ArticlePDF Available

The Open Quantum Materials Database (OQMD): Assessing the accuracy of DFT formation energies

Authors:

Abstract

The Open Quantum Materials Database (OQMD) is a high-throughput database currently consisting of nearly 300,000 density functional theory (DFT) total energy calculations of compounds from the Inorganic Crystal Structure Database (ICSD) and decorations of commonly occurring crystal structures. To maximise the impact of these data, the entire database is being made available, without restrictions, at www.oqmd.org/download. In this paper, we outline the structure and contents of the database, and then use it to evaluate the accuracy of the calculations therein by comparing DFT predictions with experimental measurements for the stability of all elemental ground-state structures and 1,670 experimental formation energies of compounds. This represents the largest comparison between DFT and experimental formation energies to date. The apparent mean absolute error between experimental measurements and our calculations is 0.096 eV/atom. In order to estimate how much error to attribute to the DFT calculations, we also examine deviation between different experimental measurements themselves where multiple sources are available, and find a surprisingly large mean absolute error of 0.082 eV/atom. Hence, we suggest that a significant fraction of the error between DFT and experimental formation energies may be attributed to experimental uncertainties. Finally, we evaluate the stability of compounds in the OQMD (including compounds obtained from the ICSD as well as hypothetical structures), which allows us to predict the existence of ~3,200 new compounds that have not been experimentally characterised and uncover trends in material discovery, based on historical data available within the ICSD.
ARTICLE OPEN
The Open Quantum Materials Database (OQMD):
assessing the accuracy of DFT formation energies
Scott Kirklin
1
, James E Saal
1
, Bryce Meredig
1
, Alex Thompson
1
, Jeff W Doak
1
, Muratahan Aykol
1
, Stephan Rühl
2
and Chris Wolverton
1
The Open Quantum Materials Database (OQMD) is a high-throughput database currently consisting of nearly 300,000 density
functional theory (DFT) total energy calculations of compounds from the Inorganic Crystal Structure Database (ICSD) and
decorations of commonly occurring crystal structures. To maximise the impact of these data, the entire database is being made
available, without restrictions, at www.oqmd.org/download. In this paper, we outline the structure and contents of the database,
and then use it to evaluate the accuracy of the calculations therein by comparing DFT predictions with experimental measurements
for the stability of all elemental ground-state structures and 1,670 experimental formation energies of compounds. This represents
the largest comparison between DFT and experimental formation energies to date. The apparent mean absolute error between
experimental measurements and our calculations is 0.096 eV/atom. In order to estimate how much error to attribute to the DFT
calculations, we also examine deviation between different experimental measurements themselves where multiple sources are
available, and nd a surprisingly large mean absolute error of 0.082 eV/atom. Hence, we suggest that a signicant fraction of the
error between DFT and experimental formation energies may be attributed to experimental uncertainties. Finally, we evaluate the
stability of compounds in the OQMD (including compounds obtained from the ICSD as well as hypothetical structures), which
allows us to predict the existence of ~ 3,200 new compounds that have not been experimentally characterised and uncover trends
in material discovery, based on historical data available within the ICSD.
npj Computational Materials (2015) 1, 15010; doi:10.1038/npjcompumats.2015.10; published online 11 December 2015
INTRODUCTION
The development of new materials is critical to continued
technological advancement, a fact that has spurred the creation
of the Materials Genome Initiative.
1
One component required to
build material innovation infrastructure and accelerate material
development is the creation of large sets of shared and
comprehensive data. In the 1960s, the development of density
functional theory (DFT)
2,3
created a theoretical framework for
accurately predicting the electronic-scale properties of a crystal-
line solid from rst principles. However, it was many years before
the rst practical DFT algorithms were constructed and calcula-
tions performed,
47
and even then it was an impressive and
noteworthy accomplishment to describe the electronic structure
of a single compound. Since then, computational resources have
advanced to the point where it is now feasible to predict
the properties of many thousands of compounds in an efcient,
high-throughput manner.
816
We extend the promise of high-
throughput DFT to its logical extreme, calculating in a consistent
and accurate manner the properties of a signicant fraction of
known crystalline solids. Using experimentally measured crystal
structures obtained from our partnership with the Inorganic
Crystal Structure Database (ICSD),
17,18
we have created a new
database of DFT-relaxed structures and total energies, which is
called the Open Quantum Materials Database (OQMD). The OQMD
has already been used to perform several high-throughput DFT
analyses for a variety of material applications.
14,1923
There are other efforts in the high-throughput calculation of
compounds from large crystal structure databases, including the
Materials Project,
11,24
the Computational Materials Repository
16
and AFLOWLIB.
25
We intend the current data set to be freely
available, in its entirety, to the scientic community without any
conditions or limitations. It is currently available for download at
www.oqmd.org/download. We envision three important benets
from making the entire data set available. First, the availability of
such a large data set of DFT data enables new and creative uses of
these results by others in the eld who lack the resources to create
their own database. This outcome is strongly in line with the goals
of the Materials Genome Initiative. Second, this data set may serve
as a nucleus from which externaland unafliatedprojects can
grow. By providing calculated electronic structures for a large
fraction of known materials, along with a utility for performing
new calculations, new projects can begin more quickly. Third,
multiple calculations of the same data set (e.g. the ICSD structures)
enable conrmation of the accuracy of the calculations across all
databases. Minor differences in the approach, i.e., the use of a
slightly different set of potentials or a different choice for GGA+U
parameters, makes it possible to see whether a particular choice
gives systematically better results.
This paper is composed of two main sections. First, we provide
the details of the construction of the OQMDa description of the
calculation settings and chemical-potential tting approachand
review the current state of the database, which contains the
DFT-predicted 0-K relaxed ground-state structures and total
energies for every calculable ICSD compound with 34 atoms or
less, a total of over 32,559 compounds. In addition, the database
contains 259,511 hypothetical compounds described in the
1
Department of Materials Science and Engineering, Northwestern University, Evanston, IL, USA and
2
FIZ KarlsruheLeibniz Institute for Information Infrastructure,
Eggenstein-Leopoldshafen, Germany.
Correspondence: C Wolverton (c-wolverton@northwestern.edu)
Received 17 September 2015; revised 31 October 2015; accepted 2 November 2015
www.nature.com/npjcompumats
All rights reserved 2057-3960/15
© 2015 Shanghai Institute of Ceramics, Chinese Academy of Sciences/Macmillan Publishers Limited
section Structures in the OQMD, based on decorations of
commonly occurring crystal structures. The database also contains
a growing number of additional structures (currently ~ 5,000)
calculated for ongoing material discovery projects such as for
structural alloys and energy materials.
14
These numbers give
OQMD a total size of 297,099 DFT calculations to date. Second, we
employ the OQMD to investigate a fundamental question of DFT:
how accurately can we use DFT to reproduce experimentally
known elemental ground states and compound-formation
energies? We use these comparisons as a basis for establishing
condence in the database, before using the database to calculate
the stability of every compound in the database. With this large
library of DFT calculations of hypothetical compounds included in
the OQMD, we are able to make predictions of compositions
where new, previously unknown compounds are likely to exist; in
this work we identify 3,231 such compositions in which one of our
hypothetical structures is predicted to be stable. Finally, we use
the breadth of our database to examine historical trends in
material discovery.
RESULTS AND DISCUSSION
The OQMD Methodology
A critical component of any high-throughput DFT database is the
infrastructure to create and access the contained knowledge,
e.g., pymatgen,
26
ASE
27
and AFLOW.
15
We have developed an
infrastructure for such high-throughput DFT calculations and
database management, dubbed qmpy. qmpy is written in python,
and it uses the django web framework as an interface to a MySQL
database. In the same Openspirit as the database itself, qmpy is
freely available for download as well. It is our goal to develop tools
that any research group can use to catalogue, access and analyse
large sets of calculations. To this end, qmpy is designed with a
decentralised modelany user can download and use it to build a
database (e.g., PostgreSQL, MySQL, sqlite or Oracle), and have
simple, programmatic access to their calculations. The package
has a built-in web interface, and, because we utilise a django
backend, it is very easy to customise the web interface depending
on the specic needs of any user. Details of qmpy and its analysis
algorithms can be found at www.oqmd.org/static/docs.
The calculation of many thousands of compounds within DFT in
a reasonable timeframe demands that optimal efciency within a
constrained standard of convergence be found. Furthermore, the
comparison of calculation results across many different types of
materials (e.g., metals, semiconductors and oxides) requires that
all the calculations be performed at a consistent level of theory,
which is acceptable for all classes of materials, e.g., consistent
plane-wave cutoff, smearing schemes and k-point densities.
To that end, extensive testing on a sample of ICSD structures
has resulted in the calculation ow described in Materials
and Methods, which ensures converged results in an efcient
manner for a variety of material classes. Furthermore, the settings
are consistent across all the calculations, ensuring that
results between different compounds are directly comparable
(e.g., predictions of energetic stability). Using DFT and DFT+U
calculations (with the parameters listed in Table 1) and the
scheme described in the Materials and Methods section, at the
time of the writing of this paper we have calculated 32,559
compounds from the ICSD and 259,511 hypothetical compounds
based on decorations of prototype structures.
Structures in the OQMD
The structures in the OQMD come from two sources. The rst
source, and the origin of the majority of our lowest-energy
structures, is the ICSD. For the ICSD structures, we start with a list
of 148,279 entries. Of those, 64,412 structures contain atomic
positions with partial occupancy, which are substantially more
difcult to treat in a high-throughput manner and thus are not
included in the current study. A further 32,202 are found to be
duplicates of other entries and have been discarded. Structural
uniqueness is determined with respect to the lattice and the
internal coordinates. Two lattices are compared by nding the
reduced primitive cells of each structure and comparing all lattice
parameters. Internal coordinates of different structures are
compared by testing all rotations allowed by each lattice, and
searching for a rotation + translation that maps atoms of the same
species onto one another within a given tolerance. Here, any two
structures in which all atoms can be mapped to within 0.2 Å of an
identical atom are considered identical. Of the remaining
structures, 13,934 have incomplete entries in the database,
missing either atomic coordinates or spacegroup information.
The removal of these structures leaves a pool of 44,506 unique,
eligible structures to be calculated. Having started with the
structures with the fewest number of atoms, the database
currently consists of calculations for all ICSD entries consisting of
less than 34 atoms and passing the above lters, a total of 32,559
structures.
The differences between what we identify as calculable, unique
structures and the total set of structures in the ICSD must be
understood in the context of the challenges of creating a
repository of measured crystal structures. There is no general
way to judge the quality of a crystal structure. Moreover, it is not
simple to determine which of the several measurements/
renements is the best for a given structure, because there are
many ways to manipulate the typical criteria used (r-values,
goodness-of-t and low estimated standard deviation). These
challenges are compounded by the fact that every measured
crystal structure is an average in time (i.e., the time over which
data are collected) and space (size of the crystal). DFT energetics
can help resolve some of these difculties by providing another
criterion for determining the physicalness of a rened structure, as
well as differentiating between structures that give similar ts to
experimental data.
28
In addition to ICSD structures, we have also calculated
decorations of many simple prototype structures over a wide
range of compositions. We dene a prototype structure to be a
crystal structure commonly observed in nature for a variety of
chemical compositions, e.g., A1 FCC and L1
2
. Table 2 gives a
complete listing of the prototype structures that we have
calculated. For every elemental prototype, we have calculated
every element for which Vienna Ab-initio Simulation Package
(VASP) includes projected augmented wave-Perdew, Burke
and Ernzerhof (PAW-PBE) potentials (89 elements). For all of the
binary prototypes, we have calculated each structure with every
combination of elements excluding the noble gases (84 elements).
Table 1. GGA+U UJvalues and their corresponding tted chemical-
potential corrections employed in OQMD calculations for oxides of the
listed elements
Element U J (eV) Correction (eV/atom)
V 3.1 2.675
Cr 3.5 2.818
Mn 3.8 1.987
Fe 4.0 2.200
Co 3.3 1.987
Ni 6.4 2.530
Cu 4.0 1.381
Th 4.0 0.999
U 4.0 2.614
Np 4.0 2.705
Pu 4.0 2.177
Abbreviation: OQMD, Open Quantum Materials Database.
The Open Quantum Materials Database
S Kirklin et al
2
npj Computational Materials (2015) 15010 © 2015 Shanghai Institute of Ceramics, Chinese Academy of Sciences/Macmillan Publishers Limited
This results in 3,486 compounds for structures with symmetrically
equivalent sites and 6,972 compounds for structures with
symmetrically distinct sites (e.g., AB
2
or A
3
B structures). Perovskite
and defect-pervoskite prototypes are only calculated for oxides,
and thus are treated as a binary, e.g., ABO
3
. We also calculated one
ternary prototype, the Heusler or L2
1
structure. We have calculated
186,596 compounds in the Heusler structure. A relatively small
number of additional hypothetical compounds (~5,000 to date)
computed for projects that utilised OQMD are also included in the
database.
