Inverse Band Structure Design via Materials Database Screening

4 days ago - ... Screening: Application to Square Planar Thermoelectrics. Eric B. Isaacs and Chris Wolverton. Department of Materials Science and Engi...
1 downloads 10 Views 784KB Size
Subscriber access provided by UNIV OF SCIENCES PHILADELPHIA

Article

Inverse band structure design via materials database screening: application to square planar thermoelectrics Eric B Isaacs, and Chris Wolverton Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.7b04496 • Publication Date (Web): 26 Feb 2018 Downloaded from http://pubs.acs.org on February 26, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Chemistry of Materials is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 12 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Inverse band structure design via materials database screening: application to square planar thermoelectrics Eric B. Isaacs and Chris Wolverton∗ Department of Materials Science and Engineering, Northwestern University, Evanston, Illinois 60208, United States E-mail: [email protected]

Abstract

promising thermoelectric power factor behavior for the compounds. This methodology is easily adapted to other targeted band structures and should be widely applicable to a variety of design problems.

Electronic band structure contains a wealth of information on the electronic properties of a solid and is routinely computed. However, the more difficult problem of designing a solid with a desired band structure is an outstanding challenge. In order to address this inverse band structure design problem, we devise an approach using materials database screening with materials attributes based on the constituent elements, nominal electron count, crystal structure, and thermodynamics. Our strategy is tested in the context of thermoelectric materials, for which a targeted band structure containing both flat and dispersive components with respect to crystal momentum is highly desirable. We screen for thermodynamically stable or metastable compounds containing d8 transition metals coordinated by anions in a square planar geometry in order to mimic the properties of recently identified oxide thermoelectrics with such a band structure. In doing so, we identify 157 compounds out of a total of over half a million candidates. After further screening based on electronic band gap and structural anisotropy, we explicitly compute the band structures for the several of the candidates in order to validate the approach. We successfully find two new oxide systems that achieve the targeted band structure. Electronic transport calculations on these two compounds, Ba2 PdO3 and La4 PdO7 , confirm

Introduction Electronic band structure, the energy  as a function of crystal momentum k for each electron band, encapsulates a wealth of fundamental information on the electronic properties of a crystalline material. Simple examples include the nature and magnitude of the fundamental band gap and the carrier effective masses (for a semiconductor) and the Fermi surface (for a metal). Band structure impacts virtually all aspects of material properties, including electronic, thermal, optical, magnetic, and mechanical behavior. It has now become a routine task to compute the band structure for a given crystal structure, in particular using Kohn-Sham density functional theory (DFT). 1,2 However, the inverse problem of finding or designing a compound with a targeted band structure is a much more difficult challenge. Franceschetti and Zunger devised an approach to this inverse problem relying on the rapid, approximate evaluation of the forward problem. 3 In this work, an optimization algorithm (simulated annealing) was used to achieve a targeted band struc-

ACS Paragon Plus Environment

1

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ture property (maximum band gap) within the configurational space of a specific superlattice or alloy system. Variants of this approach involve other optimization algorithms (e.g., genetic algorithms, particle swarm optimization), other targeted properties (e.g., minimum band gap, direct band gap, large optical transition oscillator strength), and expanded structural spaces. 4–6 In this work, we take a different approach to the inverse band structure design problem via a materials database screening based on several materials attributes. Whereas the search space in the approach of Franceschetti and Zunger is all the possible substitutional decorations of a given structure type for a particular chemistry, we search over different stoichiometries, chemistries, and a wide variety of known crystal structures. Our chemical and structural space is the more than half a million inorganic crystalline solids in the Open Quantum Materials Database (OQMD), 7,8 a database of electronic structure calculations. In addition to ∼ 38, 000 9 of the known crystalline solids from the Inorganic Crystal Structure Database (ICSD), 10,11 the OQMD currently contains electronic structure calculations of ∼ 493, 000 hypothetical crystalline compounds based on decorating known binary and ternary prototype crystal structures. 12 Our approach takes advantage of this existing, extensive collection of materials data. Though this database contains the electronic structures of all compounds, the band structures along high-symmetry directions in k-space are not computed or stored in the database. In other words, we cannot simply “search” the database directly for the targeted band structure. The materials attributes that we use in our design/screening strategy include (1) the elements contained in the material, (2) nominal valence electron count for each element, (3) the local coordination geometry of atoms in the crystal structure, and (4) thermodynamic stability. For example, here we focus (for reasons described below) on (1) compounds containing transition metals and anions, (2) transition metals in a d8 electronic configuration, (3) transition metals coordinated by anions in

