Sie sind auf Seite 1von 21

Changing Gap Type of Solid State Boric Acid by Heating: a Dispersion-Corrected

Density Functional Study of α-, β-, and γ-Metaboric Acid Polymorphs

M. Bezerra da Silvaa, A. M. Da Cunhab, R. C. R. Santosb, A. Valentinib,


E. W. S. Caetanoc, V. N. Freirea,b

a
Departamento de Física, Universidade Federal do Ceará, Caixa Postal 6030,
60440-900, Fortaleza-CE Brazil
b
Departamento de Química Analítica e Físico-Química, Universidade Federal do Ceará,60440-554,
Fortaleza-CE, Brazil
c
Instituto Federal de Educação, Ciência e Tecnologia do Ceará, DEMEL,
Campus Fortaleza, 60040-531, Fortaleza-CE, Brazil

Abstract

The metaboric acid α-, β-, and γ- polymorphs were investigated using density functional theory (DFT)
calculations employing the generalized gradient approximation improved by the Tkatchenko-Scheffler
scheme to take into account dispersive interactions. The structural, electronic, and optical properties of
these polymorphs were thus obtained, with unit cells deviations Δa, Δb, Δc (in comparison with X-ray
data) as small as -0.08, -0.09, - 0.06 Å for α-(BOH)3O3, 0.00, 0.06, -0.09 Å for β-(BOH)3O3, and 0.01,
0.01, 0.01 Å for γ-(BOH)3O3. The layered α-MA polymorph is predicted by our simulations to be a 6.26
eV direct gap material, which is close to the values found recently for the boric acid triclinic (H3BO3-2A,
6.25 eV) and trigonal (H3BO3-3T, 6.28 eV) indirect gap systems. As the α-MA structure can be obtained
by heating up the H3BO3-2A crystal above 100o C, we have the interesting result that a change of
temperature can modify the nature of the band gap of boric acid in solid state from indirect to direct. Van
der Waals interaction between the metaboric acid α-(BOH)3O3 planes (distant about 3.1±0.01 Å) is the
relevant aspect determining its energy gap, while its direct or indirect character depends on the crystal
structure and molecular components. On the other hand, the polymeric monoclinic β-(BOH)3O3 has a 6.56
eV indirect gap, while the covalent cubic γ-(BOH)3O3 has a 7.28 eV direct gap. Finally, the complex
dielectric function and optical absorption considering polarized light along the 001, 010, 100 crystal
planes and polycrystalline samples (POLY) of the three metaboric acid polymorphs were obtained.

1
1. INTRODUCTION

1,2
Many boron-containing compounds are being investigated due to their therapeutic effects ,
stimulating research efforts for an improved understanding of their molecular and crystal polymorph
properties, which are in general poorly known theoretically. The most conspicuous example is boric acid
H3BO3, with uses known to man for thousands of years and, particularly in recent times, being employed
in applications ranging from medicines to pesticides and industrial products. A striking example is the
preparation of effective traps/baits for the management of Aedes aegypti L. and Aedes albopictus Skuse
3,4
species of mosquitos , which are the main spreading vectors for worldwide health threats such as
5–7
malaria, dengue, and zika . The structural, electronic, and optical properties of their triclinic-2A and
hexagonal-3T polymorphs was only recently unveiled by density functional theory (DFT) calculations as
well 8. On the other hand, ab initio calculations suggested the possibility of forming boric acid clusters in
the laboratory, the hexamer form being quite stable and assuming preferentially a hexagonal-centered
rosette shape 9. It was suggested that metaboric acid can be a precursor for nanosized tubular or cage-like
10
structures with potential applications in tribology , and boric and metaboric acid molecules confined
inside a gallium-oxygen open-framework produced a system with photocatalytic activity to split water
molecules 11.

