causes the
phenomenon known as catastrophic forgetting [156],
where the model loses performance on old and new
data. A solution is to retrain it from scratch on new data,
but this entails capturing that data and retraining the
model, which increases operating costs and risks in hard
to predict ways. Alternative solutions exist but they
come with their own drawbacks. Training a model on a
dataset representative of the whole system’s life cycle
can mitigate the issue, but requires larger models and
better data capture at the project’s start.
Lastly, models also overfit the specifics of the
system they’re trained for. A model trained for one
specific satellite may have issues adapting to another
satellite instance, or model. Version improvements such
as equipment changes may bring performance hits with
them too. While research fields like transfer learning,
domain adaptation and domain generalization [157]
attempt to address the issue, they are far from
universally reliable at the moment. This is particularly
concerning for the space industry, where mass
manufacturing and standardized equipment is the
exception rather than the norm and can pose a serious
challenge to the industry’s adoption of ML
technologies. Sometimes, when adapting to new
platforms, new input data will be available or new
output data may be required. In this case, the field of
transfer learning is applicable, which includes both
domain adaptation and domain generalization.
In short, despite the aforementioned techniques, ML
models are extremely brittle to deviations in input data
from the training dataset, and it can be assumed that
deviations from the training dataset will break the
system. Therefore, building propper datasets is a key
IAC-21-B.1.4.4 Page 20 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
task of any ML system designer or operator, a topic
which we address in the next section.
7.2.3. Dataset Construction
Datasets are the lifeblood of ML. Therefore it is only
right to have standards assigned for data to avoid
anomalies and have a perfect collection that will help
produce the right results.
Cappi et al. [158] propose a Dataset Definition
Standard (DDS) which, while not specifically geared
towards space activities, can be applied to EO data from
either payload or satellites. It aims to provide a standard
for training, validating or testing datasets. It explains in
detail the recommendations to be followed not only
while collecting data but also how to annotate it and
perform functions. The paper talks about many
important aspects any dataset should possess, from how
it must cover as many situations as possible that could
be encountered during model development to how a
history of every single change to every data must be
kept to help with traceability and avoid discrepancy.
The paper provides clear recommendations for labelling
and annotation of data and how the dataset should be
segregated for training, validation and testing.
The US Geological Survey [159] provides dataset
standards for their various operations like Biological,
Climate and Forecast and Mapping. Cleansing “dirty
data” is mentioned as a common problem faced by data
scientists. They also take it a step further with
geological mapping by producing a set of Parameter
standards to be followed while collecting data which
define a set of rules for individual parameters within the
dataset. The parameter standards cover a wide range of
qualities like the date/time, geographic coordinates,
codes, etc. the satellite data should contain. Report
[159] explains how exactly a topographical map of
anything in the US should be produced and one
important aspect of it is the data standards including
which standards the file formats of the data should be
stored in.
For the data quality standards, they delve into it by
discussing various components like currency,
consistency, completeness, and accuracy. The paper
covers every aspect of mapping data from dealing with
off grid and oversized maps, data sources and
resolutions to how cartographic features should be
interpreted.
GIS, remote sensing and satellite imagery are the
primary space operations for Earth. Hence they are also
the primary data-producing activities. Therefore the
standards for remote sensing are relatively well
developed compared to any other ML operations in the
space industry. [160] and [161] go in-depth about all
standards of remote sensing including the dataset
standards.
Di and Kobler [161] introduce NASA’s well
developed EO Systems’ Data Information Systems
(EOSDIS). As the EOSDIS will process data from
various fields it is not feasible for the system to deal
with every single data collected one by one. This has led
to EOSDIS establishing standards to deal specifically
with remote sensing data.
7.3. Recommendations
As outlined above, ML systems face a number of
issues precluding their application in many scenarios
where they would otherwise be useful.
We believe the fundamental research being carried
out on ML model robustness is of great interest and
recommend that any practitioner follow it closely. For
certain small-scale problems, work on formal
verification of ML models may already be enough to
ascertain that the network responds appropriately within
the input regime, and input data outside of this regime
can be purged by data verification systems implemented
in classical software. Further, we recommend that any
practitioner keep a careful watch for ways in which the
lifecycle operation of a system will deviate from the
training scenarios, and mitigate the risks issued from
model brittleness to these differences.
ML explainability is another core issue,
explainability of model decisions can and does take
precedence over model performance in scenarios with
high impact decisions, or where (human) learning from
the model’s decisions is key. Current model
explainability methods can offer insight into the
relevant features of input data used for a model’s
decision, but they can also provide misleading or
unhelpful signals. For applications where explainability
is an important feature of the system, dictionary, tree, or
kernel-based models and other easily explainable
methods should be compared with harder to explain
models for a performance-explainability trade-off.
National recommendations, white papers and initial
official standards in the AI and ML field attest to the
growing interest in the subject. While the scope of these
is much broader than the space sector alone, some
considerations can be applied to ML for space
applications too.
Data quality and availability will play an important
role in the adoption of ML across EO Operations, and
will certainly be demanded by supervisory and
regulatory agencies performing standardization and
certification. This need goes beyond the mere
abundance of data. Relevance, cleanliness, and
useability will require careful attention and control. To
achieve this, the industry can leverage work from other
fields such as the aforementioned dataset standards.
IAC-21-B.1.4.4 Page 21 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
8. Discussion & Conclusion
This review work covered different domains in the
field of ML in EO Operations, such as ground
operations, enhanced GNC, on-board image processing,
FDIR and standardization. It analyzed the field’s status,
which can be thought of as a baseline, and highlighted
interesting trends.
We have found that growing evidence across many
application fields show that EO missions can benefit
from onboard usage of ML. Found case studies have
shown improvements in both performance and
autonomy of platforms. New capabilities such as
automated filtering of payload data by transmitting to
ground only relevant images can reduce downlink
bandwidth requirements, which is important for small
satellites but also mitigates saturation of
radio-frequency bands. Enhanced visual processing also
enables better visual-based navigation for spacecraft,
and RL holds promise as the key to unlock better
nonlinear controllers. Autonomous FDIR operations
enable better autonomous decision making for EO
missions making it possible for existing teams to
manage more operations in a leaner way, reducing
satellite operating costs.
ML algorithms place high (and growing)
requirements on onboard computers. This increases the
size, weight, and power of the platform, leaving less
room available for payload. The issue of optimizing ML
models both at the software and hardware level for
space applications emerges as a key challenge.
Fortunately for space platform operators this mirrors
and is an instance of, the much larger challenge of
deploying ML on edge platforms. The community can
leverage a huge and expanding pool of research and
experience.
From the on-board EO application perspective, ML
has been applied mainly to cloud detection and
novelty/change detection. Many examples of these
applications rely on vision-based solutions.
Vision-based ML algorithms and solutions represent a
domain that has been extensively explored in other
technical fields, from which EO applications could
borrow technical knowledge. However, there are few
examples also for SAR-based images and this could
potentially suggest that there is room for improvement
and development with these sensors in all-weather
all-day conditions.
On a parallel, yet strongly intertwined, track to
research and applications there is standardization.
Specific standards for ML in the space sector are yet to
be set. Rigorous research is being carried out on key
topics such as explainability, robustness, and data
structure construction. EO Operators developing ML
applications should leverage this field for increased
reliability and performance. Organizations looking to
emit standards and guidelines should also take into
consideration these research areas while avoiding
overprescribing solutions that could hinder the
effectiveness of standardization development.
