scispace - formally typeset
Open AccessJournal ArticleDOI

Quantum-chemical insights from deep tensor neural networks.

TLDR
In this article, a deep tensor neural network is used to predict atomic energies and local chemical potentials in molecules, reliable isomer energies, and molecules with peculiar electronic structure.
Abstract
Learning from data has led to paradigm shifts in a multitude of disciplines, including web, text and image search, speech recognition, as well as bioinformatics. Can machine learning enable similar breakthroughs in understanding quantum many-body systems? Here we develop an efficient deep learning approach that enables spatially and chemically resolved insights into quantum-mechanical observables of molecular systems. We unify concepts from many-body Hamiltonians with purpose-designed deep tensor neural networks, which leads to size-extensive and uniformly accurate (1 kcal mol−1) predictions in compositional and configurational chemical space for molecules of intermediate size. As an example of chemical relevance, the model reveals a classification of aromatic rings with respect to their stability. Further applications of our model for predicting atomic energies and local chemical potentials in molecules, reliable isomer energies, and molecules with peculiar electronic structure demonstrate the potential of machine learning for revealing insights into complex quantum-chemical systems. Machine learning is an increasingly popular approach to analyse data and make predictions. Here the authors develop a ‘deep learning’ framework for quantitative predictions and qualitative understanding of quantum-mechanical observables of chemical systems, beyond properties trivially contained in the training data.

read more

Content maybe subject to copyright    Report

ARTICLE
Received 24 Jun 2016 | Accepted 9 Nov 2016 | Published 9 Jan 2017
Quantum-chemical insights from deep
tensor neural networks
Kristof T. Schu¨tt
1
, Farhad Arbabzadah
1
, Stefan Chmiela
1
, Klaus R. Mu¨ller
1,2
& Alexandre Tkatchenko
3,4
Learning from data has led to paradigm shifts in a multitude of disciplines, including web, text
and image search, speech recognition, as well as bioinformatics. Can machine learning enable
similar breakthroughs in understanding quantum many-body systems? Here we develop an
efficient deep learning approach that enables spatially and chemically resolved insights into
quantum-mechanical observables of molecular systems. We unify concepts from many-body
Hamiltonians with purpose-designed deep tensor neural networks, which leads to size-
extensive and uniformly accurate (1 kcal mol
1
) predictions in compositional and config-
urational chemical space for molecules of intermediate size. As an example of chemical
relevance, the model reveals a classification of aromatic rings with respect to their stability.
Further applications of our model for predicting atomic energies and local chemical potentials
in molecules, reliable isomer energies, and molecules with peculiar electronic structure
demonstrate the potential of machine learning for revealing insights into complex quantum-
chemical systems.
DOI: 10.1038/ncomms13890
OPEN
1
Machine Learning Group, Technische Universita
¨
t Berlin, Marchstr. 23, 10587 Berlin, Germany.
2
Department of Brain and Cognitive Engineering, Korea
University, Anam-dong, Seongbuk-gu, Seoul 136-713, Republic of Korea.
3
Theory Department, Fritz-Haber-Institut der Max-Planck-Gesellschaft, Faradayweg
4-6, D-14195 Berlin, Germany.
4
Physics and Materials Science Research Unit, University of Luxembourg, Luxembourg,, L-1511 Luxembourg. Correspondence
and requests for materials should be addressed to K.R.M. (email: klaus-robert.mueller@tu-berlin.de) or to A.T. (email: alexandre.tkatchenko@uni.lu).
NATURE COMMUNICATIONS | 8:13890 | DOI: 10.1038/ncomms13890 | www.nature.com/naturecommunications 1