Page 2 of 12

a square planar arrangement, and (4) compounds that are thermodynamically stable or metastable. We design the targeted band structure by querying for any crystal that simultaneously possesses all the specified materials attributes. In our approach, full computation of the band structure (the forward problem) is only necessary for a small number of candidate compounds as a validation. This is highly advantageous since band structures require additional DFT calculations and a quantitative figure of merit for a candidate band structure is not always easily computed. (a)

(b) Transition Metal

Anion

B1g x2-y2 B2g xy A1g 3z2-r2 Eg

xz, yz

Figure 1: (a) Geometry and (b) corresponding crystal field splitting of d orbitals of a transition metal in square planar coordination (D4h point group). The d8 electronic configuration is shown as an example. Note that the order of the A1g and B2g levels is sometimes reversed, as in the case of Bi2 PdO4 in Ref. 13. We test our design strategy in the context of thermoelectric materials, which can enable waste heat capture via the conversion of a temperature gradient into a electrical current. 14 The thermoelectric figure of merit is proportional to the power factor σS 2 , where σ is the electronic conductivity and S is the Seebeck coefficient. Since the power factor is strongly dependent on band structure, thermoelectrics provides an excellent context to study band structure design. In particular, as discussed below, a band structure containing both flat and dispersive parts with respect to k leads to large power factor. In our case, the materials attributes specified above related to constituent elements, nominal electron count, and structure (d8 transition metals coordinated by anions in square planar geometry, shown in Fig. 1) are chosen to mimic those of Bi2 PdO4 13 and

ACS Paragon Plus Environment

2

Page 3 of 12 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

PbPdO2 , 15,16 two oxide thermoelectrics which were recently found to exhibit such a band structure. Screening the entire OQMD, we find 157 candidates for this targeted band structure. After further screening based on electronic band gap and structural anisotropy, we compute the band structures for several compounds to verify our approach. We identify two existing oxide compounds (Ba2 PdO3 and La4 PdO7 ) that successfully achieve the targeted band structure, providing validation to our inverse band structure design approach. Additionally, electronic transport calculations indicate promising power factor behavior in these compounds, particularly for n-type doping

crystal structures in particular can lead to a large power factor due to the enhanced density of states at the band edge. 24 A previous computational study proposed Bi2 PdO4 as a possibly promising thermoelectric oxide due in part to large power factor. 13 A key component to the promising electrical properties was the pudding-mold band structure, which relates to the square planar coordination (illustrated in Fig. 1(a)) of d8 Pd by oxygen. As shown in Fig. 1(b), a d8 (as well as d4 or d6 , in principle) electronic configuration of the metallic element in this coordination can lead to a semiconducting compound according to crystal field theory, 25 as is the case for Bi2 PdO4 . In this particular compound, the stacked arrangement of square planar PdO4 units with d3z2 −r2 highest-occupied orbitals (whose z axis is aligned with the stacking direction) leads to a valence band that is dispersive in the stacking direction and flat in the other directions (a quasi-one-dimensional crystal structure). Although it has a distinct structure, PbPdO2 also has a d8 Pd square planar coordination, pudding-mold-like band structure, and promising thermoelectric properties, 15,16 which is suggestive of a connection between square planar coordination and pudding-mold band structure. We note that Co-doped PbPdO2 has also been studied as a possible spin-gapless semiconductor. 26–28 Therefore, we choose our materials attributes to mimic those of Bi2 PdO4 and PbPdO2 and screen the OQMD based on the following criteria:

Computational Details Using a Boltzmann transport approach, Kuroki and Arita showed that a band with both flat (small ∇k ) and dispersive (large ∇k ) parts along a direction in k-space can lead to high power factor if the electronic chemical potential µ lies at an energy separating the two. 17 For this special “pudding-mold” band structure, the band velocity difference across µ (proportional to S) is large, and the high band velocity of the dispersive part and large Fermi surface enable a large σ. In addition to leading to the exceptionally high power factor of Nax CoO2 , 17 the pudding-mold band structure has been suggested to contribute to the record-breaking thermoelectric performance of SnSe 18–20 and to the large computed power factors of recently proposed Fe-based Heusler compounds. 21 While the original pudding-mold band structure concept corresponds to regions of small and large ∇k  along the same direction in kspace, one can consider a broader definition in which the flat and dispersive regions can occur along in different directions in k-space. Such a band structure naturally emerges from a low-dimensional (e.g. one- or two-dimensional) crystal structure, which can be thought of as analogous to the low-dimensional nanostructures proposed by Hicks and Dresselhaus to exhibit high thermoelectric performance. 22,23 Usui and Kuroki showed that one-dimensional