Data availability represents another key factor that
links together all the ML-based EO Operations. Having
a larger amount of data to use for ML in EO Operations
could have a tremendous impact in fostering
technological progress, yielding benefits to all players,
from industry to academia to space agencies. Currently,
though, there are few open datasets and most are
specifically tailored for image processing applications
or visual navigation applications. While disclosing
private data can be against a single organization’s
short-term interests, the technological advancement
favoured by open datasets in a broader range of
applications is a very valuable long-term goal for the
space sector as a whole. We believe the field should
focus on creating and publicizing such open datasets,
and we encourage actors not subject to the profit
motive, like space agencies, to take a leading role in
achieving such a goal.
One recurrent finding across the review’s subtopics
is that, at the moment, few missions have used ML in
orbit for EO missions. This contrasts with other
industries such as the automotive sector, that have
embraced the technology, and can be attributed to the
long lead times and slow cycles of the space industry.
With the development of New Space and faster access
to space, we hope that these cycles can be accelerated as
technology demonstrators move in a short time from
test benches to orbit. Additionally, increasing the
number of missions can lead to improved data
collection and platform standardization, further
improving the effectiveness of ML deployments in the
EO Operations sector and the space industry at large.
Last but not least we have noticed that the
application of ML for EO operations is, often, behind
the ML state of the art. A number of techniques that
have seen thorough success in both research and
operational environments have yet to be publicly
applied to EO Operations problems, such as
transformer models on sequential and other data.
Likewise, we have found little evidence for the adoption
of Neural Architecture Search techniques, despite being
well suited to adapt the computational cost of NN to
resource-constrained platforms and achieve higher
performance. Similarly, very few applications can be
found also for the use of resource reduction techniques
like pruning, distillation, or quantization and with
regards to formal verification and other certified
robustness techniques.
We hope this review will be a useful resource to
researchers and operators for further deployment and
experimentation of ML in future complex, demanding
IAC-21-B.1.4.4 Page 22 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
EO missions that are more autonomous, transmit only
useful data, and require less involvement.
Acknowledgements
We were supported by three partner organisations:
Mindseed, Satsure, and CNES.
Mindseed [162] is a company focused on bridging
the gap between space and non-space organisations and
showcasing the benefits of space for all.
Satsure [163] is an innovative decision analytics
company leveraging advances in satellite remote
sensing and Machine Learning to achieve the United
Nations Sustainable Development Goals.
CNES [164] is the French National Space Agency,
with activities all over the space value chain. Their
Earth Observation Operations teams provided
invaluable feedback for our research.
Our three partners reviewed our research, and we are
deeply thankful for their collaboration.
References
[1] Q. Guo, B. Fu, P. Shi, T. Cudahy, J. Zhang, and
H. Xu, ‘Satellite Monitoring the
Spatial-Temporal Dynamics of Desertification
in Response to Climate Change and Human
Activities across the Ordos Plateau, China’,
Remote Sens., vol. 9, no. 6, Art. no. 6, Jun.
2017, doi: 10.3390/rs9060525.
[2] ‘Copernicus air quality index’, euronews.
https://www.euronews.com/embed/weather/air-q
uality (accessed Sep. 18, 2021).
[3] G. Pan, Y. Xu, and J. Ma, ‘The potential of CO2
satellite monitoring for climate governance: A
review’, J. Environ. Manage., vol. 277, p.
111423, Jan. 2021, doi:
10.1016/j.jenvman.2020.111423.
[4] A. S. Belward and J. O. Skøien, ‘Who launched
what, when and why; trends in global
land-cover observation capacity from civilian
earth observation satellites’, ISPRS J.
Photogramm. Remote Sens., vol. 103, pp.
115–128, May 2015, doi:
10.1016/j.isprsjprs.2014.03.009.
[5] A. Krizhevsky, I. Sutskever, and G. E. Hinton,
‘ImageNet Classification with Deep
Convolutional Neural Networks’, in Advances
in Neural Information Processing Systems,
2012, vol. 25. Accessed: Sep. 18, 2021.
[Online]. Available:
https://proceedings.neurips.cc/paper/2012/hash/
c399862d3b9d6b76c8436e924a68c45b-Abstract
.html
[6] B. Wang, Mesh-Transformer-JAX:
Model-Parallel Implementation of Transformer
Language Model with JAX. 2021. [Online].
Available:
https://github.com/kingoflolz/mesh-transformer-
jax
[7] V. Mnih, K. Kavukcuoglu, D. Silver, A. Graves,
I. Antonoglou, D. Wierstra, and M. Riedmiller,
‘Playing Atari with Deep Reinforcement
Learning’, ArXiv13125602 Cs, Dec. 2013,
Accessed: Sep. 18, 2021. [Online]. Available:
http://arxiv.org/abs/1312.5602
[8] F. Fourati and M.-S. Alouini, ‘Artificial
Intelligence for Satellite Communication: A
Review’, ArXiv210110899 Cs Eess, Jan. 2021,
Accessed: Feb. 20, 2021. [Online]. Available:
http://arxiv.org/abs/2101.10899
[9] ‘2018 - Mission Planning and Scheduling.pdf’.
Accessed: Dec. 14, 2020. [Online]. Available:
https://public.ccsds.org/Pubs/529x0g1.pdf
[10] X. Wang, G. Wu, L. Xing, and W. Pedrycz,
‘Agile Earth observation satellite scheduling
over 20 years: formulations, methods and future
directions’, IEEE Syst. J., vol. 15, no. 3, pp.
3881–3892, Sep. 2021, doi:
10.1109/JSYST.2020.2997050.
[11] A. Globus, J. Crawford, J. Lohn, and A. Pryor,
‘Scheduling Earth Observing Satellites with
Evolutionary Algorithms’, Dec. 2003.
[12] M. A. A. Mansour and M. M. Dessouky, ‘A
genetic algorithm approach for solving the daily
photograph selection problem of the SPOT5
satellite’, Comput. Ind. Eng., vol. 58, no. 3, pp.
509–520, Apr. 2010, doi:
10.1016/j.cie.2009.11.012.
[13] J. Li, J. Li, H. Chen, and N. Jing, ‘A data
transmission scheduling algorithm for
rapid-response earth-observing operations’,
Chin. J. Aeronaut., vol. 27, no. 2, pp. 349–364,
Apr. 2014, doi: 10.1016/j.cja.2014.02.014.
[14] J. Crawford, J. Lohn, and A. Pryor, ‘A
Comparison of Techniques for Scheduling Earth
Observing Satellites’, p. 8.
[15] J. Bonnet, M.-P. Gleizes, E. Kaddoum, S.
Rainjonneau, and G. Flandin, ‘Multi-satellite
Mission Planning Using a Self-Adaptive
Multi-agent System’, in 2015 IEEE 9th
International Conference on Self-Adaptive and
Self-Organizing Systems, Cambridge, MA,
USA, Sep. 2015, pp. 11–20. doi:
10.1109/SASO.2015.9.
[16] H. Wang, ‘Online scheduling of image satellites
based on neural networks and deep
reinforcement learning’, p. 9.
[17] S. Peng, H. Chen, C. Du, J. Li, and N. Jing,
‘Onboard Observation Task Planning for an
Autonomous Earth Observation Satellite Using
IAC-21-B.1.4.4 Page 23 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
Long Short-Term Memory’, IEEE Access, vol.
6, pp. 65118–65129, 2018, doi:
10.1109/ACCESS.2018.2877687.