C
hemistry permeates all aspects of our life, from the
development of new drugs to the food that we consume
and materials we use on a daily basis. Chemists rely
on empirical observations based on creative and painstaking
experimentation that leads to eventual discoveries of molecules
and materials with desired properties and mechanisms to
synthesize them. Many discoveries in chemistry can be guided
by searching large databases of experimental or computational
molecular structures and properties by using concepts based
on chemical similarity. Because the structure and properties
of molecules are determined by the laws of quantum mechanics,
ultimately chemical discovery must be based on fundamental
quantum principles. Indeed, electronic structure calculations
and intelligent data analysis (machine learning) have recently
been combined aiming towards the goal of accelerated discovery
of chemicals with desired properties
1–8
. However, so far the
majority of these pioneering efforts have focused on the
construction of reduced models trained on large data sets of
density-functional theory calculations.
In this work, we develop an efficient deep learning approach
that enables spatially and chemically resolved insights
into quantum-mechanical properties of molecular systems
beyond those trivially contained in the training dataset.
Obviously, computational models are not predictive if they lack
accuracy. In addition to being interpretable, size-extensive
and efficient, our deep tensor neural network (DTNN) approach
is uniformly accurate (1 kcal mol
1
) throughout compositional
and configurational chemical space. On the more fundamental
side, the mathematical construction of the DTNN model provides
statistically rigorous partitioning of extensive molecular proper-
ties into atomic contributions—a long-standing challenge
for quantum-mechanical calculations of molecules.
Results
Molecular deep tensor neural networks. It is common to use a
carefully chosen representation of the problem at hand as a basis
for machine learning
9–11
. For example, molecules can be
represented as Coulomb matrices
7,12,13
, scattering transforms
14
,
bags of bonds
15
, smooth overlap of atomic positions
16,17
or generalized symmetry functions
18,19
. Kernel-based learning
of molecular properties transforms these representations
non-linearly by virtue of kernel functions. In contrast, deep
neural networks
20
are able to infer the underlying regularities and
learn an efficient representation in a layer-wise fashion
21
.
Molecular properties are governed by the laws of quantum
mechanics, which yield the remarkable flexibility of chemical
systems, but also impose constraints on the behaviour of bonding
in molecules. The approach presented here utilizes the many-
body Hamiltonian concept for the construction of the DTNN
architecture (Fig. 1), embracing the principles of quantum
chemistry, while maintaining the full flexibility of a complex
data-driven learning machine.
DTNN receives molecular structures through a vector of
nuclear charges Z and a matrix of atomic distances D ensuring
rotational and translational invariance by construction (Fig. 1a).
The distances are expanded in a Gaussian basis, yielding a feature
vector
^
d
ij
2 R
G
, which accounts for the different nature of
interactions at various distance regimes. Similar approaches have
been applied to the entries of the Coulomb matrix for the
prediction of molecular properties before
12
.
The total energy E
M
for the molecule M composed of N atoms
is written as a sum over N atomic energy contributions E
i
, thus
satisfying permutational invariance with respect to atom index-
ing. Each atom i is represented by a coefficient vector c 2 R
B
,
where B is the number of basis functions, or features. Motivated
by quantum-chemical atomic basis set expansions, we assign
an atom type-specific descriptor vector c
Z
i
to these coefficients
c
0ðÞ
i
. Subsequently, this atomic expansion is repeatedly refined by
pairwise interactions with the surrounding atoms
c
t þ1ðÞ
i
¼c
tðÞ
i
þ
X
j i
v
ij
; ð1Þ
where the interaction term v
ij
reflects the influence of atom j at a
distance D
ij
on atom i. Note that this refinement step is seamlessly
integrated into the architecture of the molecular DTNN, and
is therefore adapted throughout the learning process. In
Supplementary Discussion, we show the relation to convolutional
neural networks that have been applied to images, speech and text
with great success because of their ability to capture local
structure
22–27
. Considering a molecule as a graph, T refinements
of the coefficient vectors are comprised of all walks of length
T through the molecule ending at the corresponding atom
28,29
.
From the point of view of many-body interatomic interactions,
subsequent refinement steps t correlate atomic neighbourhoods
with increasing complexity.
While the initial atomic representations only consider isolated
atoms, the interaction terms characterize how the basis functions
of two atoms overlap with each other at a certain distance. Each
refinement step is supposed to reduce these overlaps, thereby
embedding the atoms of the molecule into their chemical
environment. Following this procedure, the DTNN implicitly
learns an atom-centered basis that is unique and efficient with
respect to the property to be predicted.
Non-linear coupling between the atomic vector features and
the interatomic distances is achieved by a tensor layer
30–32
, such
that the coefficient k of the refinement is given by
v
ijk
¼ tanh c
tðÞ
j
V
k
^
d
ij
þ W
c
c
tðÞ
j