It was shown that the molecular stacking of H3BO3 polymorphs unveiled by X-rays diffraction
(AB… for H3BO3-2A with inter-planar distances d = 3.18 Å 12 and ABC… for H3BO3-3T with d = 3.19
Å 13) gives rise to slightly distinct electronic and optical properties 8. Like graphene and nongraphene 2D
materials, H3BO3-2A,3T can be recognized as layered van der Waals solids 14. The smallest Q→Γ indirect
band gap of H3BO3-2A was calculated to be 6.25 eV, which is in good agreement with the 5.98 eV value
measured for H3BO3-2A by optical absorption; H3BO3-2A has also two close indirect gaps β1→Γ and
β2→Γ at 6.34 eV and 6.28 eV, respectively, and a Γ→Γ direct gap of 6.35 eV. The smallest indirect gap
α2→Γ of the H3BO3-3T polymorph is also at 6.25 eV, but with two close gaps, Q→Γ and α1→Γ of 6.26
and 6.27 eV, respectively, and a Γ→Γ direct gap of 6.31 eV 8. The energy band gap-related optical
absorption of both boric acid polymorphs is due to transitions between valence p-oxygen states and
conduction s-hydrogen (remarkably) and s-boron orbitals. The polymorphs H3BO3-2A,3T are suggested
to be insulators due to their very flat valence and conduction bands 8.

In this work, we focus on the three (BOH)3O3 metaboric acid (MA) polymorphs 15–17 formed by
the dehydration of boric acid crystals (unit cells shown in Fig. 1). The orthorhombic metaboric acid
polymorph (α crystal) is obtained by heating boric acid crystals at 80-100 oC with water release. The
monoclinic metaboric acid polymorph (β crystal) is obtained by heating the α crystal up to 130-140 oC in

2
an almost sealed ampoule (to prevent dehydration). Finally, cubic metaboric acid (γ crystal) is obtained
by heating above 140 oC the H3BO3-2A, -3T polymorphs or the α and β systems.

Here we perform state-of-the-art DFT calculations to evaluate the structural, electronic, and optical
properties of the bulk α-, β-, and γ-metaboric acid polymorphs. The physical properties of the α crystals
is mostly molecular in origin, its structure being determined by attractive noncovalent interactions, e.g. in
plane hydrogen bonding and intra-planar van der Waals dispersion interactions between their (BOH)3O3
molecules. The improvement of computational power during the last two decades has allowed DFT-based
electronic structure modeling of molecular crystals, with several contributions on this subject being made
by our research group in the last few years 18–26.

27
Despite many advances in the description of van der Waals dispersion forces , it remains a
challenge to ascertain the best methodology 28. A sound way to improve the quality of simulations is to
employ dispersion correction schemes with minimal empirical fitting, a strategy which enhances optical
29
absorption curves in comparison with the experimental data . In this work, DFT-calculated unit cell
parameters for the MA polymorphs were obtained within the local density approximation (LDA) and the
generalized gradient approximation (GGA) in pure form and including the dispersion correction scheme
of Tkatchenko and Scheffler (TS) 30, their electronic band structures and frontier orbitals being calculated
for the first time. The complex dielectric function of each MA polymorph is presented as well.

2. MATERIALS AND METHODS

2.1. The Crystallographic Data of the Metaboric Acid Polymorphs

A X-ray diffraction study of the orthorhombic α-(BOH)3O3 metaboric acid polymorph (see Fig.
1(a)) was published by Peters and Milbeg 15. The structure is formed by layers of planar B3O6 motifs (see
Fig. 1(b)) made from three BO3 triangles (Fig. 1(c)), each one sharing two oxygen atoms, so as to form a
hexagonal B3O3 ring. Hydrogen bonds occur between the motifs in the same layer, while the interlayer
distance is 3.1 Å (see Fig. 1(b)), which accounts for the mica-like cleavage of its colorless rhombic crystal
plates. The α-(BOH)3O3 metaboric acid polymorph space group is Pbnm with Z = 12, lattice parameters
a = 8.02 Å, b = 9.70 Å, and c = 6.13 Å 15. The α-(BOH)3O3 atomic coordinates (experimental) are given
in the Table ST1 of the Supporting Information.

3
Figure 1. (a) The α-(BOH)3O3 metaboric acid polymorph unit cell; (b) two interacting planes of the α-
(BOH)3O3 metaboric acid crystal; (c) the ab plane of the α-(BOH)3O3 metaboric boric polymorph showing
the in plane hexagonal arrangement; (d) the β-(BOH)3O3 metaboric acid polymorph unit cell; (d) unit cell
of the β-(BOH)3O3 polymorph; (e) the polymeric chain of coupled metaboric acid molecules that gives
rise to the β-(BOH)3O3 polymorph crystal; (f) interactions between two polymeric chains of the β-
(BOH)3O3 polymorph; (g) the γ-(BOH)3O3 metaboric acid polymorph unit cell; (h) the structural element
of the γ-(BOH)3O3 polymorph and its neighborhood coupling; (i) the tetrahedrally coordinated structure
representation of the γ-(BOH)3O3 metaboric acid polymorph.