• Chemistry – Compound must contain TM and anion elements, and (for practicality) no radioactive elements. • Structure – In the compound, TMs must be coordinated by anions in square planar coordination (4 nearest neighbors and bond angles 4 × 85 − 95◦ , 2× ≥ 175◦ ). 29 • Electron count – All TM in the compound must have d4 , d6 , or d8 configuration. As shown in Fig. 1(b), for a TM in square planar coordination, such nominal

ACS Paragon Plus Environment

3

Chemistry of Materials

Results and Discussion

electron counts can lead to a semiconductor based on crystal field theory.

Characterization of identified compounds

• Thermodynamics – Compound must be thermodynamically stable, metastable (within 25 meV/atom of the convex hull), or unstable but in ICSD. This attribute relates the desired synthesizability of the compound, rather than the band structure.

6

Total Stable Metastable

20802 35188

63198

10000

16907

3347

714 800

1000

Unstable

1833

100000

431634

7208

157 108

100

25 24

Since the OQMD and the associated qmpy framework 30 do not intrinsically compute or store the coordination environments of different atoms in a crystal structure, we wrote a separate code to enable the assessment of our structural criterion. To query for a particular coordination environment, this code first computes all the distances between TMs and anions (taking into account translation vectors) to check if the coordination number is a desired value, e.g. 4 for square planar. If so, the bond lengths and bond angles for the nearestneighbor shell are computed to enable any local structural query. Details on the parameters we employ in this code, including how the coordination number is computed, are included in the Supporting Information. Additional DFT calculations are performed for the most promising candidate materials using vasp. 31–34 The generalized gradient approximation 35 is employed and the KohnSham equations are solved using a 500 eV plane wave basis with the projector augmented wave (PAW) method 36,37 and uniform k-point meshes of k-point density ≥ 700/Å−3 . The total energy and ionic forces are converged to within 10−6 eV and 0.001 eV/Å, respectively. Electronic band structures are computed for high-symmetry k-paths based on the conventions of Setwayan et al. 38 Semiclassical electronic transport calculations within the constant relaxation time approximation are performed in boltztrap with k-point density ≥ 2300/Å−3 . 39

531122

9384 4067 3456

1×10

Number of Compounds

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 12

10 1

Total

Non− Stability Chemistry d4, 6, or 8 radioactive

Square planar

Screening

Figure 2: Total number of compounds (thick bars) as a function of consecutive screening filters on a logarithmic scale, with breakdown in terms of thermodynamic categories (thin bars). Unstable compounds are only included if they are reported in ICSD. Figure 2 summarizes the number of compounds found in the OQMD to satisfy the screening criteria. As different criteria are applied successively, the number of candidates decreases from the total of over half a million to just 157. The largest fractional decreases in candidate compounds come from the thermodynamics criterion (reduction by factor of 7) and the square planar structural criterion (reduction by factor of 21). The vast majority of the compounds identified (150/157) turn out to be present in ICSD (including stable and metastable compounds). 68 of the these compounds have one of 32 structural prototypes listed in the ICSD, which are included in the Supporting Information. The structural prototypes corresponding to the most compounds are KBrF4 type (9, e.g. PbPdF4 ), La2 Sb-type (8, e.g. GdZrSb), Rb3 PdF5 -type (7, e.g. BaLa2 PtO5 ), and K2 PtCl4 -type (5, e.g. Tl2 PdCl4 ).

ACS Paragon Plus Environment

4

160 Number of Compounds

(a)