k
þ W
d
^
d
ij

k
þb
k

; ð2Þ
where b
k
is the bias of feature k and W
c
and W
d
are the weights of
atom representation and distance, respectively. The slice V
k
of the
parameter tensor V 2 R
BBG
combines the inputs
multiplicatively. Since V incorporates many parameters, using
this kind of layer is both computationally expensive as well as
prone to overfitting. Therefore, we employ a low-rank tensor
factorization, as described in (ref. 33), such that
v
ij
¼ tanh W
fc
W
cf
c
j
þb
f
1

W
df
^
d
ij
þb
f
2
hi
; ð3Þ
where represents element-wise multiplication, while W
cf
, b
f
1
,
W
df
, b
f
2
and W
fc
are the weight matrices and corresponding
biases of atom representations, distances and resulting factors,
respectively. As the dimensionality of W
cf
c
j
and W
df
^
d
ij
corresponds to the number of factors, choosing only a few
drastically decreases the number of parameters, thus solving both
issues of the tensor layer at once.
Arriving at the final embedding after a given number of
interaction refinements, two fully-connected layers predict an
energy contribution from each atomic coefficient vector, such that
their sum corresponds to the total molecular energy E
M
.
Therefore, the DTNN architecture scales with the number of
atoms in a molecule, fully capturing the extensive nature of the
energy. All weights, biases, as well as the atom type-specific
descriptors were initialized randomly and trained using stochastic
gradient descent.
Learning molecular energies. To demonstrate the versatility of
the proposed DTNN, we train models with up to three interaction
passes T ¼3 for both compositional and configurational degrees
of freedom in molecular systems. The DTNN accuracy saturates
at T ¼3, and leads to a strong correlation between atoms in
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms13890
2 NATURE COMMUNICATIONS | 8:13890 | DOI: 10.1038/ncomms13890 | www.nature.com/naturecommunications