We included these prototype structures for two purposes. First,
by including all compositions, we have a more complete picture of
the energetic landscape of phase space. It is important to
understand the energy landscape over a comprehensive range
of compositions and structures, as, in order to reliably assess the
stability of any individual compound, its energy must be
compared with the energy of all possible competing phases and
combinations of phases. These hypothetical compounds based on
common prototype structures are useful because they ensure that
our stability calculations are reasonable, even at compositions
where limited experimental data are available. Furthermore, at
compositions where prototype compounds are predicted to be
stable, it is likely that new compounds are waiting to be
discovered. Although in general the prototype compounds that
are predicted to be stable are not likely to be the true stable
ground-state structures, they indicate a region of composition
space where some new stable compounds must exist, but which
has not been found (or at least not in the OQMD set of ICSD
structures). Thus, these ground states represent predictions of new
ordered compounds that should be validated experimentally.
Second, internal interest in particular crystal structures has
motivated the calculation of some prototypes for specic
applicationsa demand that is easily accomplished using the
qmpy framework.
Elemental Ground-State Prediction
We begin by examining the phase stability of all 89 elements
included in OQMD in a variety of structure types. We determine
the lowest DFT energy structure for every element with a
VASP PAW-PBE potential (89 elements) and compare them with
the experimentally observed low-temperature structures.
29,30
In addition to the elemental structures in the ICSD, we also
Table 2. Prototype structures calculated in the OQMD
Designation Number of atoms SG Formula Description Example
Unary
A1 1 Fm3m AFCC Cu
A2 1 Im3m ABCC W
A3 2 P6
3
/mmc A HCP Mg
A3' 4 P6
3
/mmc A Distorted HCP α-La
A4 2 Fd3m A Diamond C
A5 2 I4
1
/amd A β-Sn
A6 1 I4/mmm A In
A7 2 R3m Aα-As
A8 3 P3
1
21 A γ-Se
A9 4 P6
3
/mmc A Graphite C
A10 1 R3m A Simple rhombohedral α-Hg
A11 4 Cmca A BCO α-Ga
A12 29 I43m Aα-Mn
A13 20 P4
1
32 A β-Mn
A15 8 Pm3m AW
3
Oβ-W
A17 4 Cmca A Black phosphorus P
A20 2 Cmcm A α-U
A
a
1 I4/mmm A BCT α-Pa
A
h
1Pm3m A Simple cubic α-Po
C19 3 R3m Aα-Sm
Binary
B2 2 Pm3m AB BCC superstructure CsCl
D0
3
4Fm3m A
3
B BCC superstructure AlFe
3
L1
0
2 P4/mmm AB FCC superstructure along [100] AuCu
L1
1
2R3m AB FCC superstructure along [111] CuPt
L1
2
4Pm3m A
3
B FCC superstructure with maximised unlike bonds Cu
3
Au
B
h
2 P6m2 AB HCP superstructure along [0001] WC
B19 4 Pmma AB HCP superstructure AuCd
D0
19
8P6
3
/mmc A
3
B HCP superstructure with maximised unlike bonds Ni
3
Sn
E2
1
5Pm3m ABO
3
Perovskite CaTiO
3
defect-E2
1
9 P4/mmm A
2
B
2
O
5
Defect-perovskite
D5
1
10 R3cA
2
B
3
Corundum Al
2
O
3
C4 6 P4
2
/mnm AB
2
Rutile TiO
2
B1 2 Fm3m AB Rocksalt NaCl
B3 2 F43m AB Zincblende ZnS
B4 4 P6
3
mc AB Wurtzite ZnS
D0
22
4 I4/mmm A
3
B FCC superstructure along [012] Al
3
Ti
Ternary
L2
1
4Fm3m A
2
BC Heusler BCC superstructure Cu
2
MnAl
Abbreviation: OQMD, Open Quantum Materials Database.
For each structure the number of atoms in the primitive cell, general formula, a simple description and an example compound are given.
The Open Quantum Materials Database
S Kirklin et al
3
© 2015 Shanghai Institute of Ceramics, Chinese Academy of Sciences/Macmillan Publishers Limited npj Computational Materials (2015) 15010
Table 3. Elemental ground-state structures and chemical potentials predicted by DFT at 0 K
Potential DFT ground state μ
i
(eV/atom) Exp. LT Exp. RT
ID SG Fit-none Fit-partial Fit-all ID SG ΔE(eV/atom) ID SG ΔE(eV/atom)
H I4/mmm 3.327 3.394 3.434 Gas
He A5 I41/amd 0.004 0.004 0.004 A3 P6
3
/mmc 0.006 Gas
LLsv C19 R3m 1.907 1.907 1.731 A2 Im3m 0.003
Be A3 P6
3
/mmc 3.755 3.755 3.653
BR3m 6.678 6.678 6.656
CA9P6
3
/mmc 9.217 9.217 9.044
NPa38.235 8.122 8.195 Gas
O C2/m 4.844 4.485 4.523 Gas
F A11 Cmca 1.666 1.429 1.443 C2/c 0.006 Gas
Ne A1 Fm3m 0.029 0.029 0.029 Gas
Na_pv A3 P6
3
/mmc 1.303 1.212 1.196 A1 Fm3m 0.003
Mg A3 P6
3
/mmc 1.542 1.542 1.417
Al A1 Fm3m 3.746 3.746 3.660
Si A4 Fd3m 5.425 5.425 5.386
PP15.405 5.161 5.175 A17 Cmca 0.031
S P21 4.114 3.839 3.868
Cl Cmca 1.820 1.465 1.479 Gas
Ar A3 P6
3
/mmc 0.006 0.006 0.006 A1 Fm3m 0.009 Gas
K_sv A7 R3m 1.097 1.097 0.987 A2 Im3m 0.000
Ca_pv A1 Fm3m 1.978 1.978 1.780
Sc_sv A3 P6
3
/mmc 6.328 6.328 6.344
Ti P6/mmm 7.776 7.712 7.702 A3 P6
3
/mmc 0.014
VA2Im3m 8.941 8.941 8.898
Cr A2 Im3m 9.508 9.508 9.463
Mn A12 I43m 9.027 9.027 8.898
Fe A2 Im43m 8.308 8.308 8.499
Co A3 P6
3
/mmc 7.090 7.090 7.078
Ni A1 Fm3m 5.567 5.567 5.587
Cu A1 Fm3m 3.716 3.716 3.710
Zn A3 P6
3
/mmc 1.266 1.266 1.157
Ga_d A11 Cmca 3.032 3.032 2.902
Ge_d A4 Fd3m 4.624 4.624 4.522
As A7 R3m 4.652 4.652 4.593
Se A8 P3
1
21 3.481 3.481 3.374
Br Cmca 1.606 1.317 1.333 Liquid
Kr A3 P6
3
/mmc 0.004 0.004 0.004 A1 Fm3m 0.003 Gas
Rb_sv C19 R3m 0.963 0.963 0.881 A2 Im3m 0.001
Sr_sv A1 Fm3m 1.683 1.683 1.549
Y_sv A3 P6
3
/mmc 6.464 6.464 6.449
Zr_sv A3 P6
3
/mmc 8.547 8.547 8.438
Nb_pv A2 Im3m 10.094 10.094 10.017
Mo_pv A2 Im3m 10.848 10.848 10.921
Tc_pv A3 P6
3
/mmc 10.361 10.361 10.457
Ru A3 P6
3
/mmc 9.202 9.202 9.210
Rh A1 Fm3m 7.269 7.269 7.319
Pd A1 Fm3m 5.177 5.177 5.197
Ag A3P6
3
/mmc 2.822 2.822 2.907 A1 Fm3m 0.000
Cd A3 P6
3
/mmc 0.900 0.900 0.861
In_d A6 I4/mmm 2.720 2.720 2.609
Sn_d A4 Fd3m 4.007 3.895 3.938 A5 I41/amd 0.042
Sb A7 R3m 4.118 4.118 4.155
Te A8 P3
1
21 3.142 3.142 3.027
I Cmca 1.509 1.344 1.365
Xe A6 I4/mmm 0.003 0.003 0.640 A1 Fm3m 0.004 Gas
Cs_sv A3 P6
3
/mmc 0.855 0.855 0.744 A2 Im3m 0.002
Ba_sv A2 Im3m 1.924 1.924 1.479
La A3P6
3
/mmc 4.935 4.935 4.959
Ce_3 A3P6
3
/mmc 4.777 4.777 4.564 A1 Fm3m 0.006 A3P6
3
/mmc 0.000
Pr_3 A3P6
3
/mmc 4.775 4.775 4.627
Nd_3 A3P6
3
/mmc 4.763 4.763 4.697
Pm_3 A3P6
3
/mmc 4.745 4.745 4.716
Sm_3 A3P6
3
/mmc 4.715 4.715 4.606 C19 R3m 0.004
Eu_2 A2 Im3m 1.888 1.888 1.708
Gd_3 A3P6
3
/mmc 4.655 4.655 4.718 A3 P6
3
/mmc 0.017
Tb_3 C19 R3m 4.629 4.629 4.731 A20 Cmcm 0.017 A3 P6
3
/mmc 0.012
Dy_3 C19 R3m 4.602 4.602 4.660 A20 Cmcm 0.013 A3 P6
3
/mmc 0.008
Ho_3 C19 R3m 4.577 4.577 4.573 A3 P6
3
/mmc 0.003
Er_3 A3 P6
3
/mmc 4.563 4.563 4.582
The Open Quantum Materials Database
S Kirklin et al
4
npj Computational Materials (2015) 15010 © 2015 Shanghai Institute of Ceramics, Chinese Academy of Sciences/Macmillan Publishers Limited
calculate each element in 20 elemental ground-state structures,
listed in Table 2. For each element, therefore, we have the
energetics of a large number and geometric variety of (20)
crystal structure types, and we refer to the structure with the
lowest energy as the DFT-predicted T= 0 K ground state. Previous
efforts at a comprehensive comparison of experimentally
observed elemental ground-state structures and DFT-predicted
elemental ground-state structures have been limited to HCP, FCC
and BCC structures.
31
Hence, our present study provides a more
complete, systematic investigation of the ability of DFT (at the
settings of the OQMD) to predict the correct 0-K elemental
ground-state crystal structures.
Table 3 shows the DFT-predicted ground-state structure and the
experimentally observed ground-state structure
29,30
for all 89
elements, for which a potential is available in VASP. During DFT
calculations of the 89 elements in the wide spectrum of crystal
structure types tested, it is possible for an element to relax to a
higher-symmetry parent structure starting from a lower-symmetry
prototype-based crystal structure. Therefore, for all elements in
Table 3, we analysed the OQMD-relaxed structures of elements
that are almost degenerate with the ground-state structure (taken
as within ~ 2 meV/atom) using the structure comparison approach
outlined in the section Structures in the OQMDWe found that
only a small set of structures (Al, Ni, Rh, Ir and Th in A6, K, Sr, Al
and Th in A7, Sr in A8 and Pa in A10 structures) relaxed to the
higher-symmetry A1 (FCC) structure, and only Mg and Er in the
A20 structure relaxed to the higher-symmetry A3 (HCP) structure.
Of the 89 elements in Table 3, the OQMD prediction agrees with
the experimentally observed low-temperature structure for 82
elements to within 12 meV/atom, and 77 elements to within
5 meV/atom. For the 12 elements with a discrepancy between the
OQMD-predicted ground state and the experimentally observed
ground state greater than 5 meV/atom, we look for possible
sources of error. The elements for which we fail to correctly predict
the ground-state structure are He, F, P, Ar, Ce, Gd, Tb, Dy, Yb, Hg,
Ac and Pa. There are three distinct groups among these elements:
rst are noble gases and molecular solids, i.e., He, Ar, F and P,
second are solids of elements with f-electrons, i.e., Ce, Gd, Tb, Dy,
Yb, Ac and Pa and third is Hg. For most of these elements, we
found no change in the predicted ground states with calculations
at ~ 30100% higher plane-wave basis-set cutoffs. The exceptions
are He, F and Ar for which the higher-cutoff calculations can
reproduce the experimental ground states. However, for such
noble gas and molecular solids, it is expected that van der Waals
interactions contribute a signicant portion of the total energy
and therefore this change in stabilities cannot be substantiated.
For phosphorous, which has the second largest discrepancy
between the OQMD-predicted ground state and the experimental
ground state, the experimental ground state consists of layers of
covalently bonded atoms that interact primarily through van der
Waals forces, and increasing the cutoff has no effect on the
predicted relative stability of P allotropes. For all elements in the
rst group above, the ground-state structure cannot be deter-
mined reliably with (semi)-local exchange-correlation functionals
used in DFT because of the lack of van der Waals interactions, and
can likely be corrected by their inclusion.
32
The second groupelements that contain f-electrons
presents its own set of known challenges for DFT.
3338
There are
two sources of error for these elements. For Ce, Gd, Tb, Dy and Yb
we use the frozenpotentials, meaning that the f-electrons are
frozen in the core and not treated explicitly as valence. These
potentials may introduce errors because freezing the f-electrons
into the core neglects interactions involving these electrons.