[18] Y. Du, T. Wang, B. Xin, L. Wang, Y. Chen, and
L. Xing, ‘A Data-Driven Parallel Scheduling
Approach for Multiple Agile Earth Observation
Satellites’, IEEE Trans. Evol. Comput., vol. 24,
no. 4, pp. 679–693, Aug. 2020, doi:
10.1109/TEVC.2019.2934148.
[19] X. Liu, ‘Mission schedule of agile satellites
based on Proximal Policy Optimization
Algorithm’, ArXiv200702352 Cs, Jul. 2020,
Accessed: Sep. 28, 2021. [Online]. Available:
http://arxiv.org/abs/2007.02352
[20] A. Hadj-Salah, R. Verdier, C. Caron, M. Picard,
and M. Capelle, ‘Schedule Earth Observation
satellites with Deep Reinforcement Learning’,
ArXiv191105696 Cs, Nov. 2019, Accessed: Sep.
28, 2021. [Online]. Available:
http://arxiv.org/abs/1911.05696
[21] A. Hadj-Salah, J. Guerra, M. Picard, and M.
Capelle, ‘Towards operational application of
Deep Reinforcement Learning to Earth
Observation satellite scheduling’. Aug. 30,
2020. Accessed: Sep. 28, 2021. [Online].
Available:
https://hal.archives-ouvertes.fr/hal-02925740
[22] Y. Song, Z. Zhou, Z. Zhang, F. Yao, and Y.
Chen, ‘A framework involving MEC: imaging
satellites mission planning’, Neural Comput.
Appl., vol. 32, no. 19, pp. 15329–15340, Oct.
2020, doi: 10.1007/s00521-019-04047-6.
[23] S.-F. Wu, C. J. H. Engelen, Q.-P. Chu, R.
Babuška, J. A. Mulder, and G. Ortega, ‘Fuzzy
logic based attitude control of the spacecraft
X-38 along a nominal re-entry trajectory’,
Control Eng. Pract., vol. 9, no. 7, pp. 699–707,
Jul. 2001, doi:
10.1016/S0967-0661(01)00036-3.
[24] G. Ortega, ‘Fuzzy logic techniques for
rendezvous and docking of two geostationary
satellites’, Telemat. Inform., vol. 12, no. 3, pp.
213–227, Jun. 1995, doi:
10.1016/0736-5853(95)00013-5.
[25] C.-H. Cheng, S.-L. Shu, and P.-J. Cheng,
‘Attitude control of a satellite using fuzzy
controllers’, Expert Syst. Appl., vol. 36, no. 3,
Part 2, pp. 6613–6620, Apr. 2009, doi:
10.1016/j.eswa.2008.08.053.
[26] D. Izzo, M. Märtens, and B. Pan, ‘A Survey on
Artificial Intelligence Trends in Spacecraft
Guidance Dynamics and Control’,
ArXiv181202948 Cs, Dec. 2018, Accessed: Dec.
12, 2020. [Online]. Available:
http://arxiv.org/abs/1812.02948
[27] D. Izzo and E. Öztürk, ‘Real-Time Guidance for
Low-Thrust Transfers Using Deep Neural
Networks’, J. Guid. Control Dyn., vol. 44, no. 2,
pp. 315–327, Feb. 2021, doi:
10.2514/1.G005254.
[28] L. Federici, B. Benedikter, and A. Zavoli, ‘Deep
Learning Techniques for Autonomous
Spacecraft Guidance During Proximity
Operations’, J. Spacecr. Rockets, vol. 0, no. 0,
pp. 1–12, doi: 10.2514/1.A35076.
[29] K. Hovell and S. Ulrich, ‘Deep Reinforcement
Learning for Spacecraft Proximity Operations
Guidance’, J. Spacecr. Rockets, vol. 58, no. 2,
pp. 254–264, 2021, doi: 10.2514/1.A34838.
[30] D. Lee, ‘Guidance, navigation and control
system for autonomous proximity operations
and docking of spacecraft’, p. 149.
[31] K. Hovell and S. Ulrich, ‘On Deep
Reinforcement Learning for Spacecraft
Guidance’, in AIAA Scitech 2020 Forum,
American Institute of Aeronautics and
Astronautics. doi: 10.2514/6.2020-1600.
[32] S. Koizumi, Y. Kikuya, K. Sasaki, Y. Masuda,
Y. Iwasaki, K. Watanabe, Y. Yatsu, and S.
Matsunaga, ‘Development of Attitude Sensor
using Deep Learning’, p. 8.
[33] S.-W. Kim, S.-Y. Park, and C. Park, ‘Spacecraft
attitude control using neuro-fuzzy
approximation of the optimal controllers’, Adv.
Space Res., vol. 57, no. 1, pp. 137–152, Jan.
2016, doi: 10.1016/j.asr.2015.09.016.
[34] A. Joghataie, ‘Neural Networks and Fuzzy
Logic for Structural Control’, University of
Illinois Engineering Experiment Station.
College of Engineering. University of Illinois at
Urbana-Champaign, 1994. [Online]. Available:
http://hdl.handle.net/2142/14218
[35] M. Azarbad, H. Azami, S. Sanei, and A.
Ebrahimzadeh, ‘New Neural Network-based
Approaches for GPS GDOP Classification
based on Neuro-Fuzzy Inference System, Radial
Basis Function, and Improved Bee Algorithm’,
Appl. Soft Comput., vol. 25, pp. 285–292, Dec.
2014, doi: 10.1016/j.asoc.2014.09.022.
[36] ‘Comparative study of Classical and Fuzzy PID
Attitude Control System with Extended Kalman
Filter feedback for nanosatellites. — IAF
Digital Library’.
https://dl.iafastro.directory/event/IAC-2018/pap
er/44553/ (accessed Sep. 14, 2021).
[37] Y. WANG, Z. MA, Y. YANG, Z. WANG, and L.
TANG, ‘A New Spacecraft Attitude
Stabilization Mechanism Using Deep
Reinforcement Learning Method’, p. 13 pages,
2019, doi: 10.13009/EUCASS2019-33.
[38] D. Yadava, R. Hosangadi, S. Krishna, P.
Paliwal, and A. Jain, ‘Attitude control of a
IAC-21-B.1.4.4 Page 24 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
nanosatellite system using reinforcement
learning and neural networks’, in 2018 IEEE
Aerospace Conference, Big Sky, MT, Mar. 2018,
pp. 1–8. doi: 10.1109/AERO.2018.8396409.
[39] S. Jaekel and B. Scholz, ‘Utilizing Artificial
Intelligence to achieve a robust architecture for
future robotic spacecraft’, in 2015 IEEE
Aerospace Conference, Big Sky, MT, Mar. 2015,
pp. 1–14. doi: 10.1109/AERO.2015.7119180.
[40] S. Fuertes, B. Pilastre, and S. D’Escrivan,
‘Performance assessment of NOSTRADAMUS
& other machine learning-based telemetry
monitoring systems on a spacecraft anomalies
database’, presented at the 2018 SpaceOps
Conference, Marseille, France, May 2018. doi:
10.2514/6.2018-2559.
[41] Ivano Verzola, Alessandro Donati, J.-A.
Martínez Heras, M. Schubert, and Laszlo
Somodi, ‘Project Sybil : A Novelty Detection
System for Human Spaceflight Operations, in :
Proc. Int. Conf. Space Operations’, presented at
the SpaceOps, 2016.
[42] C. O’Meara, L. Schlag, L. Faltenbacher, and M.
Wickler, ‘ATHMoS: Automated Telemetry
Health Monitoring System at GSOC using
Outlier Detection and Supervised Machine
Learning’, May 2016. doi:
10.2514/6.2016-2347.