molecules, as can be visualized by the complexity of the potential
learned by the network (Fig. 1e). For training, we employ
chemically diverse data sets of equilibrium molecular structures,
as well as molecular dynamics (MD) trajectories for small
molecules. We employ two subsets of the GDB-13 database
34,35
referred to as GDB-7, including 47,000 molecules with up to
seven heavy (C, N, O, F) atoms, and GDB-9, consisting of 133,885
molecules with up to nine heavy atoms
36
. In both cases, the
learning task is to predict the molecular total energy calculated
with density-functional theory (DFT). All GDB molecules are
stable and synthetically accessible according to organic chemistry
rules
35
. Molecular features such as functional groups or
signatures include single, double and triple bonds; (hetero-)
cycles, carboxy, cyanide, amide, amine, alcohol, epoxy, sulphide,
ether, ester, chloride, aliphatic and aromatic groups. For each of
the many possible stoichiometries, many constitutional isomers
are considered, each being represented only by a low-energy
conformational isomer.
As Supplementary Table 1 demonstrates, DTNN achieves a
mean absolute error of 1.0 kcal mol
1
on both GDB data sets,
training on 5.8 k GDB-7 (80%) and 25 k (20%) GDB-9 reference
calculations, respectively. Figure 1c shows the performance
on GDB-9 depending on the size of the molecule. We observe
that larger molecules have lower errors because of their
25
# atoms
0.0
0.5
1.0
1.5
2.0
2.5
3.0
Mean abs. error (kcal mol
−1
)
5,000
1.4
1.6
1.8
2.0
2.2
Mean abs. error
0 100 200
Time step
0
−10
−20
−30
−40
Total energy (kcal mol
−1
)
−1.702e5
O
OH
OH
CH
3
O
O
Z
1
D
12
D
13
D
1n
Z
2
D
21
D
23
D
2n
Feedback loop
c
2
(0)
c
1
(0)
Z =
D =
Z
1
D
11
D
21
Z
2
D
12
D
12
D
n2
D
n1
Z
n
D
1n
D
2n
D
nn
v
12
v
13
v
1n
v
21
v
23
v
2n
t < T
t :=t + 1
Interaction module
c
2
(t)
t =T
v
ij
c
1
(t )
t =T
E
1
E
2
t :=t + 1
Gaussian expansion
Hyperbolic tangent
Element-wise product
Element-wise sum
E
n
CH
4
C
6
H
6
C
6
C
3
H
6
C
6
H
5
CH
3
C
3
H
8
C
4
N
2
H
4
E
Molecules with 20 atoms
+
+
+
t < T
2,500
# add. calcs. 15 atoms
20
15
10
a
b
c
d
e
c
j
(t)
tanh
W
df
d
ˆ
ij
+ b
f
2
W
fc
W
cf
c
j
+ b
f
1
Figure 1 | Prediction and explanation of molecular energies with a deep tensor neural network. (a) Molecules are encoded as input for the neural
network by a vector of nuclear charges and an inter-atomic distance matrix. This description is complete and invariant to rotation and translation.
(b) Illustration of the network architecture. Each atom type corresponds to a vector of coefficients c
0ðÞ
i
, which is repeatedly refined by interactions v
ij
.
The interactions depend on the current representation c
t
ðÞ
j
, as well as the distance D
ij
to an atom j. After T iterations, an energy contribution E
i
is predicted
for the final coefficient vector c
TðÞ
i
. The molecular energy E is the sum over these atomic contributions. (c) Mean absolute errors of predictions for the
GDB-9 dataset of 133,885 molecules as a function of the number of atoms. The employed neural network uses two interaction passes (T ¼2) and 50,000
reference calculation during training. The inset shows the error of an equivalent network trained on 5,000 GDB-9 molecules with 20 or more atoms, as
small molecules with 15 or less atoms are added to the training set. (d) Extract from the calculated (black) and predicted (orange) molecular dynamics
trajectory of toluene. The curve on the right shows the agreement of the predicted and calculated energy distributions. (e) Energy contribution E
probe
(or local chemical potential O
M
H
r
ðÞ
, see text) of a hydrogen test charge on a
P
i
r r
i
kk
2
isosurface for various molecules from the GDB-9 dataset for a
DTNN model with T ¼2.
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms13890 ARTICLE
NATURE COMMUNICATIONS | 8:13890 | DOI: 10.1038/ncomms13890 | www.nature.com/naturecommunications 3