For example, for Ce, we found that the experimental low-
Table 3. (Continued )
Potential DFT ground state μ
i
(eV/atom) Exp. LT Exp. RT
ID SG Fit-none Fit-partial Fit-all ID SG ΔE(eV/atom) ID SG ΔE(eV/atom)
Tm_3 A3 P6
3
/mmc 4.475 4.475 4.451
Yb_2 A1 Fm3m 1.513 1.513 1.125 A3 P6
3
/mmc 0.007 A1 Fm3m 0.000
Lu_3 A3 P6
3
/mmc 4.524 4.524 4.549
Hf_pv A3 P6
3
/mmc 9.955 9.955 9.902
Ta_pv A2 Im3m 11.853 11.853 11.941
W_pv A2 Im3m 12.960 12.960 13.130
Re A3 P6
3
/mmc 12.423 12.423 12.378
Os_pv A3 P6
3
/mmc 11.226 11.226 11.374
Ir A1 Fm3m 8.855 8.855 8.953
Pt A1 Fm3m 6.056 6.056 6.162
Au A1 Fm3m 3.267 3.267 3.283
Hg A12 I43m 0.298 0.376 0.374 A
a
I4/mmm 0.074 Liquid
Tl_d A3 P6
3
/mmc 2.359 2.359 2.480
Pb_d A1 Fm3m 3.704 3.704 3.951
Bi_d A7 R3m 4.039 4.039 4.199
Ac A3P6
3
/mmc 4.106 4.106 4.106 A1 Fm3m 0.012
Th A1 Fm3m 7.413 7.413 7.237
Pa A1 Fm3m 9.496 9.496 9.497 A
a
I4/mmm 0.017
U A20 Cmcm 11.292 11.292 11.032
Np Pnma 12.940 12.940 12.797
Pu P2
1
/m 14.298 14.298 13.950
Abbreviations: DFT, density functional theory; Exp. LT, experimentally observed lowest temperature; Exp. RT, experimentally observed room temperature
PAW-PBE, projected augmented wave-Perdew, Burke and Ernzerhof; VASP, Vienna Ab-initio Simulation Package.
The Potentialcolumn corresponds to VASP PAW-PBE potential names for the elements used in the current work. The ID and SG columns list the
Strukturbericht designation and the space group of the crystal structures, respectively. The t-nonechemical potentials are the DFT ground-state total
energies of each element, whereas the t-partialand t-allcorrespond to the chemical-potential correction schemes described in the section Formation
Energy Calculation.TheExp. LTand/or Exp. RTground states
29,30
are given if they differ from the DFT-predicted 0-K ground state (RT only provided if it
differs from LT), along with the difference between the DFT-predicted total energy of the Exp. LT or RT structure and DFT ground-state energy (energy
difference only provided if the DFT ground state differs from Exp. LT or Exp. RT).
The Open Quantum Materials Database
S Kirklin et al
5
© 2015 Shanghai Institute of Ceramics, Chinese Academy of Sciences/Macmillan Publishers Limited npj Computational Materials (2015) 15010
temperature structure (α-Ce) can be reproduced as the ground
state using the Ce potential with f-electrons in the valence.
However, in the OQMD, frozen potentials are used because the
high degree of correlation of f-electrons and related self-
interaction error when they are included as valence is hard to
accurately treat with DFT. Pa and Ac are examples where frozen
potentials are not available, and therefore the freef-electrons
lead to errors. The issues with f-electrons in DFT, as well as the
trade-offs that come with using frozen potentials, are thoroughly
discussed elsewhere.
3337
Lastly, it is already known that the
relative stabilities of allotropes of Hg cannot be reproduced
with local density approximation or generalized gradient approx-
imation (GGA),
39
and relativistic effects such as spinorbit
coupling (excluded from our DFT calculations in the OQMD) are
essential for accurate treatment of Hg.
40
Formation Energies of Compounds
One of the most useful quantities that we have calculated for each
compound is its formation energythe energy required to form
(positive formation energy), or given off by forming (negative
formation energy), a compound from its constituent elements.
Compound-formation energies are required to predict compound
stability, generate phase diagrams, calculate reaction enthalpies
and voltages and determine many other material properties.
Because this quantity is so ubiquitous, it is important to determine
the trustworthiness of our predictions. Although previous large-
scale investigations of DFTs accuracy in predicting formation
energies have been performed, these investigations have either
been limited in the scope of material type considered
4144
or in
the quantity of structures assessed.
45
The database of formation
energies for 297,099 structures in the OQMD allows for the most
comprehensive assessment of the ability of DFT to predict
formation energies for solids to date.
As DFT calculations are performed at 0 K and experimental
formation energies are typically measured at room temperature,
we must consider how much we can expect the formation energy
to change between 0 and 300 K. The largest source of differences
between 0 and 300 K formation energies is the existence of
phase transformations in this temperature range. These phase
transformations can take the form of solidliquid, solidgas or
solidsolid transformations and lead to signicant changes in
energetics. In addition to the elements that are gaseous or liquid
at room temperature, at least ve elements are known to exhibit a
solidsolid transformation below 300 K: Ce, Na, Li, Ti and Sn.
29
As shown in Table 3, the energy differences between the 0-K-
and room-temperature structures of Ce, Na and Li are less than
7 meV/atom. For Ti and Sn, however, the energy differences are 14
and 42 meV/atom, respectively, which will introduce systematic
errors in the OQMD-predicted formation energies when compar-
ing with experimental formation energies. For this reason, the
chemical potentials of Ti and Sn have been t to experimental
data, as discussed in the section Formation Energy Calculation.
In the following sections we make extensive use of a large
collection of experimental formation energies. These experimental
formation energies come from two sources: the SGTE Solid
SUBstance (SSUB) database,
46
from which we obtain 1,702
compound-formation energies, and the thermodynamic database
at the Thermal Processing Technology Center at the Illinois
Institute of Technology (IIT),
47
from which we obtain 994
compound-formation energies. The SSUB contains many oxides
(680), nitrides (75), hydrides (102) and halides (369), and relatively
few intermetallics (272). In contrast, the IIT database is exclusively
intermetallic compounds. By combining these databases we have
a total of 2,712 experimental formation energies to compare with.
For Th, Pa and Np oxides, which do not appear in either of these
databases, we also include formation energies reported in a
review of actinide thermodynamics,
48
which includes measured
formation energies for ThO
2
,Np
2
O
5
and NpO
2
and estimated
formation energies for PaO
2
and NpO
3
from trends among similar
actinide oxides.
Because the different experimental data sources specialise in
different types of materials, they have relatively limited overlap,
but they do have some compositions in common. As a result, after
controlling for double counting of compositions within and
between databases, we still have 2,233 distinct compositions with
experimental formation energies. The number of comparisons is
further reduced because some of the compositions for which we
have experimental formation energies have no known corre-
sponding crystal structure. One might wish to ascertain whether,
for a given stoichiometry, a given experimental formation energy
Figure 1. Comparison between the OQMD and 1,670 experimentally measured formation energies for three different sets of elemental
chemical potentials. (a) Fit-none reference states: the DFT energy of the OQMD ground-state structure is taken as the chemical potential for
each element. (b) Fit-partial reference states: chemical potentials for 13 elements where the DFT ground-state structures are known to poorly
represent the STP elemental chemical potentials are t to experimental formation energies (for all other elements, the reference state is the
DFT energy). (c) Fit-all reference states: the chemical potentials for all elements are t to experimental formation energies. The solid red line in
each plot corresponds to the average error between DFT and experiment. The dashed red lines indicate the rst and second s.d.s. The curves
in the lower plots correspond to normal distributions computed from the mean and s.d. of each data set.
The Open Quantum Materials Database
S Kirklin et al
6
npj Computational Materials (2015) 15010 © 2015 Shanghai Institute of Ceramics, Chinese Academy of Sciences/Macmillan Publishers Limited
corresponds precisely to the phase with crystal structure reported
in the ICSD. However, this comparison is not possible because the
thermodynamic databases often do not have any explicit crystal
structure information. We therefore assume that the existence of a
corresponding crystal structure in the ICSD is sufcient to include
a composition in the formation energy comparison, and, if more
than one crystal structure is available, we always compare with the
lowest-energy OQMD structure. As a result, we have a total of
1,670 formation energies to compare between the OQMD and
experimentthe largest comparison of this kind that has been
performed to date.
Formation Energy Calculation. In general, the formation energy
for a compound is given by,
ΔHf¼Etot -X
i
μixið1Þ
where E
tot
is the DFT total energy of the compound, μ
i
is the
chemical potential of element iand x
i
is the quantity of element i
in the compound. The standard convention is to take the chemical
potential of each species to be the DFT total energy of the
elemental ground state. With this choice, the computed formation
energy is valid only for 0 K. However, as the available experimental
formation energies are typically measured at room temperature
or above, it is useful to understand how well equation 1
approximates the standard temperature and pressure (STP)
formation energy. In order to answer this question, we calculate
the formation energy of all OQMD-calculated compounds and
compare the results with experimentally measured formation
energies.
Using the elemental DFT total energies as chemical potentials,
we nd that the OQMD-calculated formation energies have an
average error of 0.105 eV/atom with respect to experimentally
measured formation energies and a mean absolute error (MAE) of
0.136 eV/atom. The 1,670 formation energies used in this
comparison include a wide range of compositions, including
oxides, semiconductors and intermetallics. This error is plotted
against experimental formation energy in Figure 1a. There is a
clear systematic error: compounds with more-negative formation
energies underestimate formation energies compared with those
having less-negative formation energies. This trend is easily
understood; highly stable compounds, such as oxides and halides,
frequently contain elements whose 0-K ground states differ from
the STP stable phases, e.g., gases. In order to improve these
formation energies, we t chemical potentials to experimental
formation energies.
The use of experimental data to t elemental reference energies
has been shown to reduce systematic error in DFT formation
energies.
11,44,45,49,50
In this work we perform simultaneous least
squares tting in the manner of the Fitted Elemental Reference
Energies method,
45
which was developed for nding chemical
potentials for GGA+U oxides; however, in our work we extend this
method to a wider set of elements involving a much wider set of
compounds. In our approach we apply two distinct corrections;
therefore, we perform two independent ts. The rst tisto
calculate chemical potentials for elements with ground states that
are not applicable to STP, and the second tistond corrections
to chemical potentials for elements to which we apply GGA+U
(Table 1). In the rst step, we correct all the elemental chemical
potentials simultaneously, tting to the experimental formation
energies (binary, ternary, quaternary and so on) of every
applicable compound that does not contain a GGA+U element,
and in the second step we determine values for the chemical
potentials of GGA+U elements (described below).
To evaluate the efcacy of chemical-potential tting, we dene
three t sets that we will compare throughout this section and the
next. The rst t set is empty, which we will label t-none,and
corresponds to completely uncorrected DFT. The second t set we
will label t-partial,which refers to the elements for which we can
rationally argue that the DFT energy of the T= 0 K ground state is
not an accurate reference state for STP formation energies. We
have identied ve groups of elements to include in this t set.
These groups are room-temperature diatomic gases (H, N, O, F
and Cl), room-temperature liquids (Br and Hg), molecular solids
(P, S and I), several elements with structural phase transformations
between 0 and 298 K (Na, Ti and Sn) and elements employing
GGA+U for oxide-formation energies (Table 1). The third t set,
which we will label t-all,has the chemical potential for every
element t to experiment; none of the chemical potentials come
from DFT.
For the case of GGA+U elements, the chemical potentials of the
elements in structures with the U correction (e.g., Fe in Fe
3
O
4
) and
without the U correction (e.g., Fe in BCC Fe) are different.
11
To
address this, we t corrections for the GGA+U elements (Table 1)
using the method of ref. 11 after all other corrections described in
this section; no compounds for which GGA+U was applied were
used in the tting of chemical potentials for non-GGA+U species.
For GGA+U compounds, all corrected chemical potentials (e.g., O
2
)
Figure 2. Corrections to chemical potentials (μ
t
μ
DFT
) as determined by tting OQMD formation energies to experimental formation
energies. The corrections in blue were obtained by tting only the chemical potentials of elements whose STP phase differs signicantly from
the 0-K phase, while the corrections in red were obtained by tting the chemical potentials of all elements simultaneously.
The Open Quantum Materials Database
S Kirklin et al
7
© 2015 Shanghai Institute of Ceramics, Chinese Academy of Sciences/Macmillan Publishers Limited npj Computational Materials (2015) 15010
were applied rst, and then the GGA+U correction was
determined.
The difference between the tted chemical potential (for both
t-partial and t-all sets) and DFT ground-state energy for each
element is shown in Figure 2, and the values of the chemical
potentials for each element in each t set are provided in Table 3.
The t-partial corrections are consistently larger than the t-all
corrections for the same elements. However, compared with the
magnitude of the corrections, the difference between t-partial
and t-all corrections is quite small in all cases as observed in
Figure 2. Adding more tting parameters in t-all does not
substantially alter the tted chemical potentials from t-partial,
suggesting that the corrections in t-partial capture the majority
of the error associated with those particularly problematic
elemental chemical potentials.
In the t-all set, some elements appear to have surprisingly
large corrections. These are elements that have relatively few
compounds to ttothe actinides and some lanthanides have
comparatively large corrections. The small pool of formation
energies may lead to overcorrection, and less-predictive calcu-
lated formation energies. For this reason, the formation energies
that are available through www.oqmd.org/download are all
computed using the t-partial correction scheme. In the section
Comparison of OQMD to Experiment, we next provide a detailed
analysis of the agreement between the OQMD and experimental
formation energies, and how the agreement with experiment
depends on the t set being used.