4
The crystal structure of monoclinic β-metaboric acid polymorph was published originally by
16 31
Zachariasen . A precise redetermination was performed by Freyhardt, Wiebcke, and Felsche . As
depicted in Fig. 1(d)-(f), the β-MA unit cell contains two structures forming chains linked through intra-
layer H3…O3 hydrogen bonds between exocyclic O6-H3 hydroxyl groups and endocyclic O3 atoms. On
the other hand, the six-membered rings are linked through their O4 atoms into polymeric zigzag chains
extended along the [010] direction, which in turn are arranged in layers parallel to (102) planes. The β-
(BOH)3O3 crystal structure has space group P21/a, Z = 12 and unit cell dimensions a = 7.12 Å, b = 8.84
Å, c = 6.77 Å, and β = 93.26o. The β-(BOH)3O3 atomic coordinates are shown in Table ST2 (see
Supporting Information).

Lastly, the crystal structure of the covalent cubic γ-metaboric acid polymorph, (see Fig. 1(g)) was
17
published originally by Zachariasen and followed by a precise redetermination done by Freyhardt,
Wiebcke, and Felsche 31. In this structure, each boron atom is bonded tetrahedrally to four oxygen atoms
(two O1 and two O2), as depicted in Fig. 1(h), and each oxygen atom to two boron atoms, creating a three
dimensional network of BO4 tetrahedra. The six-membered B3O3 ring has 3(C3) symmetry in a flat chair
conformation, as shown in Fig. 1(i). The γ-(BOH)3O3 polymorph has space group P3n, its unit cell has 24
molecules with a = 8.886 Å, and its atomic coordinates are shown in Table ST3 (see Supporting
Information).

2.2. DFT Computational Details

Computational calculations for the three crystalline phases of metaboric acid were performed
32–34
within the density functional theory (DFT) framework using the crystallographic data found in
15,31
references , as mentioned in the previous section. The plane wave basis set CASTEP code 35,36 was
employed to run total energy calculations and structural optimization. Two exchange-correlation
functionals were applied: the total local-density approximation (LDA) 37,38 and the generalized gradient
39
approximation (GGA) parametrized by Perdew, Burke and Ernzerhof (PBE) . As the standard PBE
functional cannot describe the van der Waals interactions between the molecules or planes of the crystals,
we have also adopted the dispersion energy correction scheme of Tkatchenko and Scheffler (TS) 30, which
include the van der Waals interactions 40 due to electron density fluctuations 41. The inclusion of dispersive
forces is essential to provide an accurate structural description of molecular crystals and their optical
properties42.

Norm-conserving pseudopotentials 43 were used to describe the core electrons of all non-hydrogen
atoms. The geometry optimizations of the metaboric acid unit cells were performed adopting plane wave

5
basis sets with kinetic-energy cutoffs of 500 and 830 eV. In the simulations, the atomic positions, lattice
lengths, and cell angles were allowed to relax. Convergence tolerances were: total energy variation smaller
than 0.50 × 10−5 eV/atom, maximum force per atom below 0.01 eV/Å, pressure smaller than 0.02 GPa,
and maximum atomic displacement smaller than 0.10 × 10−3 Å. The BFGS minimizer 44 was employed to
perform the unit cell optimizations. After trying different basis set energy cutoffs, the 830 eV value was
taken to be safe to perform accurate calculations of the electronic and optical properties of all the
optimized GGA+TS unit cells.

Monkhorst−Pack 45 k-point grids of 2 × 1 × 2 (α-) and 2 × 2 × 2 (β-, γ-) were used to sample the
Brillouin zones in order to evaluate reciprocal space integrals. Kohn-Sham frontier orbitals of the isolated
metaboric acid molecule (HOMO and LUMO) and their crystal structures at the Γ-point were obtained
using the GGA+TS optimized crystals using the DMOL3 code 46. The calculations were performed using
the same GGA+TS exchange-correlation functional and a Double Numerical plus Polarization (DNP)
basis set with accuracy similar to the plane wave basis set adopted for the CASTEP computations. The
optical absorption α(ω) and the complex dielectric function ϵ(ω)=ϵ1(ω)+iϵ2(ω) curves for the α-, β-, and
γ-metaboric acid polymorphs, in particular, were calculated at the GGA+TS830 level following the same
scheme described in previous works 19,21,23–25.