abundance in the training data. However, when predicting larger
molecules than present in the training set, the errors increase.
This is because the molecules in the GDB-9 set are quite small,
so we considered all atoms to be in each other’s chemical
environment. Imposing a distance cutoff to interatomic interac-
tions of 3 Å leads to a 0.1 kcal mol
1
increase in the error.
However, this distance cutoff restricts only the direct interactions
considered in the refinement steps. With multiple refinements,
the effective cutoff increases by a factor of T because of indirect
interactions over multiple atoms. Given large enough molecules,
so that a reasonable distance cutoff can be chosen, scaling
to larger molecules will require only to have well-represented
local environments. For now, we observe that at least a few larger
molecules are needed to achieve a good prediction accuracy.
Following this train of thought, we trained the network on a
restricted subset of 5 k molecules with 420 atoms. By adding
smaller molecules to the training set, we are able to reduce the
test error from 2.1 kcal mol
1
to o1.5 kcal mol
1
(see inset in
Fig. 1c). This result demonstrates that our model is able to
transfer knowledge learned from small molecules to larger
molecules with diverse functional groups.
While only encompassing conformations of a single molecule,
reproducing MD simulation trajectories poses a radically different
challenge to predicting energies of purely equilibrium structures.
We learned potential energies for MD trajectories of benzene,
toluene, malonaldehyde and salicylic acid, carried out at a rather
high temperature of 500 K to achieve exhaustive exploration of
the potential-energy surface of such small molecules. The neural
network yields mean absolute errors of 0.05, 0.18, 0.17 and
0.39 kcal mol
1
for these molecules, respectively (Supplementary
Table 1). Figure 1d shows the excellent agreement between
the DFT and DTNN MD trajectory of toluene, as well as the
corresponding energy distributions. The DTNN errors are much
smaller than the energy of thermal fluctuations at room
temperature (B0.6 kcal mol
1
), meaning that DTNN potential-
energy surfaces can be utilized to calculate accurate molecular
thermodynamic properties by virtue of Monte Carlo simulations.
Supplementary Figs 1 and 2 illustrate how the performance
of DTNN depends on the number of employed reference
calculations and refinement steps (Supplementary Discussion).
The ability of DTNN to accurately describe equilibrium structures
within the GDB-9 database and MD trajectories of selected
molecules of chemical relevance demonstrates the feasibility of
developing a universal machine learning architecture that
can capture compositional as well as configurational degrees of
freedom in the vast chemical space. While the employed
architecture of the DTNN is universal, the learned coefficients
are different for GDB-9 and MD trajectories of single molecules.
Local chemical potential. Beyond predicting accurate
energies, the true power of DTNN lies in its ability to provide
novel quantum-chemical insights. In the context of DTNN, we
define a local chemical potential O
M
A
rðÞas an energy of a certain
atom type A, located at a position r in the molecule M. While the
DTNN models the interatomic interactions, we only allow the
atoms of the molecule act on the probe atom, while the probe
does not influence the molecule. The spatial and chemical
sensitivity provided by our DTNN approach is shown in Fig. 1e
for a variety of fundamental molecular building blocks. In this
case, we employed hydrogen as a test charge, while the results for
O
M
C;N;O
rðÞare shown in Fig. 2. Despite being trained only on total
energies of molecules, the DTNN approach clearly grasps
fundamental chemical concepts such as bond saturation and
different degrees of aromaticity. For example, the DTNN model
predicts the C
6
O
3
H
6
molecule to be ‘more aromatic’ than benzene
or toluene (Fig. 1e). Remarkably, it turns out that C
6
O
3
H
6
does
have higher ring stability than both benzene and toluene and
DTNN predicts it to be the molecule with the most stable
aromatic carbon ring among all molecules in the GDB-9 database
(Fig. 3). Further chemical effects learned by the DTNN model are
shown in Fig. 2 that demonstrates the differences in the chemical
potential distribution of H, C, N and O atoms in benzene,
toluene, salicylic acid and malonaldehyde. For example, the
chemical potentials of different atoms over an aromatic ring are
qualitatively different for H, C, N and O atoms—an evident fact
for a trained chemist. However, the subtle chemical differences
described by DTNN are accompanied by chemically accurate
predictions—a challenging task for humans.
Because DTNN provides atomic energies by construction, it
allows us to classify molecules by the stability of different building
blocks, for example aromatic rings or methyl groups. An example
of such classification is shown in Fig. 3, where we plot the
molecules with most stable and least stable carbon aromatic rings
in GDB-9. The distribution of atomic energies is shown in
Supplementary Fig. 3, while Supplementary Fig. 4 lists the full
stability ranking. The DTNN classification leads to interesting
stability trends, notwithstanding the intrinsic non-uniqueness of
atomic energy partitioning. However, unlike atomic projections
employed in electronic-structure calculations, the DTNN
approach has a firm foundation in statistical learning theory.
In quantum-chemical calculations, every molecule would corre-
spond to a different partitioning depending on its self-consistent
electron density. In contrast, the DTNN approach learns the
partitioning on a large molecular dataset, generating a transfer-
able and global ‘dressed atom’ representation of molecules in
chemical space. Recalling that DTNN exhibits errors below
1 kcal mol
1
, the classification shown in Fig. 3 can provide useful
guidance for the chemical discovery of molecules with desired
properties. Analytical gradients of the DTNN model with respect
to chemical composition or O
M
A
rðÞ could also aid in the
exploration of chemical compound space
37
.
Energy predictions for isomers. The quantitative accuracy
achieved by DTNN and its size extensivity paves the way to
the calculation of configurational and conformational energy
differences—a long-standing challenge for machine learning
approaches
7,12,13,38
. The reliability of DTNN for isomer energy
predictions is demonstrated by the energy distribution in Fig. 4
for molecular isomers with C
7
O
2
H
10
chemical formula (a total of
6,095 isomers in the GDB-9 data set).
Training a common model for chemical as well as conforma-
tional freedoms requires a more complex model. Furthermore,
it comes with technical challenges like sampling and multiscale
issues since the MD trajectories form clusters of small variation
within the chemical compound space. As a proof of principle, we
trained the DTNN to predict various MD trajectories of the
C
7
O
2
H
10
isomers. To this end, we calculated short MD
trajectories of 5,000 steps each for 113 randomly picked isomers
as well as consistent total energies for all equilbrium structures.
The training set is composed of all isomers in equilibrium as well
as 50% of each MD trajectory. The remaining MD calculations
are used for validation and testing. Despite the added complexity,
our model achieves a mean absolute error of 1.7 kcal mol
1
.
Discussion
DTNNs provide an efficient way to represent chemical environ-
ments allowing for chemically accurate predictions. To this end,
an implicit, atom-centered basis is learned from reference
calculations. Employing this representation, atoms can be
embedded in their chemical environment within a few refinement
ARTICLE NATURE COMMUNICATIONS | DOI: 10.1038/ncomms13890
4 NATURE COMMUNICATIONS | 8:13890 | DOI: 10.1038/ncomms13890 | www.nature.com/naturecommunications