Comparison of OQMD to Experiment. To understand the limits of
applicability of our large database of calculated thermochemical
data, it must be comprehensively validated against a known
reliable source. In the case of the OQMD, we attempt to validate
the predicted energies of formation by comparing with as many
experimental formation energies as possible. In this section, we
compare OQMD values for formation energies to experiment in
several ways. First, we will look at the statistics of the agreement
between OQMD and experiment and investigate the outliers in
the data set. For this analysis we will also consider the effects of
different t sets, described in the section Formation Energy
Calculation, and determine which set of chemical potentials we
nd to be most trustworthy. Then, we explore variation in the
error for various material classes. Finally, we consider how to
correctly distribute error between DFT and experiment, and how
to improve the predictive power of DFT-formation energies.
In Figure 1 we compare the OQMD formation energies to
experimental formation energies, and in Table 4 we present
detailed statistics for the same data. In Figure 1ac, we show the
difference between OQMD and experiment for the t-none, t-
partial and t-all chemical-potential sets, respectively. In the case
of t-none, the average difference is 0.105 eV/atom, with a MAE of
Table 4. Comparison of errors between experimental and OQMD-predicted formation energies for a variety of material classes, under different sets
of elemental chemical potentials
Category ΔHEXP
fNumber of comps. Fit-none Fit-partial Fit-all
Err MAE Err MAE Err MAE
Magnetism
Magnetic 1.231 327 0.135 0.151 0.052 0.102 0.023 0.097
Non-magnetic 1.234 1343 0.098 0.132 0.011 0.093 0.005 0.076
Bandgap
Metallic (E
g
=0) 0.706 921 0.056 0.098 0.031 0.086 0.005 0.077
Semi-conductor (0oE
g
o2) 1.104 249 0.137 0.155 0.017 0.115 0.007 0.093
Wide bandgap (E
g
42) 2.175 491 0.180 0.195 0.003 0.101 0.006 0.083
Number of components
Binary 0.935 1376 0.089 0.124 0.025 0.097 0.002 0.082
Ternary 2.187 259 0.179 0.191 0.002 0.094 0.002 0.082
Quaternary 2.439 33 0.207 0.214 0.019 0.085 0.020 0.071
Bonding type
IVII (ionic) 2.225 18 0.224 0.224 0.090 0.090 0.052 0.057
IIIV (covalent) 0.577 15 0.144 0.146 0.095 0.123 0.056 0.090
Intermetallic binary (metallic) 0.434 278 0.006 0.071 0.002 0.069 0.013 0.069
Binary compounds that contain a/an:
a
Alkali metal 1.047 87 0.120 0.126 0.052 0.081 0.017 0.056
Alkali earth metal 1.256 92 0.161 0.191 0.098 0.128 0.009 0.092
Transition metal 0.727 855 0.051 0.094 0.002 0.079 0.006 0.074
f-block metal 1.106 414 0.098 0.128 0.060 0.098 0.010 0.074
Semi-metal
b
0.573 317 0.046 0.093 0.027 0.093 0.009 0.082
Post-transition metal
c
0.651 233 0.078 0.101 0.028 0.093 0.005 0.075
Halide 1.717 258 0.199 0.214 0.015 0.135 0.014 0.122
Compounds w/o corrections
d
0.514 780 0.031 0.081 0.031 0.081 0.009 0.066
Overall 1.232 1670 0.105 0.136 0.020 0.096 0.002 0.081
Abbreviations: comps, compounds; Err, average error; MAE, mean absolute error; OQMD, Open Quantum Materials Database; w/o, with out.
The denitions of the different ts are given in the section Formation Energy Calculation. Average experimental formation energies (ΔHEXP
f), Err and MAE
are given in eV/atom.
a
At least one of the components in the binary compound belongs to a given subcategory of elements, and the second component can be any element.
b
We take semimetal elements to be B, Si, Ge, As, Sb and Te.
c
We take post-TM elements to be Al, Ga, In, Sn, Tl, Pb and Bi.
d
Includes all compounds that do not contain any elements that require GGA+U or chemical-potential corrections.
The Open Quantum Materials Database
S Kirklin et al
8
npj Computational Materials (2015) 15010 © 2015 Shanghai Institute of Ceramics, Chinese Academy of Sciences/Macmillan Publishers Limited
0.136 eV/atom. Using chemical potentials from the t-partial set,
we nd that the average error is reduced to 0.020 eV/atom and
the MAE is 0.096 eV/atom. Finally, we nd that, by tting
the chemical potentials of all elements, the average error is
0.002 eV/atom and MAE is 0.081 eV/atom, a slight improvement
compared with the difference between t-none and t-partial
chemical potentials.
With such a diverse database of experimental formation
energies to compare with, we are able to look for trends in the
errors across various material classes as a function of chemical-
potential tting. Table 4 compares the errors between experiment
and the OQMD formation energies for a variety of material classes
for the t-none, t-partial and t-all chemical-potential sets. Note
that for all classes of compounds in the t-none set, the average
error (OQMDEXP) is positive, which agrees with the expectation
that, in the generalised gradient approximation, on average DFT
underbinds or underestimates the stability of compounds.
Below, we discuss in more detail the comparison between the
OQMD and experimental formation energies for the specic
classes of materials in Table 4:
Magnetism: The MAE of magnetic-structure formation energies
is larger than that of non-magnetic structures regardless of
chemical-potential choice0.097 vs. 0.076 eV/atom, using the
t-all chemical potentials. Including more complicated magnetic
ordering beyond the ferromagnetism assumed here could lead
to lower-energy structures, reducing the MAE of magnetic-
compound-formation energies.
Bandgap: Without any tting, the difference in the formation
energy of a compound between the OQMD and experiment is
largest for wide bandgap insulators, smaller for semiconductors
and smallest for metals. As expected, formation energy error
magnitudes follow the trend in absolute formation energy
magnitudes for these classes of compounds. Wide bandgap
compounds are expected to be found in material classes such as
nitrides, uorides or oxides, for which the uncorrected chemical
potentials are unreasonable, which results in systematically
skewed formation energies. Once corrections to the chemical
potentials are applied, however, the error becomes largely
independent of band gap.
Number of components: We nd that the average and the MAE
between OQMD-calculated and experimental formation energies
(using the t-none chemical-potential set) increase as the number
of elements in the compound increases from 0.124 to 0.191 to
0.214 eV/atom for the MAE of binary, ternary and quaternary
compounds, respectively. This trend disappears when either the
t-partial or t-all chemical potentials are used. There are only 33
quaternary compound-formation energies to compare between
OQMD and experiment, and all of them are oxides, causing the
effect of chemical-potential tting to be large for this category.
Bonding type: We compare DFT accuracy for three groups
of binary compounds: alkali-metalhalide (IVII), IIIV and inter-
metallic (both elements are metals, including alkali metals, alkaline
earth, lanthanides, actinides, transition metals and poor metals).
These groups correspond roughly to ionic, covalent and metallic
bonding characters, respectively. Without any tting, intermetallic
binary compound-formation energies have the smallest MAE
with respect to experimental formation energies, followed by
covalent binary compound-formation energies, and nally ionic
compound-formation energies have the largest MAE of the three
bonding types. In addition, the accuracy of intermetallic formation
energies is almost completely unaffected by chemical-potential
tting. In contrast, ionic compounds have formation energies that
are systematically more positive than experiment (Err. equals MAE
for t-none and t-partial chemical potential sets in Table 4), and
ionic compound-formation energies were signicantly improved
by both chemical-potential corrections, with their MAE shrinking
from 0.224 to 0.090 to 0.057 eV/atom for t-none, t-partial and
t-all, respectively.
Binary compounds that contain a/an:
Alkali metal: Binary compounds containing alkali metals have
formation energies with MAEs slightly below the overall average
for all chemical-potential sets, and have reductions in MAE from
t-none to t-partial and from t-partial to t-all of 0.045 and
0.025 eV/atom, respectively.
Alkaline earth metal: Alkaline earth binaries have the largest
reduction in MAE between t-partial and t-allan improvement
of 0.036 eV/atom. Most categories have less than a 0.02-eV/atom
improvement over the same range, suggesting that the alkaline
earth elements have systematic errors in their DFT reference state
energies.
Transition metal: As the largest binary compound data set with
855 compounds, transition metal containing binary compounds
have agreement with experiment that is slightly below the overall
average, and show little improvement between the t-partial and
t-all chemical potentials.
f-block element: Binary compounds containing f-block elements
show consistent improvement in MAE across all tting levels and
have MAEs comparable to the overall MAE.
Semi-metal: Semi-metal containing binaries have one of the
lowest MAEs before tting and one of the smallest changes in
MAE between the t-none and t-all chemical-potential sets,
suggesting that the largest error component in these compounds
cannot be addressed by simply adjusting chemical potentials.
Post-transition metal: These compounds show consistent
improvement in accuracy across all levels of tting, and overall
average accuracy.
Halide: Across all tting levels, binary halides have the largest
MAE of all binary compound categories.
Next, we consider several possible sources of error between the
OQMD and experiment. First, there are several ways in which the
OQMD formation energies can be expected to be improved.
One signicant approximation made during our calculations is the
assumption that all magnetic structures are ferromagnetic.
Another potential improvement we could make would be to
assign U-values for more elements and systems. In this study we
Figure 3. Illustration of the lack of agreement between the IIT
47
and
SSUB
46
experimental thermochemical databases. Plots average error
(ΔHSSUBIIT
f) against IIT formation energy (ΔHIIT
f), with the distribution of
errors summarised in the histogram at the bottom. The signicant
range of ΔHIIT
fvalues demonstrates the surprising degree of
disagreement between these experimental formation energy data-
bases. This experiment to experiment comparison for the 75
intermetallic compounds common to both databases gives a MAE
of 0.082 eV/atom, whereas the MAE of OQMD formation energies for
intermetallic compounds is 0.071 eV/atom (using t-nonechemical
potentials).
The Open Quantum Materials Database
S Kirklin et al
9
© 2015 Shanghai Institute of Ceramics, Chinese Academy of Sciences/Macmillan Publishers Limited npj Computational Materials (2015) 15010
applied DFT+U correction only to transition metal oxides and
actinide oxides; however, by applying this correction to more
compositionsboth applying GGA+U corrections to additional
cations and applying those corrections in the presence of
additional anionsfurther improvements in formation energies
may be achieved. For example, local-environment (anion and
oxidation state)-dependent GGA+U calculations have recently
been shown to provide improved thermochemical accuracy in
transition metal oxides and uorides.
51
For systems where
dispersion interactions are important, such as molecular or layered
crystals, more accurate predictions may require van der Waals
inclusive methods beyond GGA.
32,52
Lastly, we should note that
DFT-predicted bulk formation energies serve as a T= 0 K starting
point for further thermodynamic analysis. Enthalpic and entropic
contributions at nite temperatures such as lattice dynamics,
congurational defects and orderdisorder transitions can be
captured more accurately with relevant statistical, mechanical and
DFT methods.
Assessing the Accuracy of Experimental Formation Energies.
Improvements to the DFT calculation scheme may lead to some
reductions in the discrepancy between the OQMD and experi-
mental formation energies. However, some of the errors can also
be attributed to the experimental formation energies themselves.
We wish to ascertain the size of this error or uncertainty. With
multiple experimental data sources to draw from, we can compare
experimental measurements of a given compound with one
another. Figure 3 shows the discrepancies between formation
energies from different experimental sources for the same
compounds. The resulting MAE of the experimental values is
surprisingly large, 0.082 eV/atom. This experimental error is
calculated based on a comparison between the 75 compounds
common to both of the experimental databases used in this study.
Note that this comparison is limited by the fact that the IIT
database is strictly intermetallic compounds, and therefore all of
the energies compared are for intermetallics. Of course we
acknowledge that not all experimental data are equally reliable or
accurate and that advances in techniques can yield more accurate
data. However, these comparisons are for curated databases, and
therefore might reasonably be expected to represent a high
degree of experimental accuracy.
The OQMD formation energies (using the t-nonechemical
potentials, i.e., uncorrected DFT formation energies) have a MAE
relative to experiment of 0.071 eV/atom for similar compounds,
i.e., intermetallics, which is slightly less than the experimental error
with a second experiment. For this same set of 75 compounds, we
compare the OQMD formation energy with each of the experi-
mental values. The minimum MAE between experiment and DFT
(i.e., comparing DFT with the experimental formation energy
closer in energy to the DFT formation energy for each compound)
is 0.057 eV/atom, while the maximum MAE between experiment
and DFT (i.e., comparing DFT with the experimental formation
energy farther away in energy from the DFT formation energy for
each compound) is 0.116 eV/atom. From this result it is clear that,
where experiments disagree, DFT is often signicantly closer to
one experiment than the other.
Given this level of disparity in experimental formation energies,
it is highly unlikely that all of the errors should be attributed to
DFT. In fact, without additional information, it is impossible to
fairly determine which values are in error. We conclude from these
comparisons that there remains a need for additional experimental
thermochemical data. This is particularly urgent as many
computational schemes
45,49,50
rely on these experimental values
to obtain elemental correction factors.