3. RESULTS AND DISCUSSION

3.1. Structural Features of the Metaboric Acid Polymorphs

In Table 1 we have a comparison between the experimental parameters of the unit cell of each
metaboric acid polymorph and the results obtained using the pure LDA and GGA exchange-correlation
functionals and the GGA+TS dispersion corrected approach for a plane wave basis energy cutoff of 830
eV. For the α-polymorph, the LDA lattice parameters are underestimated, with values worsening as the
energy cutoff increases, a result already expected from the typical LDA overbinding effect. The pure GGA
functional, on the other hand, gives an improved description of hydrogen bonds, but usually predicts
lattice parameters a bit large. The best agreement with the measured data is observed when one takes into
account the dispersive interactions employing the TS scheme, with errors smaller than 0.1 Å for a, b, and
c. As a matter of fact, the predicted unit cell volume according to the GGA+TS simulation is only 2.9%
smaller than the experimental unit cell lengths. This indicates that incorporating the effect of van der

6
Waals interactions is an essential feature required to provide a good assessment of the structural properties
of this polymorph.

Table 1. Lattice parameters for the orthorhombic, monoclinic and cubic metaboric acid polymorphs
according to the LDA, GGA-PBE and GGA+TS calculations with 830 eV of cutoff energy. a, b, c are
given in Å, while angles α, β, γ are in degrees and volumes (V) in Å3.

Crystal Polymorph 𝒂𝒂 Δ𝒂𝒂 𝒃𝒃 Δ𝒃𝒃 𝒄𝒄 Δ𝒄𝒄 α Δα β Δβ γ Δγ V ΔV

488.04
α-(BOH)3O3 – Exp 15 8.05 – 9.69 – 6.26 – 90.0 – 90.0 – 90.0 –
α-(BOH)3O3 LDA 7.74 -0.31 9.30 -0.39 5.80 -0.46 90.0 – 90.0 – 90.0 – 417.60 -70.45
α-(BOH)3O3 GGA 8.16 0.11 9.59 -0.10 7.45 1.19 90.0 – 90.0 – 90.0 – 583.05 95.01
α-(BOH)3O3 GGA+TS 7.97 -0.08 9.60 -0.09 6.20 -0.06 90.0 – 90.0 – 90.0 – 474.05 -13.99

β-(BOH)3O3 – Exp 16 7.12 – 8.84 – 6.77 – 90.0 – 90.0 – 93.26 – 425.70 –

β-(BOH)3O3 LDA 6.77 -0.35 8.77 -0.08 6.57 -0.20 90.0 – 90.0 – 94.13 0.87 388.97 -36.73
β-(BOH)3O3 GGA 7.63 0.51 8.89 0.05 6.88 0.11 90.0 – 90.0 – 89.52 -3.74 466.97 41.27
β-(BOH)3O3 GGA+TS 7.12 0.00 8.90 0.06 6.68 -0.09 90.0 – 90.0 – 94.15 0.89 422.41 -3.29

γ-(BOH)3O3 – Exp 31 8.88 – 8.88 – 8.88 – 90.0 – 90.0 – 90.0 – 700.5 –

γ-(BOH)3O3 LDA 8.76 -0.12 8.76 -0.12 8.76 -0.12 90.0 – 90.0 – 90.0 – 671.5 -29.00
γ-(BOH)3O3 GGA 8.94 0.06 8.94 0.06 8.94 0.06 90.0 – 90.0 – 90.0 – 714.1 13.60
γ-(BOH)3O3 GGA+TS 8.89 0.01 8.89 0.01 8.89 0.01 90.0 – 90.0 – 90.0 – 702.6 2.11