steps. Furthermore, DTNNs have the advantage that the
embedding is built recursively from pairwise distances. Therefore,
all necessary invariances (translation, rotation, permutation) are
guaranteed to be exploited by the model. In addition, the learned
embedding can be used to generate alchemical reaction paths
(Supplementary Fig. 5).
In previous approaches, potential-energy surfaces were
constructed by fitting many-body expansions with neural
networks
39–41
. However, these methods require a separate NN
for each non-equivalent many-body term in the expansion. Since
DTNN learns a common basis in which the atom interact, higher-
order interactions can obtained more efficiently without separate
treament.
Approaches like smooth overlap of atomic positions
16,17
or manually crafted atom-centered symmetry functions
18,19,42
are, like DTNN, based on representing chemical environments.
All these approaches have in common that size-extensivity
regarding the number of atoms is achieved by predicting atomic
energy contributions using a non-linear regression method
(for example, neural networks or kernel ridge regression).
However, the previous approaches have a fixed set of basis
functions describing the atomic environments. In contrast,
DTNNs are able to adapt to the problem at hand in a
−110
−80
−50
−150 −115 −80
−140 −100 −60
−145 −105 −65
Oxygen
NitrogenCarbonHydrogen
Ω
M
A
(r) in kcal mol
–1
Figure 2 | Chemical potentials O
M
A
rðÞfor A ¼{C, N, O, H} atoms. The isosurface was generated for
P
i
r r
i
kk
2
¼3.8 Å
2
(the index i is used to sum
over all atoms of the corresponding molecule). The molecules shown are (in order from top to bottom of the figure): benzene, toluene, salicylic acid and
malondehyde. Atom colouring: carbon ¼black, hydrogen ¼white, oxygen ¼red.
–859.9 –858.3 –857.8
–857.3 –856.9 –856.8
–845.1 –841.9
–841.7 –841.2
# 1 – 10
# 281 – 290
E
ring
in kcal mol
–1
E
ring
in kcal mol
–1
–857.4–857.4
–856.6–856.8
–841.7 –841.4 –841.1
–841.9–842.1–843.8
Figure 3 | Classification of molecular carbon ring stability. Shown are
20 molecules (10 most stable and 10 least stable) with respect to the
energy of the carbon ring predicted by the DTNN model. Atom colouring:
carbon ¼black; hydrogen ¼white; oxygen ¼red; nitrogen ¼blue;
fluorine ¼yellow.
Kendall rank correlation
coefficient = 0.969
–1,750
–1,800
–1,850
–1,900
–1.900 –1,850 –1,800
–1,750
Atomization energy (DFT)
Atomization energy (NN)
–1,900 –1,850 –1,800 –1,750
Atomization energy (kcal mol
–1
)
Figure 4 | Isomer energies with chemical formula C
7
O
2
H
10
. DTNN
trained on the GDB-9 database is able to acurately discriminate between
6,095 different isomers of C
7
O
2
H
10
, which exhibit a non-trivial spectrum of
relative energies.
NATURE COMMUNICATIONS | DOI: 10.1038/ncomms13890 ARTICLE
NATURE COMMUNICATIONS | 8:13890 | DOI: 10.1038/ncomms13890 | www.nature.com/naturecommunications 5