Other Halide Oxide

31

140 120 100

16

56

80

46

60 40

70 46

20

9 14

0

2

13 1 10

le

e

bl ta

le

b ta ns

U

ab St

l

as et M

The lack of non-ICSD compounds found might be explained by the fact that few of the structural prototypes used to generate hypothetical compounds in the OQMD contain square planar bonding. Only two of the prototype structures used to systematically and comprehensively (i.e., considering nearly all periodic table elements) generate hypothetical compounds contain any square planar coordination. The only one currently in the OQMD in which all sites for an element are in square planar coordination is the L12 structure (e.g. Cu in CuAu3 ). The other, the D022 prototype, has square planar coordination for 1/3 of the sites of one of the elements (e.g. Al in Al3 Ti). The 7 non-ICSD compounds found by our query, (Ce/Pr/Nd/Sm/Gd/Tb/Dy)2 PdO4 , actually correspond to neither the L12 nor the D022 prototype. These compounds have the distinct layered Ca2 RuO4 prototype structure, which was previously explored for a very small number (∼ 25) of Pd oxide compounds to search for and assess the stability of materials similar to Bi2 PdO4 . 13 The chemistry of the 157 identified compounds is summarized in Fig. 3(a). Oxides are the most prevalent chemistry, with 70 compounds. Note that we include compounds with polyatomic anions in this category if the oxygen of the polyatomic anion is bonded to the transition metal. Examples include iodates, nitrates, sulfates, selenates, and pyrophosphates. There are also a significant number of halides (56). The remaining 31 compounds include sulfides, selenides, phosphides, nitrides, carbonyls, and cyanides. Our query is able to find arbitrarily complex stoichiometries and crystal structures, as exemplified by H8 CrSiO4 F6 . Each Cr is bonded to 2 hydroxide groups and 2 F in this compound, which we choose to count as an oxide in Fig. 3(a). The crystal structure also contains H2 , H2 O2 , and SiF6 units. Fig. 3(b) illustrates the particular elements involved in the square planar bonding. Pd oxides are dominant with 37 occurrences, and Pd halides also occur significantly (29 times) in the group of compounds. This is consistent with the prevalence of Pd2+ in square planar environments as has been found previously. 40 Au

Al

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Br

(b)

Fe

C

F

I

Rh Pd

Mn

Ru O

Pt

N Ni P Zr S Cr Sc

Se

(c)

Cu

Au

Sb

No. of Compounds

Page 5 of 12

Cl

Ag

Ge

35 30 25 20 15 10 5 0 0

1

2

3

4

5

DFT Band Gap (eV)

Figure 3: Characterization of the 157 candidate compounds identified. (a) Distribution of stabilities split up in terms of oxides, halides, and other chemistries, (b) Network of chemical bonding of the square planar environment in which node size represents elemental occurrence and edge thickness is proportional to bonding frequency, (c) Histogram of DFT band gaps (greyed out for values larger than 1.1 eV).

ACS Paragon Plus Environment

5

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

oxides and chlorides also feature prominently. Despite the dominance of Pd oxides, there is still overall a diverse range of transition metals and anions in the 157 compounds. We also observe significant diversity in structures. There are compounds of all dimensionalities (e.g. 0D PdCl2 , 1D PdBr2 , 2D PdS2 , 3D Mn3 Sb). Here dimensionality refers to the connectivity of the square planar units. For example, in PdBr2 there are 1D channels of edge-sharing PdBr4 square planar units. Even within a particular dimensionality of the square planar connectivity, there is significant diversity in structure. For example, for 0D (unconnected) square planar compounds, there are distinct relative orientations of the square planar units (e.g. perpendicular in PbPdF4 vs. parallel in Tl2 PdCl4 ).

Page 6 of 12

NaBi2 AuO5 , Ba2 TlCuO5 , Ba2 PdO3 , LiCuO2 , and La4 PdO7 . We focus on such “non-isotropic” compounds since the difference in bonding in different directions should lead in principle to different ∇k  in different directions of k-space as is targeted for the pudding-mold band structure.

Compounds achieving the targeted band structure Two of the compounds for which the band structure is computed, Ba2 PdO3 and La4 PdO7 , successfully achieve the targeted band structure. Ba2 PdO3 (space group Immm) and La4 PdO7 (space group C2/m) both contain one-dimensional chains of corner-sharing PdO4 square planar units (shown going into the plane for La4 PdO7 in Fig. 4). For both of these materials, the number of d electrons (roughly estimated via projection within the PAW sphere) on the transition metal (Pd) site is 8.4, consistent with the nominal d8 configuration. As shown in Fig. 4, Ba2 PdO3 and La4 PdO7 possess the pudding-mold band structure. Both show flat and dispersive parts of the low-energy spectrum both below and above the Fermi energy. For example, for Ba2 PdO3 the conduction band disperses strongly (weakly) along Γ– Y (Γ–Z). The valence band disperses strongly (weakly) along W–R (W–T). La4 PdO7 similarly shows both dispersive and flat components, though it is seen most strongly in the conduction band. We note that La4 PdO7 appears to have a smaller valence bandwidth than Ba2 PdO3 as the highest-occupied band is flat for much of the k-path shown in Fig. 4(b). The density of states in Fig. 4 shows strong Pd d character in the band edges. The fact that we find the targeted band structure in these two compounds is a strong success of our inverse design strategy. Although the square planar unit itself is twodimensional, the dimensionality of the band structure for the extended system depends strongly on the connectivity of the square planar units. For example, one-dimensional chains of PdO4 square planar units (e.g. stacked in the case of Bi2 PdO4 , corner-sharing in the case