For the monoclinic β crystal, which resembles a polymer crystal in contrast with the α system, the
GGA functional error for the a parameter is 0.51 Å at 830 eV. The other lattice parameters are more
accurate, with deviations of 0.05 and 0.11 Å relative to experiment. Interestingly, the inclusion of
dispersive forces improves these figures even more, with the a lattice length having exactly the measured
value and the other parameters, b and c, being only 0.06 Å larger and 0.09 Å smaller than the experimental
data. If one notes that the metaboric acid polymeric chains are disposed along the b axis (see Fig. 1(d)), it
is reasonable to infer that dispersive forces play an important role in their stabilization along the a and c
directions (the latter with a more significant contribution from hydrogen bonds). The β angle for the
o
GGA+TS simulation is just 0.89 larger than the measurement, while the LDA is just slightly more
accurate (0.87 o). The GGA+TS unit cell volume, however, is only 0.8% smaller than the X-ray data, in
comparison with corresponding errors of 9.7% and -8.6% using the GGA and LDA approaches,
respectively.

Looking now to covalent cubic γ-metaboric acid (bottom of Table 1), the LDA lattice parameter
predicts a unit cell length smaller by 0.12 Å in comparison to the experiment. The pure GGA functional,
in contrast, predicts a value 0.06 Å larger. Even for this structure, where van der Waals forces are not
expected to be much relevant due to the dominant covalent character of its interatomic forces, the inclusion
of dispersive effects improves the description, with the GGA+TS prediction for a being 0.01 Å larger than

7
the measurement. The GGA+TS unit cell volume is only 0.3% above the experimental value. Overall, one
can conclude that the inclusion of the dispersion correction scheme leads to significantly more accurate
structural predictions for all metaboric acid polymorphs, especially the molecular (α) and polymeric (β)
systems, but being also relevant for the mostly covalent (γ) crystal. For this reason, we will attain ourselves
only to the GGA+TS results in the remaining of our study. More detailed results on the structural
properties (atom positions, bond angles) are included in the Supporting Information of the paper (Tables
ST1 to ST6).

Figure 2 depicts the behavior of the unit cell total energy Ecell of the GGA+TS optimized crystals
as we vary independently the lattice parameters. For the α system, there is an asymmetry between the Ecell
curves for a and b, on one side, and c, on the other. As the curvature of Ecell for a and b changes is much
more significant than for c, one conclude that the α polymorph is softer along c (the stacking direction),
which is certainly caused by the weak character of its interplanar interactions. In the case of monoclinic
β metaboric acid, the behavior of Ecell is unique for each direction, with deformations along a exhibiting
the smallest variation of Ecell, deformations along c displaying intermediary curvature and deformations
along b producing the largest amount of energy variation. So, for this crystal, stiffness must decrease for
deformations along b, c, and a, in respective order. Lastly, for the cubic γ-polymorph the Ecell curves are
perfectly symmetric, revealing a very isotropic material in its mechanical properties.

Figure 2. DFT calculated (GGA+TS830 level) total energy E of the α-(BOH)3O3, β-(BOH)3O3, and γ-

(BOH)3O3 metaboric acid unit cells as a function of lattice parameter deviations Δa/ao, Δb/bo, and Δc/co

relative to the converged optimal parameters ao, bo, and co, respectively.

8
3.2 Electronic Properties

The Kohn-Sham electronic band structures obtained for the metaboric acid polymorphs adopt the
nomenclature for its high symmetry points in the first Brillouin zone following the schemes shown in Fig.
3. In Figs. 4, 5 and 6, the full band structures (top) and close-ups near the main band gap (bottom) are
presented for the α, β, and γ metaboric acid systems, respectively, with side panels depicting per-atom
and per-orbital partial densities of states (PDOS). The calculated main energy band gaps are presented in
Table 2 in comparison with the results for the 2A triclinic boric acid.

Figure 3. First Brillouin zones of the metaboric acid polymorphs α-(BOH)3O3, β-(BOH)3O3, and γ-

(BOH)3O3. High-symmetry points are indicated.

9
Figure 4. Band structure and electron densities of states of the α-(BOH)3O3 polymorph. Top: Full

electronic band structure in the -21.0 to 12.0 eV energy range (left) and the s, p, and total densities of

states (right). Bottom: close-up of the band structure (left) and H s, B s, B p, O s, and O p electron densities

of states (right) near the main band gap.