Citations
More filters
Journal ArticleDOI

A Comprehensive Survey on Graph Neural Networks

TL;DR: This article provides a comprehensive overview of graph neural networks (GNNs) in data mining and machine learning fields and proposes a new taxonomy to divide the state-of-the-art GNNs into four categories, namely, recurrent GNNS, convolutional GNN’s, graph autoencoders, and spatial–temporal Gnns.
Posted Content

Graph Neural Networks: A Review of Methods and Applications

TL;DR: A detailed review over existing graph neural network models is provided, systematically categorize the applications, and four open problems for future research are proposed.
Posted Content

Neural Message Passing for Quantum Chemistry

TL;DR: Using MPNNs, state of the art results on an important molecular property prediction benchmark are demonstrated and it is believed future work should focus on datasets with larger molecules or more accurate ground truth labels.
Journal ArticleDOI

Deep learning and process understanding for data-driven Earth system science

TL;DR: It is argued that contextual cues should be used as part of deep learning to gain further process understanding of Earth system science problems, improving the predictive ability of seasonal forecasting and modelling of long-range spatial connections across multiple timescales.
Journal ArticleDOI

Methods for interpreting and understanding deep neural networks

TL;DR: The second part of the tutorial focuses on the recently proposed layer-wise relevance propagation (LRP) technique, for which the author provides theory, recommendations, and tricks, to make most efficient use of it on real data.
References
More filters
Journal ArticleDOI

Generalized Gradient Approximation Made Simple

TL;DR: A simple derivation of a simple GGA is presented, in which all parameters (other than those in LSD) are fundamental constants, and only general features of the detailed construction underlying the Perdew-Wang 1991 (PW91) GGA are invoked.
Journal ArticleDOI

Density‐functional thermochemistry. III. The role of exact exchange

TL;DR: In this article, a semi-empirical exchange correlation functional with local spin density, gradient, and exact exchange terms was proposed. But this functional performed significantly better than previous functionals with gradient corrections only, and fits experimental atomization energies with an impressively small average absolute deviation of 2.4 kcal/mol.
Journal ArticleDOI

Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density

TL;DR: Numerical calculations on a number of atoms, positive ions, and molecules, of both open- and closed-shell type, show that density-functional formulas for the correlation energy and correlation potential give correlation energies within a few percent.
Proceedings Article

ImageNet Classification with Deep Convolutional Neural Networks

TL;DR: The state-of-the-art performance of CNNs was achieved by Deep Convolutional Neural Networks (DCNNs) as discussed by the authors, which consists of five convolutional layers, some of which are followed by max-pooling layers, and three fully-connected layers with a final 1000-way softmax.
Journal ArticleDOI

Deep learning

TL;DR: Deep learning is making major advances in solving problems that have resisted the best attempts of the artificial intelligence community for many years, and will have many more successes in the near future because it requires very little engineering by hand and can easily take advantage of increases in the amount of available computation and data.
Related Papers (5)