Further screening of oxides based on band gap The band gap, in addition to the dispersion of the bands, is critically important for thermoelectric materials. 41 In particular, highefficiency thermoelectric materials such as PbTe, SnSe, and Bi2 Te3 have band gaps no larger than around 1 eV. As such, to further screen the 157 candidate compounds, we further require a DFT band gap no larger than 1.1 eV. We choose not to put a lower limit on the band gap since semilocal DFT is well known to underestimate the electronic band gap. 42 As illustrated in the band gap distribution in Fig. 3(c), this leaves 72 remaining compounds. We focus on the 33 of these compounds that are pure oxides (i.e., containing no other anion element). These candidate compounds are listed in the Supporting Information. The known square planar thermoelectric candidate materials Bi2 PdO4 and PbPdO2 are included in these 33, which is a validation of the screening strategy. One of the 33 compounds is CaPd3 O4 , which has been studied as a possible excitonic insulator 43,44 and topological Dirac semimetal. 45 The electronic band structures are computed for five of the compounds in which the connectivity of the square planar units is not the same in all spatial directions:

ACS Paragon Plus Environment

6

Page 7 of 12

Ba2PdO3

La4PdO7

(a) 2

(b)

Total Pd d

1

Op

0 Energy (eV)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

−1 −2 −3 −4 −5 −6 Γ

X

LT W

R X1

Z

Γ

YS

W

Γ

YF

L

I

Figure 4: Electronic band structure and total and projected density of states of (a) Ba2 PdO3 and (b) La4 PdO7 . The dashed red line marks the valence band maximum. In the crystal structures shown in the insets, Pd, O, and Ba/La atoms are shown in blue, red, and green, respectively. primarily corresponding to the TM d orbitals, though it does contain some pudding-mold-like features.

of Ba2 PdO3 ) lead to a one-dimensional band structure with dispersion primarily in one direction, whereas a two-dimensional layer of square planar PdO4 units (e.g. La2 PdO4 ) leads to a two-dimensional band structure with dispersion primarily in two directions. This suggests our type of inverse band structure design for pudding-mold thermoelectrics might also have success for other coordinations, in addition to square planar, that lend themselves to lowdimensional crystal structures. We note that our design strategy did not uncover any candidate compounds with a pudding-mold band structure that contain a lone pair cation, which would also be desirable for a thermoelectric by lowering lattice thermal conductivity. 46,47 There are several compounds with Zr bonded to Sb in square planar coordination (e.g. GdZrSb), but here Sb is nominally the anion. In nearly all of the other compounds identified, either a metal is found (e.g. Ba2 TlCuO5 , Li8 Bi2 PdO10 ) or the electronic band gap is too large (e.g. 1.4 eV for Pb2 PdCl6 , 2.3 eV for Tl2 Ni(CN)4 ). For NaBi2 AuO5 , the valence band edge does not

Electronic transport properties of Ba2 PdO3 and La4 PdO7 Finally, as a confirmation that the designed electronic band structure leads to promising thermoelectric behavior, we perform electronic transport calculations via the constant relaxation time approximation. To isolate the effect of band structure shape, rather than the magnitude of the band gap, we compare the transport properties of Ba2 PdO3 and La4 PdO7 to Bi2 PdO4 at a fixed band gap of 1.41 eV. This value corresponds to the band gap of Bi2 PdO4 from hybrid DFT calculations and was used in Ref. 13. Figure 5(a) shows the behavior of the power factor (divided by the relaxation time) for these 3 materials as a function of doping at 700 K. Typical doping magnitudes leading to peak thermoelectric performance that can be experimentally achieved are 1019 to 1021 cm−3 . 48

ACS Paragon Plus Environment

7

Chemistry of Materials

(a)