[43] B. Pilastre, ‘Estimation parcimonieuse et
apprentissage de dictionnaires pour la détection
d’anomalies multivariées dans des données
mixtes de télémesure satellites’, phd, 2020.
Accessed: Sep. 13, 2021. [Online]. Available:
https://oatao.univ-toulouse.fr/27738/
[44] D. L. Iverson, ‘System Health Monitoring for
Space Mission Operations’, in 2008 IEEE
Aerospace Conference, Big Sky, MT, USA, Mar.
2008, pp. 1–8. doi:
10.1109/AERO.2008.4526646.
[45] B. W. Ricks, ‘METHODS FOR
PROBABILISTIC FAULT DIAGNOSIS: AN
ELECTRICAL POWER SYSTEM CASE
STUDY’, Mar. 2021, vol. Vol 1. No. 1.
[46] J.-G. Meß, ‘Techniques of Artificial Intelligence
for Space Applications - A Survey’, presented at
the European Workshop on On-Board Data
Processing (OBDP2019), Feb. 2019.
[47] J. Schumann, O. J. Mengshoel, and T. Mbaya,
‘Integrated Software and Sensor Health
Management for Small Spacecraft’, in 2011
IEEE Fourth International Conference on Space
Mission Challenges for Information Technology,
Palo Alto, CA, USA, Aug. 2011, pp. 77–84. doi:
10.1109/SMC-IT.2011.25.
[48] Ruonan Liu, Boyuan Yang, Enrico Zio, and
Xuefeng Chen, ‘Artificial intelligence for fault
diagnosis of rotating machinery: A review’.
[49] S.P.M (Sander) Voss, ‘Application of deep
learning for spacecraft fault detection and
isolation’, Master Thesis.
[50] D. Codetta-Raiteri and L. Portinale, ‘Dynamic
Bayesian Networks for Fault Detection,
Identification, and Recovery in Autonomous
Spacecraft’, IEEE Trans. Syst. Man Cybern.
Syst., vol. 45, no. 1, pp. 13–24, Jan. 2015, doi:
10.1109/TSMC.2014.2323212.
[51] ‘Periodic Reporting for period 1 - MiRAGE
(Enhanced mission autonomy through on-board
Artificial Intelligence for next generation
spacecraft) | H2020 | CORDIS | European
Commission’.
https://cordis.europa.eu/project/id/816442/report
ing (accessed Sep. 19, 2021).
[52] ‘Georges Labrèchea, Tanguy Sotoa, Evridiki
Ntagioua, David Evansa, In-Flight Training of a
FDIR Model with Online Machine Learning on
the OPS-SAT Spacecraft, 2021-
SpaceOps-2021,13,x1450’.
https://github.com/georgeslabreche/opssat-orbita
i/find/main (accessed Sep. 12, 2021).
[53] olanleed, MochiMochi. 2021. Accessed: Sep.
29, 2021. [Online]. Available:
https://github.com/olanleed/MochiMochi
[54] S. Mahajan and B. Fataniya, ‘Cloud detection
methodologies: variants and development—a
review’, Complex Intell. Syst., vol. 6, no. 2, pp.
251–261, Jul. 2020, doi:
10.1007/s40747-019-00128-0.
[55] Z. Zhang, A. Iwasaki, G. Xu, and J. Song,
‘Cloud detection on small satellites based on
lightweight U-net and image compression’, J.
Appl. Remote Sens., vol. 13, no. 2, p. 026502,
Apr. 2019, doi: 10.1117/1.JRS.13.026502.
[56] R. Hinz, J. I. Bravo, M. Kerr, C. Marcos, A.
Latorre, and F. Membibre, ‘EO-ALERT:
Machine Learning-Based On-Board Satellite
Processing for Very-Low Latency Convective
Storm Nowcasting’, p. 1.
[57] A. N. Srivastava, Onboard Detection of Snow,
Ice, Clouds and Other Geophysical Processes
Using Kernel Methods. 2003.
[58] G. Giuffrida, L. Diana, F. de Gioia, G. Benelli,
G. Meoni, M. Donati,and L. Fanucci,
‘CloudScout: A Deep Neural Network for
On-Board Cloud Detection on Hyperspectral
Images’, Remote Sens., vol. 12, no. 14, Art. no.
14, Jan. 2020, doi: 10.3390/rs12142205.
[59] A. Maskey and M. Cho, ‘CubeSatNet: Ultralight
Convolutional Neural Network designed for
on-orbit binary image classification on a 1U
CubeSat’, Eng. Appl. Artif. Intell., vol. 96, p.
103952, Nov. 2020, doi:
IAC-21-B.1.4.4 Page 25 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
10.1016/j.engappai.2020.103952.
[60] M. Ireland, ‘Integrating AI Techniques Into
Future Nanosatellite Onboard Data Processing’,
p. 30.
[61] R. Castaño, S. A. Chien, K. L. Wagstaff, and T.
M. Stough, ‘On-board analysis of uncalibrated
data for a spacecraft at mars’, presented at the
International Conference on Knowledge
Discovery and Data Mining, San Jose,
California, USA, Aug. 2007. doi:
10.1145/1281192.1281291.
[62] F. Férésin, E. Kervennic, Y. Bobichon, E.
Lemaire, N. Abderrahmane, G. Bahl, I. Grenet,
M. Moretti, and M. Benguigui, ‘In space image
processing using AI embedded on system on
module : example of OPS-SAT cloud
segmentation’, presented at the 2nd European
Workshop on On-Board Data Processing, 2021.
Accessed: Sep. 28, 2021. [Online]. Available:
https://www.irt-saintexupery.com/two-premieres
-in-on-board-artificial-intelligence-on-satellites/
[63] K. L. Wagstaff, S. Chien, A. Altinok, U.
Rebbapragada, D. Thompson, S. Schaffer, and
D. Tran, ‘Cloud Filtering and Novelty Detection
using Onboard Machine Learning for the EO-1
Spacecraft’, Int. Jt. Conf. Artif. Intell., p. 4,
2017.
[64] S. Chien, J. Doubleday, D. R. Thompson, K. L.
Wagstaff, J. Bellardo, C. Francis, E.
Baumgarten, A. Williams, E. Yee, E. Stanton
and J. Piug-Suari, ‘Onboard Autonomy on the
Intelligent Payload Experiment CubeSat
Mission’, J. Aerosp. Inf. Syst., vol. 14, no. 6, pp.
307–315, 2017, doi: 10.2514/1.I010386.
[65] P. Deepan and L. R. Sudha, ‘Object
Classification of Remote Sensing Image Using
Deep Convolutional Neural Network’, in The
Cognitive Approach in Cloud Computing and
Internet of Things Technologies for Surveillance
Tracking Systems, Elsevier, 2020, pp. 107–120.
doi: 10.1016/B978-0-12-816385-6.00008-8.
[66] A. P. Arechiga, A. J. Michaels, and J. T. Black,
‘Onboard Image Processing for Small
Satellites’, in NAECON 2018 - IEEE National
Aerospace and Electronics Conference, Jul.
2018, pp. 234–240. doi:
10.1109/NAECON.2018.8556744.
[67] J. Manning, D. Langerman, B. Ramesh, E.
Gretok, C. Wilson, A. George, J. MacKinnon,
and G. Crum, ‘Machine-Learning Space
Applications on SmallSat Platforms with
TensorFlow’, Small Satell. Conf., Aug. 2018,
[Online]. Available:
https://digitalcommons.usu.edu/smallsat/2018/al
l2018/458
[68] J. Goodwill, D. Wilson, S. Sabogal, A. D.