10
In the α-phase, one can see a set of valence bands between -8.0 eV and 0.0 eV with strong p-character
(top of Fig. 4), and deep bands between -20.0 and -17.0 eV mostly originated from s states. The conduction
band starts at about 6.0 eV with a nearly parabolic minimum at the Γ point and a dense set of bands nearly at
8.0 eV with a high density of states, mostly due to p orbitals. A close-up near the main band gap reveals that
the top of its valence band is very flat, especially along the Γ-Z, Γ-T, and Γ-Y lines, with maximum at the Γ
point, implying in a direct valence-to-conduction band gap of 6.26 eV. The valence band maximum originates
mainly from O 2p states, with smaller contribution from B 2p levels, while the bottom of the conduction band
has contributions from B 2p, H 1s, and O 2p atomic orbitals. This can be contrasted with the 2A boric acid
polymorph, which exhibits an indirect gap of 6.25 eV and a direct gap (Γ→Γ) of 6.35 eV, with similar orbital
contributions for the frontier bands, as shown in reference 8. As the 2A boric acid transforms into the metaboric
α-phase by heating past the temperature of 80-100 oC, an electronic gap type transition from indirect to direct
can be promoted just by varying the temperature of a boric acid solid sample. One must remark, however, that
DFT calculations tend to underestimate band gap values significantly in comparison with experimental results
47
.

Qualitatively, the band structure of the monoclinic polymeric β polymorph (see Fig. 5) shares some
similarities with the orthorhombic α phase. For example, the valence bands extend approximately between the
same energy ranges and are derived from nearly the same atomic states. However, the density of valence bands
in the β system is higher, with less gaps being observed between -8.0 and 0.0 eV. The conduction band
minimum still occurs at the Γ point with a significant amount of dispersion, but with the conduction bands
nearby being a bit energetically shifted down towards it (Fig. 5, bottom). This minimum originates mostly
from B 2p levels. Two indirect gaps of 6.56 eV occur involving the U→Γ and α→Γ transitions, while the
direct Γ→Γ gap is 6.60 eV. The uppermost valence bands are less flat than the ones observed for the α crystal.

Table 2. Calculated energy gaps (eV) using GGA+ TS approach for the main electronic transitions involving

valence band (VB) and conduction band (CB) of the α-, β- and γ- metaboric acid polymorphs. Data for the

H3BO3-2A are also presented for comparison.

Crystalline structures
Transition H3BO3- 2A α-(BOH)3O3 β-(BOH)3O3 γ-(BOH)3O3
VB CB (triclinic)8 (orthorhombic) (monoclinic) (cubic)
Z Γ 6.34 (i) - - -
Q Γ 6.25 (i) - - -
F Γ 6.28 (i) - - -
Γ Γ 6.35 (d) 6.26 (d) 6.60 (d) -
T Γ - 6.29 (i) - 7.22 (d)
U Γ - - 6.56 (i) -
R Γ - - 6.56 (i) 7.25 (i)

Gap types (d: direct/i: indirect)


11
Figure 5. Band structure and electron densities of states of the β-(BOH)3O3 polymorph. Top: Full electronic

band structure in the -21.0 to 12.0 eV energy range (left) and the s, p, and total densities of states (right).

Bottom: close-up of the band structure (left) and H s, B s, B p, O s, and O p electron densities of states (right)

near the main band gap.

Considering now the covalent cubic γ-polymorph (Fig. 6), one can discern a set of valence bands below
-17.0 eV (mostly s-like) and two sets of dense valence bands between -9.0 and -4.0 eV and between -3.5 eV
and 0.0 eV, respectively. These two sets have strong p-character, with some s contribution mostly for the
lowest energy values. In the conduction band, we have two distinct curves between 7.0 and 8.0 eV and a set
of dense bands above 9.0 eV (Fig. 6, bottom). Both the conduction band minimum and the valence band
maximum are at the Γ point, with a direct Γ→Γ gap of 7.22 eV and a secondary direct gap of 7.98 eV at the
M point. As occurred for the other polymorphs, the top of the valence band originates from O 2p and B 2p
states, while the bottom of the conduction band has significant contribution from H 1s, B 2p, and B 2s states,
in decreasing order of relative importance.

12
Figure 6. Band structure and electron densities of states of the γ-(BOH)3O3 polymorph. Top: Full electronic

band structure in the -21.0 to 12.0 eV energy range (left) and the s, p, and total densities of states (right).

Bottom: close-up of the band structure (left) and H s, B s, B p, O s, and O p electron densities of states (right)

near the main band gap.