For p-type doping, the power factor magnitude of Ba2 PdO3 and La4 PdO7 is similar to that of Bi2 PdO4 (∼ 1012 W/(m K2 s)). Bi2 PdO4 achieves a peak in S 2 σ/τ of 1.36 ×1012 W/(m K2 s) for a carrier concentration of 5.3 × 1020 cm−3 . One of the new compounds (Ba2 PdO3 ) achieves a higher peak value at a higher doping (3.5 × 1021 cm−3 ), but for more reasonable dopings (< 1021 cm−3 ), the value for Bi2 PdO4 is larger than those of Ba2 PdO3 and La4 PdO7 by roughly 50%. In contrast, for n-type doping Ba2 PdO3 and La4 PdO7 exhibit larger power factor behavior. Ba2 PdO3 has the largest peak value of 1.48 ×1012 W/(m K2 s) at doping of 4.4 × 1020 cm−3 . The improved n-type behavior of the new compounds can be explained by the presence of pudding-mold band structure in the conduction band, which is lacking in Bi2 PdO4 , 13 in addition to the valence band. The corresponding plots of power factor versus temperature for a fixed doping of 1020 cm−3 shown in Fig. 5(b) also illustrate the superior n-type performance for the new compounds as compared to Bi2 PdO4 . Such transport calculations confirm that the pudding-mold band structures achieved lead to significant promise for thermoelectric applications, for p- and especially n-type doping. We performed hybrid DFT calculations (using HSE 49 ) of Ba2 PdO3 and La4 PdO7 exhibit band gaps of ≈ 1.5 − 1.6 eV (close to the 1.41 eV of Bi2 PdO4 ). Therefore, we believe incorporating band gap differences between the three materials to not significantly affect the comparison in Fig. 5.

Temperature: 700 K 12

1.4×10

2

2

S σ/τ [W/(m K s)]

12

1.2×10

12

1×10

11

8×10

11

6×10

11

Bi2PdO4 Ba2PdO3 La4PdO7

4×10

11

2×10

0 21 −1×10

20

n (cm ) Doping: 10

12

1.4×10

12

2

5×10 −3

20

(b)

2

20

0

−5×10 n−type

S σ/τ [W/(m K s)]

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

21

1×10

p−type −3

cm

Thin lines: p−type Thick lines: n−type

1.2×10

12

1×10

11

8×10

11

6×10

11

4×10

11

2×10

0 300

400

500

600 700 T (K)

800

900

Page 8 of 12

1000

Figure 5: Power factor σS 2 divided by relaxation time τ for Ba2 PdO3 and La4 PdO7 as compared to that of Bi2 PdO4 . (a) σS 2 /τ as a function of doping at 700 K, (b) σS 2 /τ as a function of temperature for p and n-type doping of 1020 cm−3 . To isolate the effect of band structure shape, for all cases the band gap is scissor shifted to the HSE band gap of Bi2 PdO4 (1.41 eV) as in Ref. 13 .

Conclusions Using a materials database and screening criteria based on several material attributes, we have devised a method for inverse band structure design in the search space of chemistry, stoichiometry, and structure. Our method does not require the evaluation of the band structure except as a validation. As an example case, we apply the band structure design to the pudding-mold band structure desirable for thermoelectric materials with high power

ACS Paragon Plus Environment

8

Page 9 of 12 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

factor. Out of the over half a million candidate materials, 157 candidate systems are identified. After further screening based on electronic band gap and structural anisotropy, we compute the band structures for several of the candidate compounds as a validation of the approach. Two palladium oxide systems, Ba2 PdO3 and La4 PdO7 , successfully achieve the targeted band structure and are found to exhibit promising thermoelectric power factor behavior. Inverse band structure design via materials database screening represents a powerful, versatile approach for the rational identification and design of materials. This general approach opens up many possibilities for designing desirable band structures for a broad range of applications.

verse band-structure problem of finding an atomic configuration with given electronic properties. Nature 1999, 402, 60–63. (4) Kim, K.; Graf, P. A.; Jones, W. B. A genetic algorithm based inverse band structure method for semiconductor alloys. J. Comput. Phys. 2005, 208, 735–760. (5) Dudiy, S. V.; Zunger, A. Searching for Alloy Configurations with Target Physical Properties: Impurity Design via a Genetic Algorithm Inverse Band Structure Approach. Phys. Rev. Lett. 2006, 97, 046401. (6) Xiang, H. J.; Huang, B.; Kan, E.; Wei, S.H.; Gong, X. G. Towards Direct-Gap Silicon Phases by the Inverse Band Structure Design Approach. Phys. Rev. Lett. 2013, 110, 118702.

Acknowledgement We acknowledge support from the U.S. Department of Energy under Contracts DE-SC0014520 (overall design strategy and thermoelectric transport calculations) and DE-SC0015106 (development of software tools for high-throughput screening). Computational resources were provided by the Quest high performance computing facility at Northwestern University and the National Energy Research Scientific Computing Center (U.S. Department of Energy Contract DE-AC0205CH11231).