George, and C. Wilson, ‘Adaptively Lossy
Image Compression for Onboard Processing’, in
2020 IEEE Aerospace Conference, Mar. 2020,
pp. 1–15. doi:
10.1109/AERO47225.2020.9172536.
[69] T. Vladimirova and S. Atek, ‘A New Lossless
Compression Method for Small Satellite
On-Board Imaging’, University of Surrey
Guildford, Surrey, GU2 7 XH United Kingdom,
2002. doi: 10.1142/9789812776266_0038.
[70] Y. Cai, Y. Hu, M. Siegel, S. J. Gollapalli, A. R.
Venugopal, and U. Bardak, ‘Onboard Feature
Indexing from Satellite Lidar Images’, p. 4.
[71] N. D. Lane, S. Bhattacharya, P. Georgiev, C.
Forlivesi, and F. Kawsar, ‘An Early Resource
Characterization of Deep Learning on
Wearables, Smartphones and Internet-of-Things
Devices’, in Proceedings of the 2015
International Workshop on Internet of Things
towards Applications, Seoul South Korea, Nov.
2015, pp. 7–12. doi: 10.1145/2820975.2820980.
[72] H. Han, S. Lee, J. Im, M. Kim, M.-I. Lee, M. H.
Ahn, and S.-R. Chung, ‘Detection of Convective
Initiation Using Meteorological Imager Onboard
Communication, Ocean, and Meteorological
Satellite Based on Machine Learning
Approaches’, Remote Sens., vol. 7, no. 7, Art.
no. 7, Jul. 2015, doi: 10.3390/rs70709184.
[73] D. Blalock, J. J. G. Ortiz, J. Frankle, and J.
Guttag, ‘What is the State of Neural Network
Pruning?’, ArXiv200303033 Cs Stat, Mar. 2020,
Accessed: Aug. 06, 2021. [Online]. Available:
http://arxiv.org/abs/2003.03033
[74] H. Wang, C. Qin, Y. Zhang, and Y. Fu,
‘Emerging Paradigms of Neural Network
Pruning’, ArXiv210306460 Cs, Mar. 2021,
Accessed: Aug. 06, 2021. [Online]. Available:
http://arxiv.org/abs/2103.06460
[75] J. Frankle, G. K. Dziugaite, D. M. Roy, and M.
Carbin, ‘Pruning Neural Networks at
Initialization: Why are We Missing the Mark?’,
ArXiv200908576 Cs Stat, Mar. 2021, Accessed:
Aug. 06, 2021. [Online]. Available:
http://arxiv.org/abs/2009.08576
[76] T. He, Y. Fan, Y. Qian, T. Tan, and K. Yu,
‘Reshaping deep neural network for fast
decoding by node-pruning’, in 2014 IEEE
International Conference on Acoustics, Speech
and Signal Processing (ICASSP), Florence,
Italy, May 2014, pp. 245–249. doi:
10.1109/ICASSP.2014.6853595.
[77] Y. Tang, J. Ji, S. Gao, H. Dai, Y. Yu, and Y.
Todo, ‘A Pruning Neural Network Model in
Credit Classification Analysis’, Comput. Intell.
Neurosci., vol. 2018, pp. 1–22, 2018, doi:
10.1155/2018/9390410.
IAC-21-B.1.4.4 Page 26 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
[78] A. Lazarevic and Z. Obradovic, ‘Effective
pruning of neural network classifier ensembles’,
in IJCNN’01. International Joint Conference on
Neural Networks. Proceedings (Cat.
No.01CH37222), Washington, DC, USA, 2001,
vol. 2, pp. 796–801. doi:
10.1109/IJCNN.2001.939461.
[79] ‘Pruning in Keras example | TensorFlow Model
Optimization’, TensorFlow.
https://www.tensorflow.org/model_optimization
/guide/pruning/pruning_with_keras (accessed
Aug. 06, 2021).
[80] D. Browne, M. Giering, and S. Prestwich,
‘PulseNetOne: Fast Unsupervised Pruning of
Convolutional Neural Networks for Remote
Sensing’, Remote Sens., vol. 12, no. 7, Art. no.
7, Jan. 2020, doi: 10.3390/rs12071092.
[81] M. Castelluccio, G. Poggi, C. Sansone, and L.
Verdoliva, ‘Land Use Classification in Remote
Sensing Images by Convolutional Neural
Networks’, ArXiv150800092 Cs, Aug. 2015,
Accessed: Aug. 07, 2021. [Online]. Available:
http://arxiv.org/abs/1508.00092
[82] T. Kavzoglu and P. M. Mather, ‘Pruning
artificial neural networks: An example using
land cover classification of multi-sensor
images’, Int. J. Remote Sens., vol. 20, no. 14,
pp. 2787–2803, Jan. 1999, doi:
10.1080/014311699211796.
[83] E. Maggiori, Y. Tarabalka, G. Charpiat, and P.
Alliez, ‘Convolutional Neural Networks for
Large-Scale Remote-Sensing Image
Classification’, IEEE Trans. Geosci. Remote
Sens., vol. 55, no. 2, pp. 645–657, Feb. 2017,
doi: 10.1109/TGRS.2016.2612821.
[84] G. Pitsis, G. Tsagkatakis, C. Kozanitis, I.
Kalomoiris, A. Ioannou, A. Dollas, M. G.H.
Katevenis, and P. Tsakalides, ‘Efficient
Convolutional Neural Network Weight
Compression for Space Data Classification on
Multi-fpga Platforms’, in ICASSP 2019 - 2019
IEEE International Conference on Acoustics,
Speech and Signal Processing (ICASSP),
Brighton, United Kingdom, May 2019, pp.
3917–3921. doi:
10.1109/ICASSP.2019.8682732.
[85] S. Zhang, G. Wu, J. Gu, and J. Han, ‘Pruning
Convolutional Neural Networks with an
Attention Mechanism for Remote Sensing
Image Classification’, Electronics, vol. 9, no. 8,
p. 1209, Jul. 2020, doi:
10.3390/electronics9081209.
[86] Q. Tan, J. Ling, J. Hu, X. Qin, and J. Hu,
‘Vehicle Detection in High Resolution Satellite
Remote Sensing Images Based on Deep
Learning’, IEEE Access, vol. 8, pp.
153394–153402, 2020, doi:
10.1109/ACCESS.2020.3017894.
[87] N. Ma, X. Yu, Y. Peng, and S. Wang, ‘A
Lightweight Hyperspectral Image Anomaly
Detector for Real-Time Mission’, Remote Sens.,
vol. 11, no. 13, p. 1622, Jul. 2019, doi:
10.3390/rs11131622.
[88] A. Goel, C. Tung, Y.-H. Lu, and G. K.
Thiruvathukal, ‘A Survey of Methods for
Low-Power Deep Learning and Computer
Vision’, in 2020 IEEE 6th World Forum on
Internet of Things (WF-IoT), Jun. 2020, pp. 1–6.
doi: 10.1109/WF-IoT48130.2020.9221198.
[89] F. N. Iandola, S. Han, M. W. Moskewicz, K.
Ashraf, W. J. Dally, and K. Keutzer,
‘SqueezeNet: AlexNet-level accuracy with 50x
fewer parameters and <0.5MB model size’,
ArXiv160207360 Cs, Nov. 2016, Accessed: Sep.
30, 2021. [Online]. Available:
http://arxiv.org/abs/1602.07360
[90] A. G. Howard, M. Zhu, B. Chen, D.