To explore the source of disparity between DFT and experiment
further, we looked at several of the compounds that have the
poorest agreement with experiment, and searched the literature
for alternative values for their formation energies. For the cases
where we were able to nd another value, we show the
composition, SSUB
46
formation energy and the literature forma-
tion energy in (Table 5) these compounds with very large
discrepancies between experimental formation energiesthe
second value we found in literature is often closer to the
OQMD-predicted formation energy. As a result, we conclude that
some of the very signicant disagreements between DFT and
experiment are more likely to be due to experimental or
transcription errors than to problems in DFT. On the basis of
these ndings, we believe that for many other compounds with
large formation-energy errors, and for which no alternative
formation energies could be found, the source of error might
also be the experimental measurement, rather than only the DFT
calculation.
Comparison of OQMD to Other DFT Databases and the Miedema
Model. The Miedema model has historically been widely used
to provide estimates of formation enthalpies of solid alloys
and intermetallic compounds.
66
The Miedema model is a semi-
empirical approach wherein atoms are conceptually treated as
space-lling polyhedra. Chemical bonding is treated by consider-
ing the overlap of the surface areas of neighbouring atomic
polyhedra, weighted by the difference in charge density of each
atom at the boundary and the electronegativity difference
between the atoms.
66
The model contains several element-
dependent parameters that Miedema t to trends in the formation
energies of a range of binary intermetallic compounds, as well as
elemental properties (bulk modulus, molar volume and work
function), which were adjusted to give the best t to experimental
Table 5. Comparison of SSUB database
46
and alternative sources for
experimental formation energies
Composition SSUB ΔH
f
(eV/atom) alternative OQMD
LiNbO
3
5.660 2.774
a
2.739
Eu
2
O
3
3.437 3.423
b
2.574
SiB
6
0.173 0.181
c
0.455
EuF
3
4.104 4.195
d
3.483
MnSe 0.889 0.922
e
0.363
InN 0.500 0.148
f
0.021
AlB
2
0.521 0.055
g
0.044
PtZr 1.561 0.99
d
1.086
Ba
2
Pb 1.012 0.694
h
0.540
PuCl
3
2.614 2.486
i
2.176
PrO
2
3.280 3.316
j
2.846
CaAl
2
0.759 0.346
k
0.335
BaSiO
3
3.354 3.330
l
2.945
CeN 1.715 1.763
m
1.326
Abbreviations: OQMD, Open Quantum Materials Database; SSUB, SGTE
Solid SUBstance.
Comparison made for outlying compounds that show a large disagree-
ment between SSUB data and OQMD-predicted formation energies (using
t-partial chemical potentials).
Differences between SSUB and alternative-source formation energies
ranging from 0.008 eV/atom (SiB
6
) to 2.885 eV/atom (LiNbO
3
).
a
Reference 53.
b
Reference 54.
c
Reference 55.
d
Reference 56.
e
Reference 57.
f
Reference 58.
g
Reference 59.
h
Reference 60.
i
Reference 61.
j
Reference 62.
k
Reference 63.
l
Reference 64.
m
Reference 65.
The Open Quantum Materials Database
S Kirklin et al
10
npj Computational Materials (2015) 15010 © 2015 Shanghai Institute of Ceramics, Chinese Academy of Sciences/Macmillan Publishers Limited
formation energies. A comparison of the accuracy of the Miedema
model with the accuracy of the OQMD is important as the
Miedema model is still actively employed.
67,68
Table 6 contains a
comparison of Miedema model predictions for formation energy
to experiment. The MAE between the Miedema model and
experiment is 0.199 eV/atom, greater than twice that of the OQMD
for the same set of compounds, 0.090 eV/atom. This result
indicates that, in addition to the inherent drawbacks of the
Miedema model (i.e., applicable only to binary intermetallics), the
OQMD is a much more accurate predictor than the Miedema
model for formation energies.
The Materials Project
11,24
was one of the rst high-throughput
databases to be developed. As the Materials Project database uses
slightly different calculation parameters from OQMD, and a
different chemical potential correction scheme, there is an
interesting opportunity to directly compare the results of different
DFT databases with one another. Table 6 shows the statistics of
the agreement between Materials Project and experimental
formation energies. The average error of formation energies for
the Materials Project is 0.006 eV/atom, smaller than the average
error of the OQMD formation energies, 0.032 eV/atom. However,
the MAE for the Materials Project is 0.133 eV/atom, which is larger
than that of the OQMD over the same compounds, which is
0.108 eV/atom.
We attribute the difference in MAE with experiment between
OQMD and Materials Project to the difference in chemical-
potential tting procedures for the two data sets.
11,24
The
chemical-potential tting used in the OQMD is performed on
the same set of compounds on which the computed accuracy is
based, which gives the OQMD a natural advantage.Further
evidence to support this argument can be found by calculating
the mean absolute difference between the OQMD and Materials
Project formation energies for all compounds for which we do not
t any chemical potentials. For this set of 563 compounds, the
mean absolute difference between the OQMD and Materials
Project is 0.028 eV/atom, much smaller than the difference
between the OQMD and experiment (0.093 eV/atom for these
563 compounds) or between Materials Project and experiment
(0.086 eV/atom for these 563 compounds). As a result, we
conclude that in general the two databases contain very similar
results, and that different choices for DFT parameters have a much
smaller impact on compound-formation energies than do the
different approaches to chemical-potential tting. Finally, as new
calculations are continuously added to both OQMD and Materials
Project, the analysis above corresponds only to certain snapshots
of each database; however, we expect this conclusion to be valid
as long as the chemical-potential tting approaches are not
signicantly revised.
Historical Trends in Material/Compound Discovery
A large thermodynamic database of energetics and phase stability
of the type presented here can be used to address many
interesting general trends. For instance, we leverage the fact that
we have evaluated a signicant fraction of known ground-state
compounds to answer several questions about trends and
patterns in material discovery and stability. Without a large
database such as the OQMD it would otherwise be impossible to
answer many of these questions.
How many stable compounds are in the database? How many of
these are experimentally known versus theoretically predicted? In
order to answer these questions, rst we determine phase
stability. Phase stability is determined by constructing the energy
convex hull of a given region of composition space.
69
Once this
has been determined using existing computational geometry
algorithms,
70
(Kirklin, S. & Wolverton, C. (2015, unpublished)) every
phase that lies on the convex hull is stable at T= 0 K (i.e., it is lower
in energy than any other phase or combination of phases in the
database). Of the 297,099 calculated compounds in the OQMD, we
nd that 19,757 are thermodynamically stable at T= 0 K. Of these,
16,526 were from the ICSD with the remaining 3,231 being
prototype structures.
All the 3,231 compounds that we predict to be stable, but are
not in the ICSD, represent new compounds to be discovered.
The prototype compounds were constructed from commonly
occurring, simple, crystal structures, and do not represent an
exhaustive crystal structure determination for each predicted
compound. For this reason, we do not assert that in all cases the
predicted compounds are stable in the crystal structure we list.
Rather, in these cases, our predicted convex hull is an upper
bound to the true ground-state hull. Thus, for all the 3,231 cases,
we predict that some new compound(s) are awaiting experi-
mental discovery in these (binary and ternary) systems. A detailed
crystal structure search in such systems can be made using
evolutionary
71
or minima hopping
72
methods. Furthermore,
because the prototype compounds have identied so many holes
in our knowledge of ground-state phase stability, we expect that
by including the prototype structures in our list of ground-state
structures we are providing a better estimate of the energy
landscape where we do not have experimentally measured
structures.
What is the rate of stable material discovery? New materials and
compounds are being discovered all the time, some of which are
stable and some of which are not. Utilising the publication data
associated with ICSD records, we can study the historical rate of
material discovery. In Figure 4a we plot the total rate of material
discovery as the number of new compounds reported per year
since 1910. Each compound only appears in the year in which it
was rst reported. In Figure 4b, we plot the number of stable ICSD
compounds discovered per year. The material discoverydata in
Figure 4a come directly from the ICSD (no DFT is required).
However, to classify a compound as stableis not possible from
the ICSD alone but requires some measures of energetics, as well
as those of competing (combinations of) phases. The latter is
possible only with a large material database containing formation
energies, such as the OQMD.
We nd that the rate of compound discovery is increasing with
timein most years, more compounds are discovered than the
year before. In contrast, the rate of stable material discovery has
been fairly constant since the 1960s. By decomposing the number
of discovered materials into binary, ternary, quaternary and
pentanary compounds, we observe that the number of stable
binary compounds discovered each year has been dwindling since
the 1970s, when the number of ternary compounds discovered
began to signicantly rise. As of the 1990s, the number of stable
ternaries has also stagnated, while the number of stable
quaternary compounds began to increase.
Table 6. Comparison of predicted formation energies with
experimental values for two DFT databases (the OQMD with t-partial
corrections and Materials Project) and an empirical model for
intermetallic compounds (Miedema model)
Materials Project
a
Miedema
Number of compounds 1,386 820
ΔHXEXP
fAvg Err 0.006 0.033
MAE 0.133 0.199
ΔHOQMDEXP
fAvg Err 0.032 0.029
MAE 0.108 0.090
Abbreviations: API, application programming interface; Avg Err, average
error; DFT, density functional theory; MAE, mean absolute error; OQMD,
Open Quantum Materials Database.
Average error (ΔHXEXP
f) and MAE (9ΔHXEXP
f9) are given in eV/atom.
a
Data from the Materials Project used in this comparison were retrieved via
the Materials Project API
24,26
on 16/12/2013.
The Open Quantum Materials Database
S Kirklin et al
11
© 2015 Shanghai Institute of Ceramics, Chinese Academy of Sciences/Macmillan Publishers Limited npj Computational Materials (2015) 15010
What compositions are most likely to be stable? On the basis of
the same stability data, we can also look at which compositions
are most frequently stable. In Figures 5 and 6 we provide
histograms showing the frequency at which compositions are
stable for binary and ternary compositions, respectively. We nd
that the most commonly stable binary composition is at 1:1 (AB),
followed by 3:1 (A
3
B) and 2:1 (A
2
B). For ternary compounds, the
most common composition is 2:1:1 (A
2
BC), which is the
composition of the L2
1
(Heusler) prototype. We believe that the
preponderance of newly predicted L2
1
compounds is primarily
because L2
1
is the only ternary prototype that we have calculated
in a wide range of compositions. Many ternary systems that have
favourable ordering, but for which the true ground state is
unknown, could yield predictions that the L2
1
compound is stable
(the upper bound to the convex hull), demonstrating the need for
further exploration of a wide range of composition spaces.
Following this interpretation, we checked every stable ternary L2
1
prototype in the OQMD and searched for any ternary ICSD
compound in that systemthat is, if Ca
2
GaLi is a stable Heusler, is
there any ternary Ca-Ga-Li compound reported in the ICSD? Of the
2,290 stable Heusler prototypes in the OQMD, only 781, or 22%,
have any ICSD structure in that region of phase space. In the
remaining ~ 1,500 ternary systems, we predict the existence of
stable compounds waiting to be discovered.
In order to facilitate the discovery of new, stable compounds in
the thousands of regions of composition space where we predict
stable compounds to exist, we provide a full list of compositions
where we predict a prototype to be stable online. We break this
list into (i) prototypes that are more stable than an experimentally
measured structure and (ii) prototypes that have no experimental
structure at that composition. In the rst case, nite-temperature
effects may cause the formation energy of the experimental
structure to lower relative to the formation energy of the
prototype. However, in the second case, some compounds should
be found at the listed composition, although possibly not the
prototype. A current list of predicted compounds can be found
online at www.oqmd.org/materials/discovery where one can
obtain the entire list or lter by composition.
CONCLUSIONS
The OQMD is a high-throughput database of DFT calculations of
32,559 ICSD structures and 259,511 prototypical structures,
growing steadily as new structures are added continuously.
OQMD is available for download without restrictions at www.
oqmd.org/download. Included in the download is a complete
framework for performing additional calculations that are
commensurate with the database. We use the breadth of the
OQMD to compare DFT-calculated structures and formation
energies with experiment at an unprecedented scale. We nd
the following:
Elemental Ground-State Structures. Using the capabilities of qmpy
and the OQMD, we nd that for 77 out of the 89 elements DFT as
implemented in VASP at the settings of the OQMD is able to
correctly predict the observed ground-state structure as being
lowest in energy out of 20 possible structures, chosen from among
the known ground states of all elements. In all cases where DFT
nds a lower-energy structure, the observed ground state is
nearly degenerate with the lowest-energy structure (by far, the
largest errors being phosphorous and mercury, with 0.036 and
0.074 eV/atom error, respectively).
Formation Energies. In order to most accurately determine
compound-formation energies, we evaluate the effects of three
different choices of chemical potentials: using DFT ground-state
reference energies, tting chemical potentials to experiment for
elements where the DFT ground state differs signicantly from the
room temperature stable phase, and tting chemical potentials for
Figure 4. (a) The total number of compound discovery within the
ICSD by year. (b) The number of stable (T=0 K) compound discovery
in the ICSD, where the stability is assesed by the OQMD energies.
The year for a structure corresponds to the earliest publication year
for ICSD entries at that given structures composition.
Figure 5. Distribution of stable binary compounds as a function of
composition. Note the presence of large peaks at low integer ratios,
i.e., 1:1, 2:1 and 3:1.
Figure 6. Distribution of stable ternary compounds as a function of
composition. Plotted on log scale to account for the extremely high
density of phases at A
2
BC compositions because of the calculation
of over 180,000 decorations of the L2
1
structure.