(7) Saal, J. E.; 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 2013, 65, 1501–1509. (8) Kirklin, S.; Saal, J. E.; Meredig, B.; Thompson, A.; Doak, J. W.; Aykol, M.; Rühl, S.; Wolverton, C. The Open Quantum Materials Database (OQMD): assessing the accuracy of DFT formation energies. npj Comput. Mater. 2015, 1, 15010.

Supporting Information Available: Additional details on the materials database screening criteria and full list of the 157 candidate compounds identified. This material is available free of charge via the Internet at http://pubs.acs.org/.

(9) There are ∼ 44, 000 unique ICSD compounds without partial occupancy of atomic sites in the crystal structure; ∼ 38, 000 of these with the smallest number of atoms in the primitive unit cell (. 35 atoms) have been calculated in the OQMD as of April 2017.

Notes and References (1) Hohenberg, P.; Kohn, W. Inhomogeneous Electron Gas. Phys. Rev. 1964, 136, B864–B871.

(10) Bergerhoff, G.; Hundt, R.; Sievers, R.; Brown, I. D. The inorganic crystal structure data base. J. Chem. Inf. Comput. Sci. 1983, 23, 66–69.

(2) Kohn, W.; Sham, L. J. Self-Consistent Equations Including Exchange and Correlation Effects. Phys. Rev. 1965, 140, A1133–A1138.

(11) Belsky, A.; Hellenbrandt, M.; Karen, V. L.; Luksch, P. New developments in the Inorganic Crystal

(3) Franceschetti, A.; Zunger, A. The in-

ACS Paragon Plus Environment

9

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Structure Database (ICSD): accessibility in support of materials research and design. Acta Crystallogr. Sect. B 2002, 58, 364–369.

Page 10 of 12

(20) Wang, Z.; Fan, C.; Shen, Z.; Hua, C.; Hu, Y.; Sheng, F.; Lu, Y.; Fang, H.; Qiu, Z.; Lu, J.; Xu, Z.-A.; Shen, D. W.; Zheng, Y. Defects controlled hole doping and multi-valley transport in SnSe single crystals. arXiv:1706.10054 [cond-mat] 2017,

(12) As of April 2017. (13) He, J.; Hao, S.; Xia, Y.; Naghavi, S. S.; Ozolin, š, V.; Wolverton, C. Bi2 PdO4 : A Promising Thermoelectric Oxide with High Power Factor and Low Lattice Thermal Conductivity. Chem. Mater. 2017, 29, 2529–2534.

(21) Bilc, D. I.; Hautier, G.; Waroquiers, D.; Rignanese, G.-M.; Ghosez, P. LowDimensional Transport and Large Thermoelectric Power Factors in Bulk Semiconductors by Band Engineering of Highly Directional Electronic States. Phys. Rev. Lett. 2015, 114, 136601.

(14) Goldsmid, H. J. Introduction to Thermoelectricity; Springer Series in Materials Science; Springer: Berlin, Heidelberg, 2010; Vol. 121.

(22) Hicks, L. D.; Dresselhaus, M. S. Effect of quantum-well structures on the thermoelectric figure of merit. Phys. Rev. B 1993, 47, 12727–12731.

(15) Ozawa, T. C.; Taniguchi, T.; Nagata, Y.; Noro, Y.; Naka, T.; Matsushita, A. Metal–insulator transition and large thermoelectric power of a layered palladium oxide: PbPdO2 . J. Alloys Comp. 2005, 388, 1–5.

(23) Hicks, L. D.; Dresselhaus, M. S. Thermoelectric figure of merit of a onedimensional conductor. Phys. Rev. B 1993, 47, 16631–16634.

(16) Lamontagne, L. K.; Laurita, G.; Gaultois, M. W.; Knight, M.; Ghadbeigi, L.; Sparks, T. D.; Gruner, M. E.; Pentcheva, R.; Brown, C. M.; Seshadri, R. High Thermopower with Metallic Conductivity in p-Type Li-Substituted PbPdO2 . Chem. Mater. 2016, 28, 3367–3373.

(24) Usui, H.; Kuroki, K. Enhanced power factor and reduced Lorenz number in the Wiedemann–Franz law due to pudding mold type band structures. J. Appl. Phys. 2017, 121, 165101. (25) Krishnamurthy, R.; Schaap, W. B. Computing ligand field potentials and relative energies of d orbitals: A simple general approach. J. Chem. Educ. 1969, 46, 799.

(17) Kuroki, K.; Arita, R. “Pudding Mold” Band Drives Large Thermopower in Nax CoO2 . J. Phys. Soc. Jpn. 2007, 76, 083707.