Kalenichenko, W. Wang, T. Weyand, M.
Andreetto, and H. Adam, ‘MobileNets: Efficient
Convolutional Neural Networks for Mobile
Vision Applications’, ArXiv170404861 Cs, Apr.
2017, Accessed: Sep. 30, 2021. [Online].
Available: http://arxiv.org/abs/1704.04861
[91] ‘Multitask Classification of Remote Sensing
Scenes Using Deep Neural Networks | IEEE
Conference Publication | IEEE Xplore’.
https://ieeexplore.ieee.org/document/8518874
(accessed Sep. 30, 2021).
[92] ‘SqueezeNet with Attention for Remote Sensing
Scene Classification | IEEE Conference
Publication | IEEE Xplore’.
https://ieeexplore.ieee.org/abstract/document/90
96876 (accessed Sep. 30, 2021).
[93] ‘Multi-Scale Convolutional Neural Network for
Remote Sensing Scene Classification | IEEE
Conference Publication | IEEE Xplore’.
https://ieeexplore.ieee.org/abstract/document/85
00107 (accessed Sep. 30, 2021).
[94] A. Poortinga, N. S. Thwal, N. Khanal, T. Mayer,
B. Bhandari, K. Markert, A. P. Nicolau, J.
Dilger, K. Tenneson, N. Clinton, and D. Saah,
‘Mapping sugarcane in Thailand using transfer
learning, a lightweight convolutional neural
network, NICFI high resolution satellite
imagery and Google Earth Engine’, ISPRS Open
J. Photogramm. Remote Sens., vol. 1, p.
100003, Oct. 2021, doi:
10.1016/j.ophoto.2021.100003.
[95] B. Zhang, Y. Zhang, and S. Wang, ‘A
Lightweight and Discriminative Model for
Remote Sensing Scene Classification With
Multidilation Pooling Module’, IEEE J. Sel.
IAC-21-B.1.4.4 Page 27 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
Top. Appl. Earth Obs. Remote Sens., vol. 12, no.
8, pp. 2636–2653, Aug. 2019, doi:
10.1109/JSTARS.2019.2919317.
[96] D. Yu, Q. Xu, H. Guo, C. Zhao, Y. Lin, and D.
Li, ‘An Efficient and Lightweight Convolutional
Neural Network for Remote Sensing Image
Scene Classification’, Sensors, vol. 20, no. 7, p.
1999, Apr. 2020, doi: 10.3390/s20071999.
[97] T. Hoeser, F. Bachofer, and C. Kuenzer, ‘Object
Detection and Image Segmentation with Deep
Learning on Earth Observation Data: A
Review—Part II: Applications’, Remote Sens.,
vol. 12, no. 18, Art. no. 18, Jan. 2020, doi:
10.3390/rs12183053.
[98] X. He, K. Zhao, and X. Chu, ‘AutoML: A
survey of the state-of-the-art’, Knowl.-Based
Syst., vol. 212, p. 106622, Jan. 2021, doi:
10.1016/j.knosys.2020.106622.
[99] M. Chan, D. Scarafoni, R. Duarte, J. Thornton,
and L. Skelly, ‘Learning Network Architectures
of Deep CNNs Under Resource Constraints’, in
2018 IEEE/CVF Conference on Computer
Vision and Pattern Recognition Workshops
(CVPRW), Salt Lake City, UT, USA, Jun. 2018,
pp. 1784–17847. doi:
10.1109/CVPRW.2018.00222.
[100] B. Zoph and Q. V. Le, ‘Neural Architecture
Search with Reinforcement Learning’,
ArXiv161101578 Cs, Feb. 2017, Accessed: Sep.
18, 2021. [Online]. Available:
http://arxiv.org/abs/1611.01578
[101] B. Baker, O. Gupta, N. Naik, and R. Raskar,
‘Designing Neural Network Architectures using
Reinforcement Learning’, ArXiv161102167 Cs,
Mar. 2017, Accessed: Sep. 18, 2021. [Online].
Available: http://arxiv.org/abs/1611.02167
[102] T. Elsken, J. H. Metzen, and F. Hutter, ‘Neural
Architecture Search: A Survey’,
ArXiv180805377 Cs Stat, Apr. 2019, Accessed:
Sep. 18, 2021. [Online]. Available:
http://arxiv.org/abs/1808.05377
[103] W. Liang, J. Li, W. Diao, X. Sun, K. Fu, and Y.
Wu, ‘FGATR-Net: Automatic Network
Architecture Design for Fine-Grained Aircraft
Type Recognition in Remote Sensing Images’,
Remote Sens., vol. 12, no. 24, p. 4187, Dec.
2020, doi: 10.3390/rs12244187.
[104] M. Tan, B. Chen, R. Pang, V. Vasudevan, M.
Sandler, A. Howard, and Q. V. Le, ‘MnasNet:
Platform-Aware Neural Architecture Search for
Mobile’, in 2019 IEEE/CVF Conference on
Computer Vision and Pattern Recognition
(CVPR), Long Beach, CA, USA, Jun. 2019, pp.
2815–2823. doi: 10.1109/CVPR.2019.00293.
[105] L. J. Ba and R. Caruana, ‘Do Deep Nets Really
Need to be Deep?’, Dec. 2013, Accessed: Jun.
28, 2021. [Online]. Available:
https://arxiv.org/abs/1312.6184v7
[106] F. de Vieilleville, A. Lagrange, R. Ruiloba, and
S. May, ‘Towards Distillation of Deep Neural
Networks for Satellite On-Board Image
Segmentation’, Int. Arch. Photogramm. Remote
Sens. Spat. Inf. Sci., vol. 43, pp. 1553–1559,
2020.
[107] G. Chen, X. Zhang, X. Tan, Y. Cheng, F. Dai, K.
Zhu, Y. Gong, and Q. Wang, ‘Training Small
Networks for Scene Classification of Remote
Sensing Images via Knowledge Distillation’,
Remote Sens., vol. 10, no. 5, p. 719, May 2018,
doi: 10.3390/rs10050719.
[108] H. Bazzi, D. Ienco, N. Baghdadi, M. Zribi, and
V. Demarez, ‘Distilling Before Refine:
Spatio-Temporal Transfer Learning for Mapping
Irrigated Areas Using Sentinel-1 Time Series’,
IEEE Geosci. Remote Sens. Lett., vol. 17, no.
11, pp. 1909–1913, Nov. 2020, doi:
10.1109/LGRS.2019.2960625.
[109] Y. Chen, Y. Bian, X. Xiao, Y. Rong, T. Xu, and
J. Huang, ‘On Self-Distilling Graph Neural
Network’, ArXiv201102255 Cs Stat, Apr. 2021,
Accessed: Aug. 07, 2021. [Online]. Available:
http://arxiv.org/abs/2011.02255
[110] Y. Chai, K. Fu, X. Sun, W. Diao, Z. Yan, Y.
Feng, and L. Wang, ‘Compact Cloud Detection
with Bidirectional Self-Attention Knowledge
Distillation’, Remote Sens., vol. 12, no. 17, p.
2770, Aug. 2020, doi: 10.3390/rs12172770.
[111] V. Kothari, E. Liberis, and N. D. Lane, ‘The
Final Frontier: Deep Learning in Space’,
ArXiv200110362 Cs Eess, Feb. 2020, Accessed:
Jul. 15, 2021. [Online]. Available:
http://arxiv.org/abs/2001.10362
[112] R. Huq, M. Islam, and S. Siddique, ‘AI-OBC:
Conceptual Design of a Deep Neural Network
based Next Generation Onboard Computing
Architecture for Satellite Systems’, p. 8.