The Open Quantum Materials Database
S Kirklin et al
12
npj Computational Materials (2015) 15010 © 2015 Shanghai Institute of Ceramics, Chinese Academy of Sciences/Macmillan Publishers Limited
all elements (labelled t-none, t-partial and t-all). We nd that
t-partial exhibits signicant improvements over t-none, with a
reduction in MAE against experiment from 0.136 to 0.096 eV/atom.
In comparison, t-all has only marginal gains over t-partial, with
the MAE reducing to 0.081 eV/atom. However, by increasing the
number of t parameters, we also increase the risk of over-tting,
and, as a result, take t-partial to be the optimal choice of
chemical potentials for predicting formation energies.
To put the discrepancy between the OQMD and experiment
into appropriate context, we also compared formation energies
between the two database experimental formation energies used
in this work (the SSUB database,
46
and the thermodynamic
database at the Thermal Processing Technology Center at the
IIT).
47
We nd that the MAE between these two databases is
0.082 eV/atom, a value that is similar to the error between the
OQMD and experiment. As a result, it is impossible to assign all of
the errors between DFT and experiment solely to DFT, and leads
us to conclude that, in order to establish a more accurate
evaluation of the accuracy of formation energies based on DFT
total energies, more accurate measurements of formation
energies should be undertaken in future.
We also compare the OQMD with two other databases
of calculated formation energies: the Miedema model and
the Materials Project. We compare the Miedema model with
experiment for all compositions for which we have (i) an
experimental formation energy, (ii) an OQMD formation energy,
and (iii) the Miedema model is applicable. Over the resulting pool
of 820 compounds, the OQMD has a MAE of 0.090 eV/atom, less
than half the error found in the Miedema model,
0.199 eV/atom. We made an identical comparison with the
Materials Project, which had a comparison pool of 1,386
compounds. For this set of compounds, we found that the
Materials Project had an MAE of 0.133 eV/atom, which is slightly
larger than the OQMD error for the same set of compounds,
0.108 eV/atom. By comparing cases in which the formation energy
is calculated without any chemical-potential tting with cases in
which the chemical potentials have been t to experiment, we
determine that the majority of the difference between the error in
the OQMD and Materials Project is attributed to differences in the
chemical-potential-tting approach.
Historical Trends in Material/Compound Discovery. The OQMD
allows us to explore trends, both historical and stoichiometric, in
compound discovery and stability. From a historical perspective,
the number of reported structures has been increasing linearly
with time, while the number of stable structures reported annually
has remained roughly constant since the 1960s. In addition,
the scope of compound discovery has progressed from binary to
ternary to quaternary compounds over time. This trend has now
been disrupted by recent advances in structure prediction,
structure determination and high-throughput structure
calculation. In particular, in this study we predict the existence
of ~ 3,200 new compounds. Using our predictions of the
compositions at which new phases should be found, experi-
mentalists can now more efciently discover and characterise new
materials.
In this work we demonstrate a few examples of how a large
database of DFT calculations can be used to extract information
beyond what can be gleaned from many distributed collections
containing similar data. We believe that there is much more that
can be understood by looking at large-scale material property
databases, and in order to facilitate such discovery, we make the
entire database available for download, without restriction.
MATERIALS AND METHODS
All DFT calculations are performed with the VASP
73,74
(v5.3.2). The electron
exchange and correlation are described with the GGA of PBE,
75
using the
potentials supplied by VASP with the PAW method.
76
PAW-PBE potentials
for 89 elements are supplied with VASP, and those employed in the OQMD
are listed in Table 3. We follow the VASP guidelines concerning the
optimum choice of potentials.
77
Potentials where electrons have been
moved from the core and treated as valence are appended with _sv, _pv
and _d in Table 3 for s-, p- and d-electrons, respectively. For the
4f-elements, we employ potentials where the valence f-electrons are
treated as core electrons (appended with _2 and _3). For all calculations,
Γ-centred k-point meshes are constructed to have the same relative ratios
of mesh points to reciprocal lattice vector length, and with the number
of k-points such that the k-points per reciprocal atom (calculated from
N
kpoints
×N
atoms
) is as close to a target value as possible. The electronic
self-consistency (for a given set of ion positions) is converged to within
10
4
eV/atom.
Any calculation containing d-block or actinide elements are spin-
polarised with a ferromagnetic alignment of spins to capture possible
magnetism, with initial magnetic moments of 5 and 7 μ
B
for the d-block
and actinide elements, respectively. It should be noted that this approach
will not capture more complex magnetic ordering, such as antiferro-
magnetism. For many d-block oxides, a typical difference in total energy
between ferromagnetic and antiferromagnetic states is on the order of
1020 meV/atom.
45
However, given that in this work the range of
compounds being calculated is extremely broad, it is likely that in some
cases this error will be larger.
For several d- and f-block elements listed in Table 1, the GGA+U
approach is implemented to improve the exchange and correlation
description of the localised charge density when these elements are in
compounds with oxygen. We employed the Dudarev approach to GGA+
U,
78
where the only input parameter is UJ. For several transition metals,
previously determined UJvalues were used.
49
For actinide elements in
oxides, we apply a UJparameter of 4 eV. This value was chosen because
no reliable values had been reported in the literature when these
calculations were begun, and we found that the formation energies and
band gaps of compounds containing these elements are relatively
insensitive to the exact value of UJ; therefore, we elected to use a
moderate UJvalue of 4 eV for these elements. Recent lines of work have
identied UJvalues for a few of these elements,
51,79
with UJvalues
close to those used herein. All UJvalues are given in Table 1.
All calculations were completed in a two-step scheme. First, the
structures were fully relaxed, followed by a static calculation. In relaxing
an ICSD structure, we begin with the given ICSD structure parameters and
perform several relaxation runs sequentially, until the volume change
within the last relaxation run is less than 10%. The relaxation calculations
are performed at a plane-wave basis-set energy cutoff at the energy
recommended in the VASP potentials of the elements in the structure, and
6,000 k-points per reciprocal atom. The quasi-Newton scheme is used to
optimise the structure to within 10
3
eV/atom. In these relaxation steps,
Gaussian smearing is applied with a width of 0.2 eV. The nal static
calculation of the structure is performed at an energy cutoff of 520 eV
using tetrahedral k-point integration.
76
The 520-eV cutoff is chosen
because it is 25% higher than the highest recommended energy cutoff
over all of the potentials used (including Li_sv, for which we use the
version with lower recommended cutoff). This constant cutoff for all
calculations ensures that all the energies calculated in OQMD are
compatible, and can be used to evaluate the formation energies of
compounds and T= 0 ground-state phase diagrams.
ACKNOWLEDGEMENTS
We acknowledge funding support from the following sources: the Center for
Electrical Energy Storage: Tailored Interfaces, an Energy Frontier Research Center
funded by the U.S. Department of Energy, Ofce of Science, Ofce of Basic Energy
Sciences (SK); the Revolutionary Materials for Solid State Energy Conversion, an
Energy Frontier Research Center funded by the U.S. Department of Energy, Ofce of
Science, Ofce of Basic Energy Sciences under Award Number DE-SC00010543 (JWD);
the Ford-Boeing-Northwestern Alliance (JES); the Department of Defense through the
National Defense Science and Engineering Graduate Fellowship Program with further
support by DOE under Grant No. DE-FG02-07ER46433 (BM and AT); The Dow
Chemical Company (MA) and the National Science Foundation under grant
DRL-1348800 (CW). We acknowledge the openness of the Materials Project for
making it possible to obtain a large set of energies with which to compare with the
OQMD and experiment. We acknowledge a large allocation on Northwestern
Universitys Quest high-performance computing system. Calculations were also
performed on resources of the National Energy Research Scientic Computing
The Open Quantum Materials Database
S Kirklin et al
13
© 2015 Shanghai Institute of Ceramics, Chinese Academy of Sciences/Macmillan Publishers Limited npj Computational Materials (2015) 15010
Center, which is supported by the Ofce of Science of the U.S. Department of Energy
under Contract No. DE-AC02-05CH11231.
CONTRIBUTIONS
SK developed DFT automation and thermodynamic analysis tools, and had primary
writing responsibilities. JWD carried out comparison with Miedema model. JES
carried out comparison with Materials Project, evaluated SSUB formation energies at
STP. BM and AT selected DFT+U parameterisations and VASP potentials. MA analysed
the elemental ground states. BM and SR organised the partnership with the ICSD. MA,
JWD, JES and CW extensively edited the manuscript. CW conceptualised, guided all
aspects and led the project.
COMPETING INTERESTS
The authors declare no conict of interest.
REFERENCES
1. National Science and Technology Council. Materials Genome Initiative for Global
Competitiveness Tech. Rep. http://www.whitehouse.gov/blog/2011/06/24/mate
rials-genome-initiative-renaissance-american-manufacturing (2011).
2. Hohenberg, P. & Kohn, W. Inhomogeneous electron gas. Phys. Rev. 136,
864871 (1964).
3. Kohn, W. & Sham, L. J. Self-consistent equations including exchange and
correlation effects. Phys. Rev. 140, 11331138 (1965).
4. Perdew, J. P. & Zunger, A. Self-Interaction correction to density-functional
approximations for many-electron systems. Phys. Rev. B 23, 50485079 (1981).
5. Ihm, J., Zunger, A. & Cohen, M. L. Momentum-space formalism for the total
energy of solids. J. Phys. C Solid State Phys. 12, 44094422 (1979).
6. Ceperley, D. M. & Alder, B. J. Ground state of the electron gas by a
stochastic method. Phys. Rev. Lett. 45,566569 (1980).
7. Ihm, J., Yin, M. T. & Cohen, M. L. Quantum mechanical force calculations in solids:
the phonon spectrum of Si. Solid State Commun. 37,491494 (1981).
8. Levy, O., Hart, G. L. W. & Curtarolo, S. Uncovering compounds by synergy of
cluster expansion and high-throughput methods. J. Am. Chem. Soc. 132,
48304833 (2010).
9. Curtarolo, S. et al. The high-throughput highway to computational
materials design. Nat. Mater. 12, 191201 (2013).
10. Hautier, G., Fischer, C., Ehrlacher, V., Jain, A. & Ceder, G. Data mined ionic sub-
stitutions for the discovery of new compounds. Inorg. Chem. 50,656663 (2010).
11. Jain, A. et al. A high-throughput infrastructure for density functional theory
calculations. Comput. Mater. Sci. 50, 22952310 (2011).
12. Armiento, R., Kozinsky, B., Fornari, M. & Ceder, G. Screening for high-performance
piezoelectrics using high-throughput density functional theory. Phys. Rev. B 84,
014103 (2011).
13. Setyawan, W., Gaume, R. M., Lam, S., Feigelson, R. S. & Curtarolo, S.
High-throughput combinatorial database of electronic band structures for
inorganic scintillator materials. ACS Comb. Sci. 13, 382390 (2011).
14. Saal, J., Kirklin, S., Aykol, M., Meredig, B. & Wolverton, C. Materials design and
discovery with high-throughput density functional theory: the Open Quantum
Materials Database (OQMD). JOM 65, 15011509 (2013).
15. Curtarolo, S. et al. AFLOW: an automatic framework for high-throughput materials
discovery. Comput. Mater. Sci. 58,218226 (2012).
16. Landis, D. D. et al. The computational materials repository. Comput. Sci. Eng. 14,
5157 (2012).
17. Bergerhoff, G. & Brown, I. D. Crystallographic Databases. chap. Inorganic
147156 (International Union of Crystallography: Chester, 1987).
18. Belsky, A., Hellenbrandt, M., Karen, V. L. & Luksch, P. New developments in the
Inorganic Crystal Structure Database (ICSD): accessibility in support of materials
research and design. Acta Crystallogr. B 58, 364369 (2002).
19. Saal, J. E. & Wolverton, C. Thermodynamic stability of Mg-Y-Zn long-period
stacking ordered structures. Scr. Mater. 67,798801 (2012).
20. Kirklin, S., Meredig, B. & Wolverton, C. High-throughput computational screening
of new Li-ion battery anode materials. Adv. Energy Mater. 3, 252262 (2013).
21. Meredig, B. et al. Combinatorial screening for new materials in unconstrained
composition space with machine learning. Phys. Rev. B 89, 094104 (2014).
22. Aykol, M., Kirklin, S. & Wolverton, C. Thermodynamic aspects of cathode coatings
for lithium-ion batteries. Adv. Energy Mater. 4, 1400690 (2014).
23. Kirklin, S., Chan, M., Trahey, L., Thackeray, M. M. & Wolverton, C. M.
High-throughput screening of high-capacity electrodes for hybrid Li-ion/Li-O
2 cells. Phys. Chem. Chem. Phys. 16, 2207322082 (2014).
24. Jain, A. et al. Commentary: The Materials Project: a materials genome approach to
accelerating materials innovation. APL Mater. 1, 011002 (2013).
25. Curtarolo, S. et al. AFLOWLIB.ORG: a distributed materials properties repository
from high-throughput ab initio calculations. Comput. Mater. Sci. 58,227235
(2012).
26. Ong, S. P. et al. Python Materials Genomics (pymatgen): a robust, open-source
python library for materials analysis. Comput. Mater. Sci. 68,314319 (2013).