(26) Wang, X. L. Proposal for a New Class of Materials: Spin Gapless Semiconductors. Phys. Rev. Lett. 2008, 100, 156404.

(18) Zhao, L.-D.; Lo, S.-H.; Zhang, Y.; Sun, H.; Tan, G.; Uher, C.; Wolverton, C.; Dravid, V. P.; Kanatzidis, M. G. Ultralow thermal conductivity and high thermoelectric figure of merit in SnSe crystals. Nature 2014, 508, 373–377.

(27) Chen, S. W.; Huang, S. C.; Guo, G. Y.; Lee, J. M.; Chiang, S.; Chen, W. C.; Liang, Y. C.; Lu, K. T.; Chen, J. M. Gapless band structure of PbPdO2: A combined first principles calculation and experimental study. Appl. Phys. Lett. 2011, 99, 012103.

(19) Kutorasinski, K.; Wiendlocha, B.; Kaprzyk, S.; Tobola, J. Electronic structure and thermoelectric properties of n- and p-type SnSe from first-principles calculations. Phys. Rev. B 2015, 91, 205201.

(28) Kurzman, J. A.; Miao, M.-S.; Seshadri, R. Hybrid functional electronic structure of PbPdO 2 , a small-gap semiconductor.

ACS Paragon Plus Environment

10

Page 11 of 12 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

J. Phys.: 465501.

Condens. Matter 2011, 23,

(40) Waroquiers, D.; Gonze, X.; Rignanese, G.M.; Welker-Nieuwoudt, C.; Rosowski, F.; Göbel, M.; Schenk, S.; Degelmann, P.; André, R.; Glaum, R.; Hautier, G. Statistical Analysis of Coordination Environments in Oxides. Chem. Mater. 2017, 29, 8346– 8360.

(29) We do not require, however, that all the anions must be present in this bonding environment. (30) qmpy. https://github.com/ wolverton-research-group/qmpy.

(41) Sofo, J. O.; Mahan, G. D. Optimum band gap of a thermoelectric material. Phys. Rev. B 1994, 49, 4565–4570.

(31) Kresse, G.; Hafner, J. Ab initio molecular-dynamics simulation of the liquid-metal–amorphous-semiconductor transition in germanium. Phys. Rev. B 1994, 49, 14251–14269.

(42) Perdew, J. P. Density functional theory and the band gap problem. Int. J. Quantum Chem. 1985, 28, 497–523.

(32) Kresse, G.; Hafner, J. Ab initio molecular dynamics for liquid metals. Phys. Rev. B 1993, 47, 558–561.

(43) Hase, I.; Nishihara, Y. CaPd3 O4 as an excitonic insulator. Phys. Rev. B 2000, 62, 13426–13429.

(33) Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169–11186.

(44) Ichikawa, S.; Terasaki, I. Metal-insulator transition in Ca1−x Lix Pd3 O4 . Phys. Rev. B 2003, 68, 233101.

(34) Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50.

(45) Li, G.; Yan, B.; Wang, Z.; Held, K. Topological Dirac semimetal phase in Pd and Pt oxides. Phys. Rev. B 2017, 95, 035102. (46) Skoug, E. J.; Morelli, D. T. Role of Lone-Pair Electrons in Producing Minimum Thermal Conductivity in NitrogenGroup Chalcogenide Compounds. Phys. Rev. Lett. 2011, 107, 235901.

(35) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. (36) Blöchl, P. E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953– 17979.

(47) Nielsen, M. D.; Ozolin, š, V.; Heremans, J. P. Lone pair electrons minimize lattice thermal conductivity. Energy Environ. Sci. 2013, 6, 570–578.

(37) Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758–1775.

(48) Snyder, G. J.; Toberer, E. S. Complex thermoelectric materials. Nat. Mater. 2008, 7, 105–114. (49) Heyd, J.; Scuseria, G. E.; Ernzerhof, M. Hybrid functionals based on a screened Coulomb potential. J. Chem. Phys. 2003, 118, 8207–8215.

(38) Setyawan, W.; Curtarolo, S. Highthroughput electronic band structure calculations: Challenges and tools. Comput. Mater. Sci. 2010, 49, 299–312. (39) Madsen, G. K. H.; Singh, D. J. BoltzTraP. A code for calculating band-structure dependent quantities. Comput. Phys. Commun. 2006, 175, 67–71.

ACS Paragon Plus Environment

11

Chemistry of Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Graphical TOC Entry

ACS Paragon Plus Environment

12

Page 12 of 12