[113] M. Esposito, S. S. Conticello, M. Pastena, and
B. C. Domínguez, ‘In-orbit demonstration of
artificial intelligence applied to hyperspectral
and thermal sensing from space’, in CubeSats
and SmallSats for Remote Sensing III, Aug.
2019, vol. 11131, p. 111310C. doi:
10.1117/12.2532262.
[114] F. C. Bruhn, N. Tsog, F. Kunkel, and O. Flordal,
‘Enabling radiation tolerant heterogeneous
GPU‑based onboard data processing in space’,
Vol01234567891 3CEAS Space Journal, vol. 12,
pp. 551–564, 2020.
[115] S. Mittal, ‘A Survey on optimized
implementation of deep learning models on the
NVIDIA Jetson platform’, J. Syst. Archit., vol.
97, pp. 428–442, Aug. 2019, doi:
IAC-21-B.1.4.4 Page 28 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
10.1016/j.sysarc.2019.01.011.
[116] N. Buonaiuto, M. Louie, R. Mital, D. Mateik, R.
Sivilli, A. Bhopale, C. Kief, and B. Zufelt,
‘Satellite Identification Imaging for Small
Satellites Using NVIDIA’, p. 12, 2017.
[117] J. J. Hernández-Gómez, G. A. Yañez-Casas, A.
M. Torres-Lara, C. Couder-Castañeda, M. G.
Orozco-del-Castillo, J. C. Valdiviezo-Navarro, I.
Medina, A. Solís-Santomé, D.
Vázquez-Álvarez, and P. I. Chávez-López,
‘Conceptual low-cost on-board high
performance computing in CubeSat
nanosatellites for pattern recognition in Earth’s
remote sensing’, pp. 114–104. doi:
10.29007/8d25.
[118] J. Goodwill, G. Crum, J. MacKinnon, C.
Brewer, M. Monaghan, T. Wise, and C. Wilson,
‘NASA SpaceCube Edge TPU SmallSat Card
for Autonomous Operations and Onboard
Science-Data Analysis’, p. 13.
[119] Y. Guo, ‘A Survey on Methods and Theories of
Quantized Neural Networks’, ArXiv180804752
Cs Stat, Dec. 2018, Accessed: Sep. 18, 2021.
[Online]. Available:
http://arxiv.org/abs/1808.04752
[120] H. Qin, R. Gong, X. Liu, X. Bai, J. Song, and N.
Sebe, ‘Binary Neural Networks: A Survey’,
Pattern Recognit., vol. 105, p. 107281, Sep.
2020, doi: 10.1016/j.patcog.2020.107281.
[121] T. Simons and D.-J. Lee, ‘A Review of
Binarized Neural Networks’, Electronics, vol. 8,
no. 6, p. 661, Jun. 2019, doi:
10.3390/electronics8060661.
[122] M. Courbariaux, Y. Bengio, and J.-P. David,
‘BinaryConnect: Training Deep Neural
Networks with binary weights during
propagations’, in Advances in Neural
Information Processing Systems, 2015, vol. 28.
[Online]. Available:
https://proceedings.neurips.cc/paper/2015/file/3
e15cc11f979ed25912dff5b0669f2cd-Paper.pdf
[123] M. Rastegari, V. Ordonez, J. Redmon, and A.
Farhadi, ‘XNOR-Net: ImageNet Classification
Using Binary Convolutional Neural Networks’,
ArXiv160305279 Cs, Aug. 2016, Accessed: Sep.
18, 2021. [Online]. Available:
http://arxiv.org/abs/1603.05279
[124] C. Zhu, S. Han, H. Mao, and W. J. Dally,
‘Trained Ternary Quantization’,
ArXiv161201064 Cs, Feb. 2017, Accessed: Sep.
18, 2021. [Online]. Available:
http://arxiv.org/abs/1612.01064
[125] K. Simonyan and A. Zisserman, ‘Very Deep
Convolutional Networks for Large-Scale Image
Recognition’, ArXiv14091556 Cs, Apr. 2015,
Accessed: Sep. 18, 2021. [Online]. Available:
http://arxiv.org/abs/1409.1556
[126] S. Han, H. Mao, and W. J. Dally, ‘Deep
Compression: Compressing Deep Neural
Networks with Pruning, Trained Quantization
and Huffman Coding’, ArXiv151000149 Cs,
Feb. 2016, Accessed: Sep. 18, 2021. [Online].
Available: http://arxiv.org/abs/1510.00149
[127] ‘Post-training quantization | TensorFlow Lite’.
https://www.tensorflow.org/lite/performance/pos
t_training_quantization (accessed Sep. 28,
2021).
[128] ‘Quantization — PyTorch 1.9.1 documentation’.
https://pytorch.org/docs/stable/quantization.html
(accessed Sep. 28, 2021).
[129] ‘Larq | Binarized Neural Network
development’. https://larq.dev/ (accessed Sep.
28, 2021).
[130] A. G. Anderson and C. P. Berg, ‘The
High-Dimensional Geometry of Binary Neural
Networks’, ArXiv170507199 Cs, May 2017,
Accessed: Sep. 18, 2021. [Online]. Available:
http://arxiv.org/abs/1705.07199
[131] D. Molchanov, A. Ashukha, and D. Vetrov,
‘Variational Dropout Sparsifies Deep Neural
Networks’, ArXiv170105369 Cs Stat, Jun. 2017,
Accessed: Sep. 18, 2021. [Online]. Available:
http://arxiv.org/abs/1701.05369
[132] A. Polino, R. Pascanu, and D. Alistarh, ‘Model
compression via distillation and quantization’,
ArXiv180205668 Cs, Feb. 2018, Accessed: Sep.
18, 2021. [Online]. Available:
http://arxiv.org/abs/1802.05668
[133] Y. Gong, L. Liu, M. Yang, and L. Bourdev,
‘Compressing Deep Convolutional Networks
using Vector Quantization’, ArXiv14126115 Cs,
Dec. 2014, Accessed: Sep. 18, 2021. [Online].
Available: http://arxiv.org/abs/1412.6115
[134] A. Mishra, J. J. Cook, E. Nurvitadhi, and D.
Marr, ‘WRPN: Training and Inference using
Wide Reduced-Precision Networks’,
ArXiv170403079 Cs, Apr. 2017, Accessed: Sep.
18, 2021. [Online]. Available:
http://arxiv.org/abs/1704.03079
[135] ‘ISO - ISO/IEC JTC 1/SC 42 - Artificial
intelligence’.
https://www.iso.org/committee/6794475/x/catal
ogue/p/1/u/1/w/0/d/0 (accessed Sep. 04, 2021).
[136] J. Törnblom and S. Nadjm-Tehrani, ‘Formal
Verification of Random Forests in
Safety-Critical Applications’, in Formal
Techniques for Safety-Critical Systems, Cham,
2019, pp. 55–71. doi:
10.1007/978-3-030-12988-0_4.
[137] W. Wahlster and Cristoph Winterhalter,
‘GERMAN STANDARDIZATION ROADMAP
ON ARTIFICIAL INTELLIGENCE’, p. 226,
IAC-21-B.1.4.4 Page 29 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
Nov. 2020.
[138] ‘White Paper on Artificial Intelligence: a
European approach to excellence and trust’,
European Commission - European Commission.
https://ec.europa.eu/info/publications/white-pap
er-artificial-intelligence-european-approach-exc
ellence-and-trust_en (accessed Sep. 11, 2021).