27. Bahn, S. R. & Jacobsen, K. W. An object-oriented scripting interface to a legacy
electronic structure code. Comput. Sci. Eng. 4,5666 (2002).
28. Meredig, B. & Wolverton, C. A hybrid computational-experimental approach for
automated crystal structure solution. Nat. Mater. 12,123
127 (2013).
29. Young, D. A. Phase Diagrams of the Elements (University of California Press:
Berkeley and Los Angeles, 1991).
30. King, H. W. in CRC Handbook of Chemistry and Physics 95th edn (ed. Haynes W. M.)
Ch. 12, 1518 (CRC Press, Taylor & Francis, 2012).
31. Wang, Y. et al. Ab initio lattice stability in comparison with CALPHAD lattice
stability. Calphad 28,7990 (2004).
32. Klimeš,J.&Michaelides,A.Perspective:advances and challenges in treating van der
Waals dispersion forces in density functional theory. J. Chem. Phys. 137, 120901 (2012).
33. Tao, X. et al. Phase stability of magnesium-rare earth binary systems from
rst-principles calculations. J. Alloys Compd. 509, 68996907 (2011).
34. Gao, M. C., Rollett, A. D. & Widom, M. First-principles calculation of lattice stability
of C15M2R and their hypothetical C15 variants (M = Al, Co, Ni; R = Ca, Ce, Nd, Y).
Calphad 30,341348 (2006).
35. Mao, Z., Seidman, D. N. & Wolverton, C. First-principles phase stability, magnetic
properties and solubility in aluminumare-earth (AlRE) alloys and compounds.
Acta Mater. 59, 36593666 (2011).
36. Temmerman, W. et al. Handbook on the Physics and Chemistry of Rare Earths
vol. 39 (Elsevier, 2009); http://www.sciencedirect.com/science/article/pii/
S0168127308000019.
37. Zhou, F. & Ozoliņš, V. Obtaining correct orbital ground states in f-electron
systems using a nonspherical self-interaction-corrected LDA+U method. Phys.
Rev. B 80, 125127 (2009).
38. Duthie, J. C. & Pettifor, D. G. Correlation between d-band occupancy and crystal
structure in the rare earths. Phys. Rev. Lett. 38,564567 (1977).
39. Biering, S. & Schwerdtfeger, P. High-pressure transitions in bulk mercury:
a density functional study. Theor. Chem. Acc. 130, 455462 (2011).
40. Mishra, V., Gyanchandani, J., Chaturvedi, S. & Sikka, S. Effect of spin-orbit coupling
on the ground state structure of mercury. Solid State Commun. 186,3841 (2014).
41. Wolverton, C. Crystal structure and stability of complex precipitate phases in
Al-Cu-Mg-(Si) and Al-Zn-Mg alloys. Acta Mater. 49, 31293142 (2001).
42. Wolverton, C., Ozoliņš, V. & Asta, M. Hydrogen in aluminum: First-principles
calculations of structure and thermodynamics. Phys. Rev. B 69, 144109 (2004).
43. Lany, S. Semiconductor thermochemistry in density functional calculations. Phys.
Rev. B 78, 245207 (2008).
44. Hautier, G., Ong, S. P., Jain, A., Moore, C. J. & Ceder, G. Accuracy of density
functional theory in predicting formation energies of ternary oxides from binary
oxides and its implication on phase stability. Phys. Rev. B 85, 155208 (2012).
45. Stevanović, V., Lany, S., Zhang, X. & Zunger, A. Correcting density functional
theory for accurate predictions of compound enthalpies of formation: tted
elemental-phase reference energies. Phys. Rev. B 85, 115104 (2012).
46. SGTE. Thermodynamic Properties of Inorganic Materials Vol. 19. (Springer-Verlag:
Berlin, Heidelberg, 1999).
47. Nash, P. Thermodynamic database https://tptc.iit.edu/index.php/thermo-data
base (2013).
48. Konings, R. J. M., Morss, L. R., Fuger, J. in The Chemistry of the Actinide and
Transactinide Elements 3rd edn (eds Morss L. R. et al.) Ch. 19, 21132224 (Springer:
Dordrecht, 2006); http://www.springerlink.com/index/l17213068108mk22.pdf.
49. Wang, L., Maxisch, T. & Ceder, G. Oxidation energies of transition metal oxides
within the GGA+U framework. Phys. Rev. B 73, 195107 (2006).
50. Grindy, S., Meredig, B., Kirklin, S., Saal, J. E. & Wolverton, C. Approaching chemical
accuracy with density functional calculations: diatomic energy corrections. Phys.
Rev. B 87, 075150 (2013).
51. Aykol, M. & Wolverton, C. Local environment dependent GGA+U method for
accurate thermochemistry of transition metal compounds. Phys. Rev. B 90,
115105 (2014).
52. Aykol, M., Kim, S. & Wolverton, C. van der waals interactions in layered lithium
cobalt oxides. J. Phys. Chem. C 119, 1905319058 (2015).
53. Pozdnyakova, I., Navrotsky, A., Shilkina, L. & Reznitchenko, L. Thermodynamic and
structural properties of sodium lithium niobate solid solutions. J. Am. Ceram. Soc.
85,379384 (2004).
54. CRC. Handbook of Chemistry and Physics, 93rd edn 20122013
http://www.hbcpnetbase.com/ (2012).
55. Gordienko, S. P. Enthalpies of formation for boron silicides. Powder Metall. Met.
Ceram. 34,660662 (1995).
56. Kubaschewski, O., Alcock, C. B. & Spencer, P. J. Materials Thermochemistry 6th edn
(Pergamon Press, 1993).
The Open Quantum Materials Database
S Kirklin et al
14
npj Computational Materials (2015) 15010 © 2015 Shanghai Institute of Ceramics, Chinese Academy of Sciences/Macmillan Publishers Limited
57. Olin, Å., Noläng, B., Öhman, L.-O., Osadchii, E. & Rosén, E. Chemical Thermo-
dynamics (Elsevier Science, 2005).
58. Ranade, M. R., Tessier, F., Navrotsky, A. & Marchand, R. Calorimetric determination
of the enthalpy of formation of InN and comparison with AlN and GaN. J. Mater.
Res. 16, 28242831 (2001).
59. van Setten, M. & Fichtner, M. On the enthalpy of formation of aluminum
diboride, AlB2. J. Alloys Compd. 477, L11L12 (2009).
60. Sommer, F., Borzone, G., Parodi, N. & Ferro, R. Enthalpy of formation of CaPb and
BaPb alloys. Intermetallics 14, 287296 (2006).
61. Lemire, R. J. Chemical Thermodynamics of Neptunium and Plutonium Vol. C
(Elsevier Science, 2001).
62. Gramsch, S. A. & Morss, L. R. Standard molar enthalpies of formation of PrO2 and
SrPrO3: the unusual thermo dynamic stability of APrO3 (A = Sr,Ba). J. Chem.
Thermodyn. 27, 551560 (1995).
63. Notin, M., Mejbar, J., Bouhajib, A., Charles, J. & Hertz, J. The thermodynamic
properties of calcium intermetallic compounds. J. Alloys Compd. 220,
6275 (1995).
64. Huntelaar, M. E., Cordfunke, E. H. P. & Ouweltjes, W. The standard molar
enthalpies of formation of BaSiO3 (s) and Ba2SiO4 (s). J. Chem. Thermodyn. 24,
10991102 (1992).
65. Tessier, F. et al. Thermodynamics of formation of binary and ternary nitrides in
the system Ce/Mn/N. Z. Anorg. Allg. Chem. 627,194200 (2001).
66. de Boer, F. R., Boom, R., Mattens, W. C. M., Miedema, A. R. & Niessen, A. K. Cohesion
in Metals: Transition Metal Alloys (North-Holland, Amsterdam 1988).
67. Das, N. et al. Miedema model based methodology to predict amorphous-
forming-composition range in binary and ternary systems. J. Alloys Compd. 550,
483495 (2013).
68. Ray, P. K., Akinc, M. & Kramer, M. J. Applications of an extended Miedemas model
for ternary alloys. J. Alloys Compd. 489,357361 (2010).
69. Akbarzadeh, A. R., Ozoliņš, V. & Wolverton, C. First-principles determination of
multicomponent hydride phase diagrams: application to the Li-Mg-N-H system.
Adv. Mater. 19, 32333239 (2007).
70. Barber, C., Dobkin, D. & Huhdanpaa, H. The quickhull algorithm for convex hulls.
ACM Trans. Math. Softw. 22, 469483 (1996).
71. Oganov, A. R. & Glass, C. W. Crystal structure prediction using ab initio
evolutionary algorithms: principles and applications. J. Chem. Phys. 124,
244704 (2006).
72. Amsler, M. & Goedecker, S. Crystal structure prediction using the minima
hopping method. J. Chem. Phys. 133, 224104 (2010).
73. Kresse, G. & Furthmuller, J. Efciency of ab-initio total energy calculations for
metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 6,
1550 (1996).
74. Kresse, G. & Furthmüller, J. Efcient iterative schemes for ab initio total-energy
calculations using a plane-wave basis set. Phys. Rev. B 54, 1116911186 (1996).
75. Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation
made simple. Phys. Rev. Lett. 77, 38653868 (1996).
76. Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B 50,
1795317979 (1994).
77. Kresse, G. & Joubert, D. From ultrasoft pseudopotentials to the projector
augmented-wave method. Phys. Rev. B 59, 17581775 (1999).
78. Dudarev, S. L., Botton, G. A., Savrasov, S. Y., Humphreys, C. J. & Sutton, A. P.
Electron-energy-loss spectra and the structural stability of nickel oxide. Phys. Rev.
B57, 15051509 (1998).
79. Dorado, B. & Garcia, P. First-principles DFT+U modeling of actinide-based alloys:
application to paramagnetic phases of UO2 and (U,Pu) mixed oxides. Phys. Rev. B
87, 195139 (2013).
This work is licensed under a Creative Commons Attribution 4.0
International License. The images or other third party material in this
article are included in the articles Creative Commons license, unless indicated
otherwise in the credit line; if the material is not included under the Creative Commons
license, users will need to obtain permission from the license holder to reproduce the
material. To view a copy of this license, visit http://creativecommons.org/licenses/
by/4.0/
The Open Quantum Materials Database
S Kirklin et al
15
© 2015 Shanghai Institute of Ceramics, Chinese Academy of Sciences/Macmillan Publishers Limited npj Computational Materials (2015) 15010
... Compositional space within the Ba-As binary system is heavily populated with multiple reported compositions in experimental (ICSD) and theoretical (OQMD) databases: BaAs 2 , BaAs 3 , Ba 4 As 2.6 , Ba 3 As 14 , Ba 5 As 4 , Ba 3 As 2 , BaAs, Ba 5 As 3 , and Ba 3 As 4 [34,35,[38][39][40][41][42] ( Figure 1). Note that the title Ba 3 As 4 and Ba 5 As 4 were only theoretically predicted, being accessible from OQMD, but the structural elucidation of both will be presented in this work. ...
... Compositional space within the Ba-As binary system is heavily populated with multiple reported compositions in experimental (ICSD) and theoretical (OQMD) databases: BaAs2, BaAs3, Ba4As2.6, Ba3As14, Ba5As4, Ba3As2, BaAs, Ba5As3, and Ba3As4 [34,35,[38][39][40][41][42] (Figure 1). Note that the title Ba3As4 and Ba5As4 were only theoretically predicted, being accessible from OQMD, but the structural elucidation of both will be presented in this work. ...
... Note that the title Ba3As4 and Ba5As4 were only theoretically predicted, being accessible from OQMD, but the structural elucidation of both will be presented in this work. [35][36][37]41,42]. New phases reported in this paper are highlighted in red. ...
Article
Full-text available
Three novel binary barium arsenides, Ba3As4, Ba5As4, and Ba16As11, were synthesized and their crystal and electronic structures were investigated. Structural data collected via the single-crystal X-ray diffraction method indicate that the anionic substructures of all three novel compounds are composed of structural motifs based on the homoatomic As–As contacts, with [As2]4− dimers found in Ba5As4 and Ba16As11, and an [As4]6− tetramer found in Ba3As4. Ba3As4 and Ba5As4 crystallize in the orthorhombic crystal system—with the non-centrosymmetric space group Fdd2 (a = 15.3680(20) Å, b = 18.7550(30) Å, c = 6.2816(10) Å) for the former, and the centrosymmetric space group Cmce (a = 16.8820(30) Å, b = 8.5391(16) Å, and c = 8.6127(16) Å) for the latter—adopting Eu3As4 and Eu5As4 structure types, respectively. The heavily disordered Ba16As11 structure was solved in the tetragonal crystal system with the space group P4¯21m (a = 12.8944(12) Å and c = 11.8141(17) Å). The Zintl concept can be applied to each of these materials as follows: Ba3As4 = (Ba2+)3[As4]6−, Ba5As4 = (Ba2+)5(As3−)2[As2]4−, and 2 × Ba16As11 = (Ba2+)32(As3−) ≈ 20[As2]4− ≈ 1, pointing to the charge-balanced nature of these compounds. Electronic structure calculations indicate narrow bandgap semiconducting behavior, with calculated bandgaps of 0.47 eV for Ba3As4, 0.34 eV for Ba5As4, and 0.33 eV for Ba16As11.
... In another study, Zhang et al. 24 introduced an information entropy-based metric to measure data bias and developed an entropy-targeted active learning (ET-AL) framework to mitigate it. They utilized the ET-AL framework to guide new data acquisition in density functional theory (DFT)-generated materials databases (such as OQMD 28,29 and JARVIS 30 ) by addressing the imbalanced coverage of formation energy among different crystal systems (i.e., structure-stability bias). This approach aimed to enhance the diversity of underrepresented crystal systems and, as a result, improved the performance of machine learning models 24 . ...
Preprint
Full-text available
Machine learning models are susceptible to being misled by biases in training data that emphasize incidental correlations over the intended learning task. In this study, we demonstrate the impact of data bias on the performance of a machine learning model designed to predict the synthesizability likelihood of crystal compounds. The model performs a binary classification on labeled crystal samples. Despite using the same architecture for the machine learning model, we showcase how the model's learning and prediction behavior differs once trained on distinct data. We use two data sets for illustration: a mixed-source data set that integrates experimental and computational crystal samples and a single-source data set consisting of data exclusively from one computational database. We present simple procedures to detect data bias and to evaluate its effect on the model's performance and generalization. This study reveals how inconsistent, unbalanced data can propagate bias, undermining real-world applicability even for advanced machine learning techniques.
... Data-driven materials science has been a rapidly developing field facilitating efficient analysis and use of robust datasets for accelerated design of new materials [38]. Materials database platforms such as the Materials Project (MP) [39], the Open Quantum Materials Database (OQMD) [40], and the Automatic-FLOW for materials discovery (AFLOW) [41], collect DFT data including formation enthalpy, elastic and thermal properties that can be used to train machine learning (ML) models with various compositional-feature-descriptor extraction strategies. Examples of ML models able to predict formation enthalpies for yet unknown compounds are given in Refs. ...
Article
Full-text available
Data scarcity is one of the critical bottlenecks to utilizing machine learning in material discovery. Transfer learning can use existing big data to assist property prediction on small data sets, but the premise is that there must be a strong correlation between large and small data sets. To extend its applicability in scenarios with different properties and materials, here we develop a hybrid framework combining adversarial transfer learning and expert knowledge, which enables the direct prediction of carrier mobility of two-dimensional (2D) materials using the knowledge learned from bulk effective mass. Specifically, adversarial training ensures that only common knowledge between bulk and 2D materials is extracted while expert knowledge is incorporated to further improve the prediction accuracy and generalizability. Successfully, 2D carrier mobilities are predicted with the accuracy over 90% from only crystal structure, and 21 2D semiconductors with carrier mobilities far exceeding silicon and suitable bandgap are successfully screened out. This work enables transfer learning in simultaneous cross-property and cross-material scenarios, providing an effective tool to predict intricate material properties with limited data.
Article
Full-text available
With extraordinarily high efficiency, low cost, and excellent stability, 2D perovskite has demonstrated a great potential to revolutionize photovoltaics technology. However, inefficient material structure representations have significantly hindered artificial intelligence (AI)-based perovskite design and discovery. Here we propose geometric data analysis (GDA)-based perovskite structure representation and featurization and combine them with learning models for 2D perovskite design. Both geometric properties and periodicity information of the material unit cell, are fully characterized by a series of 1D functions, i.e., density fingerprints (DFs), which are mathematically guaranteed to be invariant under different unit cell representations and stable to structure perturbations. Element-specific DFs, which are based on different site combinations and atom types, are combined with gradient boosting tree (GBT) model. It has been found that our GDA-based learning models can outperform all existing models, as far as we know, on the widely used new materials for solar energetics (NMSE) databank.
Preprint
Full-text available
Overcoming the challenge of limited data availability within materials science is crucial for the broad-based applicability of machine learning within materials science. One pathway to overcome this limited data availability is to use the framework of transfer learning (TL), where a pre-trained (PT) machine learning model (on a larger dataset) can be fine-tuned (FT) on a target (typically smaller) dataset. Our study systematically explores the effectiveness of various PT/FT strategies to learn and predict material properties with limited data. Specifically, we leverage graph neural networks (GNNs) to PT/FT on seven diverse curated materials datasets, encompassing sizes ranging from 941 to 132,752 datapoints. We consider datasets that cover a spectrum of material properties, ranging from band gaps (electronic) to formation energies (thermodynamic) and shear moduli (mechanical). We study the influence of PT and FT dataset sizes, strategies that can be employed for FT, and other hyperparameters on pair-wise TL among the datasets considered. We find our pair-wise PT-FT models to consistently outperform models trained from scratch on the target datasets. Importantly, we develop a GNN framework that is simultaneously PT on multiple properties (MPT), enabling the construction of generalized GNN models. Our MPT models outperform pair-wise PT-FT models on several datasets considered, and more significantly, on a 2D material band gap dataset that is completely out-of-distribution from the PT datasets. Finally, we expect our PT/FT and MPT frameworks to be generalizable to other GNNs and materials properties, which can accelerate materials design and discovery for various applications.
Article
Full-text available
Accelerating the discovery of advanced materials is essential for human welfare and sustainable, clean energy. In this paper, we introduce the Materials Project (www.materialsproject.org), a core program of the Materials Genome Initiative that uses high-throughput computing to uncover the properties of all known inorganic materials. This open dataset can be accessed through multiple channels for both interactive exploration and data mining. The Materials Project also seeks to create open-source platforms for developing robust, sophisticated materials analyses. Future efforts will enable users to perform ‘‘rapid-prototyping’’ of new materials in silico, and provide researchers with new avenues for cost-effective, data-driven materials design.
Article
Full-text available
The standard enthalpy of formation of InN at 298 K has been determined using high-temperature oxidative drop solution calorimetry in a molten sodium molybdate solvent at 975 K. Calorimetric measurements were performed on six InN samples with varying nitrogen contents. The samples were characterized using x-ray diffraction, chemical analysis, electron microprobe analysis, and Brunauer-Emmett-Teller surface area measurement. The variation of the enthalpy of drop solution (kJ/g) with nitrogen content is approximately linear. The data, when extrapolated to stoichiometric InN, yield a standard enthalpy of formation from the elements of -28.6 ± 9.2 kJ/mol. The relatively large error results from the deviation of individual points from the straight line rather than uncertainties in each set of data for a given sample. This new directly measured enthalpy of formation is in good agreement with the old combustion calorimetric result by Hahn and Juza (1940). However, this calorimetric enthalpy of formation is significantly different from the enthalpy of formation values derived from the temperature dependence of the apparent decomposition pressure of nitrogen over InN. A literature survey of the enthalpies of formation of III-N nitride compounds is presented.
Article
The role of van der Waals (vdW) interactions in density functional theory (DFT) + U calculations of the layered lithium-ion battery cathode LixCoO2 (x = 0-1) is investigated using (i) dispersion corrections in the Perdew-Burke-Ernzerhof (PBE) generalized gradient approximation functional, (ii) vdW density functionals, and (iii) the Bayesian error estimation functional with vdW correlation. We find that combining vdW corrections or functionals with DFT+U can yield lithiation voltages, relative stabilities, and structural properties that are in much better agreement with experiments for the phases O1-CoO2, O3-CoO2, layered-Li0.5CoO2, spinel-Li0.5CoO2, and LiCoO2 than using DFT+U or vdW-inclusive methods alone or using the hybrid Heyd-Scuseria-Ernzerhof functional. Contributions of vdW interactions to the lithiation voltages are estimated to have a similar magnitude with that of applying a typical U in the range 2-4 eV for cobalt, each accounting for 5-10% of calculated voltages relative to PBE. Relative stabilities of O1 and O3-CoO2 as well as layered- and spinel-Li0.5CoO2 are correctly predicted with vdW-inclusive methods combined with the +U correction.
Article
The standard enthalpy of formation of InN at 298 K has been determined using high-temperature oxidative drop solution calorimetry in a molten sodium molybdate solvent at 975 K. Calorimetric measurements were performed on six InN samples with varying nitrogen contents. The samples were characterized using x-ray diffraction, chemical analysis, electron microprobe analysis, and Brunauer–Emmett–Teller surface area measurement. The variation of the enthalpy of drop solution (kJ/g) with nitrogen content is approximately linear. The data, when extrapolated to stoichiometric InN, yield a standard enthalpy of formation from the elements of −28.6 ± 9.2 kJ/mol. The relatively large error results from the deviation of individual points from the straight line rather than uncertainties in each set of data for a given sample. This new directly measured enthalpy of formation is in good agreement with the old combustion calorimetric result by Hahn and Juza (1940). However, this calorimetric enthalpy of formation is significantly different from the enthalpy of formation values derived from the temperature dependence of the apparent decomposition pressure of nitrogen over InN. A literature survey of the enthalpies of formation of III–N nitride compounds is presented.
Article
We present a framework to carry out highly accurate GGA+U thermochemistry calculations by deriving effective U values from experimental data. The U values predicted in this approach are applied to metal cations, and depend not only on (i) the chemical identity and the band to which the U correction is applied, but also on the local environment of the metal described by (ii) its oxidation state and (iii) the surrounding ligand. We predict such local environment dependent (LD) U values for the common oxidation states of 3d metals M = Ti, V, Cr, Mn, Fe, Co, and Ni in their oxides and fluorides. We implement the GGA/GGA+U mixing method [Jain et al. Phys. Rev. B 84, 045115 (2011)] to establish the total energy compatibility among the GGA+U calculations involving M treated with different LD-U values. Using the presented framework, formation enthalpies of 52 transition metal bearing oxides (which are not used during the LD-U parametrization) are predicted with a remarkably small mean absolute error of similar to 19 meV/atom, which is on the order of the experimental chemical accuracy. In addition, we present applications of the method in redox processes of important 3d-metal oxide and fluoride systems such as LixCoO2, LixV6O13, LixFeF3, and VO1.5+x, and show that LD-GGA+U can overcome several drawbacks of using constant-U values in conventional GGA+U.
Article
This paper deals with the ground state of an interacting electron gas in an external potential v(r). It is proved that there exists a universal functional of the density, F[n(r)], independent of v(r), such that the expression E≡∫v(r)n(r)dr+F[n(r)] has as its minimum value the correct ground-state energy associated with v(r). The functional F[n(r)] is then discussed for two situations: (1) n(r)=n0+ñ(r), ñn0≪1, and (2) n(r)=ϕ(rr0) with ϕ arbitrary and r0→∞. In both cases F can be expressed entirely in terms of the correlation energy and linear and higher order electronic polarizabilities of a uniform electron gas. This approach also sheds some light on generalized Thomas-Fermi methods and their limitations. Some new extensions of these methods are presented.
Article
From the earlier works on the prediction of amorphous forming composition range (AFCR) using Miedema based model and also, on mechanical alloying experiments it has been observed that all amorphous forming compositions of a given alloy system falls within a linear band when the chemical contribution to enthalpy of the solid solution (ΔHss) is plotted against the enthalpy of mixing in the amorphous phase (ΔHamor). On the basis of this observation, a methodology has been proposed in this article to identify the AFCR of a ternary system that is likely to be more precise than what can be obtained using ΔHamor − ΔHss < 0 criterion. MA experiments on various compositions of Al–Ni–Ti system, producing amorphous, crystalline, and mixture of amorphous plus crystalline phases have been carried out and the phases have been characterized using X-ray diffraction and transmission electron microscopy techniques. Data from the present MA experiments and, also, from the literature have been used to validate the proposed approach. Also, the proximity of compositions, producing a mixture of amorphous and crystalline phases to the boundary of AFCR in the Al–Ni–Ti ternary has been found useful to validate the effectiveness of the prediction.
Article
Recent experiments have shown that lithium and oxygen can be electrochemically removed from Li5FeO4 (5Li2O•Fe2O3) and re-accommodated during discharge, creating the possibility of its use as a high-capacity electrode in a hybrid Li-ion/Li-O2 electrochemical cell. Taking this novel chemistry as a model, we use density functional theory (DFT) within a high-throughput framework to screen for analogous reactions in other materials. We search for candidate materials possessing high capacity, voltages compatible with existing electrolytes, and reasonable electrical conductivity. We identify several promising candidate materials that may operate by a similar reaction mechanism and are worthy of investigation, such as Li6MnO4, Li6CoO4, Li4MoO5 and Li8IrO6. This work paves the way for accelerated exploration of this intriguing new battery chemistry.
Article
We use density functional theory (DFT) in conjunction with grand canonical linear programming (GCLP), a powerful automated tool for analyzing ground state thermodynamics, to exhaustively enumerate the 515 thermodynamically stable lithiation reactions of transition metal silicides, stannides and phosphides, and compute cell potential, volume expansion, and capacity for each. These reactions comprise an exhaustive list of all possible thermodynamically stable ternary conversion reactions for these transition metal compounds. The reactions are calculated based on a library DFT energies of 291 compounds, including all transition metal silicides, phosphides and stannides found in the Inorganic Crystal Structure Database (ICSD). We screen our computational database for the most appealing anode properties based on gravimetric capacity, volumetric capacity, cell potential, and volume expansion when compared with graphitic carbon anodes. This high-throughput computational approach points towards several promising anode compositions with properties significantly superior to graphitic carbon, including CoSi2, TiP and NiSi2.