[139] Faisal D’Souza, ‘The National Artificial
Intelligence Research and Development
Strategic Plan: 2019 Update’, p. 50, 2019.
[140] Gregory Flandin, ‘White Paper Machine
Learning in Certified System’, p. 113, Feb.
2021.
[141] C. Molnar, Interpretable Machine Learning.
2021. Accessed: Sep. 28, 2021. [Online].
Available:
https://christophm.github.io/interpretable-ml-bo
ok/
[142] P. Linardatos, V. Papastefanopoulos, and S.
Kotsiantis, ‘Explainable AI: A Review of
Machine Learning Interpretability Methods’,
Entropy, vol. 23, no. 1, Art. no. 1, Jan. 2021,
doi: 10.3390/e23010018.
[143] E. Tjoa and C. Guan, ‘A Survey on Explainable
Artificial Intelligence (XAI): Towards Medical
XAI’, IEEE Trans. Neural Netw. Learn. Syst.,
pp. 1–21, 2020, doi:
10.1109/TNNLS.2020.3027314.
[144] ‘AC 25.1309-1A - System Design and Analysis
– Document Information’, Jun. 21, 1988.
https://www.faa.gov/regulations_policies/adviso
ry_circulars/index.cfm/go/document.informatio
n/documentid/22680 (accessed Sep. 11, 2021).
[145] C.-H. Cheng and R. Yan, ‘Testing Autonomous
Systems with Believed Equivalence
Refinement’, ArXiv210304578 Cs, Aug. 2021,
Accessed: Sep. 18, 2021. [Online]. Available:
http://arxiv.org/abs/2103.04578
[146] G. Katz, C. Barrett, D. Dill, K. Julian, and M.
Kochenderfer, ‘Reluplex: An Efficient SMT
Solver for Verifying Deep Neural Networks’,
ArXiv170201135 Cs, May 2017, Accessed: Sep.
04, 2021. [Online]. Available:
http://arxiv.org/abs/1702.01135
[147] S. Wang, K. Pei, J. Whitehouse, J. Yang, and S.
Jana, ‘Formal Security Analysis of Neural
Networks using Symbolic Intervals’,
ArXiv180410829 Cs, Jul. 2018, Accessed: Sep.
04, 2021. [Online]. Available:
http://arxiv.org/abs/1804.10829
[148] M. Mirman, T. Gehr, and M. Vechev,
‘Differentiable Abstract Interpretation for
Provably Robust Neural Networks’, in
International Conference on Machine Learning,
Jul. 2018, pp. 3578–3586. Accessed: Sep. 04,
2021. [Online]. Available:
https://proceedings.mlr.press/v80/mirman18b.ht
ml
[149] M. N. Müller, G. Makarchuk, G. Singh, M.
Püschel, and M. Vechev, ‘PRIMA: Precise and
General Neural Network Certification via
Multi-Neuron Convex Relaxations’, p. 20, 2021.
[150] I. J. Goodfellow, J. Shlens, and C. Szegedy,
‘Explaining and Harnessing Adversarial
Examples’, ArXiv14126572 Cs Stat, Mar. 2015,
Accessed: Sep. 04, 2021. [Online]. Available:
http://arxiv.org/abs/1412.6572
[151] H. Chen, H. Zhang, D. Boning, and C.-J. Hsieh,
‘Robust Decision Trees Against Adversarial
Examples’, ArXiv190210660 Cs Stat, Jun. 2019,
Accessed: Sep. 04, 2021. [Online]. Available:
http://arxiv.org/abs/1902.10660
[152] S. Shalev-Shwartz, Understanding Machine
Learning: From Theory to Algorithms, 1st
edition. New York, NY, USA: Cambridge
University Press, 2014.
[153] Z. Yang, L. Li, X. Xu, B. Kailkhura, T. Xie, and
B. Li, ‘On the Certified Robustness for
Ensemble Models and Beyond’,
ArXiv210710873 Cs, Jul. 2021, Accessed: Sep.
29, 2021. [Online]. Available:
http://arxiv.org/abs/2107.10873
[154] T. Pang, K. Xu, C. Du, N. Chen, and J. Zhu,
‘Improving Adversarial Robustness via
Promoting Ensemble Diversity’, in Proceedings
of the 36th International Conference on
Machine Learning, May 2019, pp. 4970–4979.
Accessed: Sep. 29, 2021. [Online]. Available:
https://proceedings.mlr.press/v97/pang19a.html
[155] K. Shweta, ‘A Survey on Classification of
Concept Drift with Stream Data’, Mar. 2019.
Accessed: Sep. 11, 2021. [Online]. Available:
https://hal.archives-ouvertes.fr/hal-02062610
[156] C. V. Nguyen, A. Achille, M. Lam, T. Hassner,
V. Mahadevan, and S. Soatto, ‘Toward
Understanding Catastrophic Forgetting in
Continual Learning’, ArXiv190801091 Cs Stat,
Aug. 2019, Accessed: Sep. 26, 2021. [Online].
Available: http://arxiv.org/abs/1908.01091
[157] S. Zhao X. Yue, S. Zhang, B. Li, H. Zhao, B.
Wu, R. Krishna, J. E. Gonzalez, A. L.
Sangiovanni-Vincentelli, Sanjit A. Seshia, and
K. Keutzer, ‘A Review of Single-Source Deep
Unsupervised Visual Domain Adaptation’,
ArXiv200900155 Cs Eess, Sep. 2020, Accessed:
Sep. 04, 2021. [Online]. Available:
http://arxiv.org/abs/2009.00155
[158] C. Cappi, C. Chapdelaine, L. Gardes, E. Jenn,
B. Lefevre, S. Picard, and T. Soumarmon,
‘Dataset Definition Standard (DDS)’,
ArXiv210103020 Cs, Jan. 2021, Accessed: Sep.
19, 2021. [Online]. Available:
IAC-21-B.1.4.4 Page 30 of 31
72nd International Astronautical Congress (IAC), Dubai, United Arab Emirates, 25-29 October 2021.
Copyright © [IAC 2021] by Mr. Pablo Miralles. Published by the IAF, with permission and released to the IAF to publish in all forms.
http://arxiv.org/abs/2101.03020
[159] L. R. Davis, K. A. Fishburn, H. Lestinsky, L. R.
Moore, and J. L. Walter, ‘US Topo Product
Standard’, Techniques and Methods, 2019.
[Online]. Available:
https://pubs.usgs.gov/tm/tm11b2/
[160] L. Di, ‘Standards, Critical Evaluation of Remote
Sensing’, in Encyclopedia of GIS, S. Shekhar
and H. Xiong, Eds. Boston, MA: Springer US,
2008, pp. 1128–1135. doi:
10.1007/978-0-387-35973-1_1346.
[161] L. Di and B. Kobler, ‘NASA Standards for
Earth Remote Sensing Data’, Jan. 2000.
https://www.researchgate.net/publication/22895
3572_NASA_Standards_for_Earth_Remote_Se
nsing_Data (accessed Sep. 04, 2021).
[162] ‘Home’, Mindseed. https://www.mindseed.ie/
(accessed Sep. 18, 2021).
[163] ‘SatSure’. https://satsure.co/ (accessed Sep. 18,
2021).
[164] ‘cnes | Le site du Centre national d’études
spatiales’. https://cnes.fr/fr/ (accessed Sep. 18,
2021).
IAC-21-B.1.4.4 Page 31 of 31