In order to better understand the nature of the Kohn-Sham orbitals near the main band gap, we have
calculated the frontier molecular orbitals, HOMO (Highest Occupied Molecular Orbital) and LUMO (Lowest
Unoccupied Molecular Orbital), for the metaboric acid molecule using the GGA+TS exchange-correlation
functional. Their plots are shown in Fig. 7, together with an estimated HOMO-LUMO gap of 8.09 eV.
Equivalent orbitals for the crystals, HOVB (Highest Occupied Valence Band at Γ) and LUCB (Lowest
Unoccupied Conduction Band at Γ) are shown for the three polymorphs in Fig. 8. The HOMO orbital of the
metaboric acid molecule presents a set of three large p-like lobes at the oxygen atoms O3, O5, and O6 (atom
labels at the top of Fig. 1) , and a secondary set with smaller size lobes at O1 and O2. There is a very small
pair of p lobes at O4. The LUMO orbital, on the other hand, exhibits three π-like structures involving the three
boron atoms and their adjacent oxygen atoms in the B3O3 ring. The p lobes at O1 are larger than the observed
for the HOMO, while the O3, O5, and O6 lobes are smaller. One can say therefore that for the HOMO state,
the electrons are more localized outside the B3O3 ring, while for the LUMO they prefer to stay in it.

13
Figure 7. GGA+TS calculated HOMO and LUMO orbitals for the metaboric acid (BOH)3O3 molecule. The

HOMO-LUMO transition energy of 8.09 eV is indicated.

Looking now to the HOVB and LUCB orbitals of the α polymorph (Fig. 8, top), one can see that, for
the HOVB, we have a set of diffuse p-like structures at the O4 and O5 atoms contained within the metaboric
acid layers and forming a zig-zag pattern across the structure. More concentrated p-like structures can be seen
at the O1, O2, and O3, with the first and the latter being involved in hydrogen bonds. For the LUCB, on the
other hand, we note the existence of an alternate pattern of sp-like surfaces around the oxygen atoms with
positive (blue) phase and the boron atoms with negative (yellow) phase. The π-like pattern observed in the
free molecule is absent. Concentrated sp lobes can also be seen at the H1 and H3 atoms, which form hydrogen
bonds. In the polymeric β crystal (Fig. 8, middle) it is not very easy to discern the electronic configurations,
but a careful analysis reveals that the HOVB state has a set of extended p-like isosurfaces around the oxygen
atoms O1, O3, and more concentrated ones at O2, O4, O5, and O6 (atom labels shown in Fig. 1, middle). For
the LUCB level, the oxygen atoms exhibit sp-like structures and the boron atoms are surrounded by large sp-
like lobes above and below the molecular plane extending to the nearest adjacent polymeric chain. The
hydrogen atoms also exhibit some sp isosurfaces around them, resembling in many respects the situation
obtained for the α system. Finally, for the γ-metaboric acid polymorph (Fig. 8, bottom), the HOVB state is
formed from p-like isosurfaces at the oxygen atoms O2 in the B3O3 ring (see Fig. 1, bottom), and the O1 atom
connecting them. As these isosurfaces are oriented along distinct directions in three dimensions for each
molecular unit, the electronic environment for this orbital must be approximately isotropic. The LUCB orbital,
on the other hand, has sp-like features at the O1 and O2 atoms and an overlap of s orbitals for neighbor
hydrogen atoms (blue lobes), while the boron atoms do not contribute significantly.

14
Figure 8. GGA+TS calculated highest occupied valence band (HOVB, left column) and lowest unoccupied

conduction band (LUCB, right column) orbitals at the Γ point of the α-(BOH)3O3 (top), β-(BOH)3O3 (middle),

and γ-(BOH)3O3 (bottom) metaboric acid polymorphs.

3.3 Optical Properties

The energy-dependent electronic complex dielectric function 𝜖𝜖(𝜔𝜔) = 𝜖𝜖1 (𝜔𝜔) + 𝑖𝑖𝜖𝜖2 (𝜔𝜔) of the α, β, and
γ metaboric acid crystals was calculated considering incident polarized light along the (100), (010), and (001)
planes and a polycrystalline sample. As all electronic transitions are estimated to occur above 6.2 eV in the
metaboric acid polymorphs, hardware limitations have prevented us to perform optical absorption
measurements to compare with our theoretical results. Besides, to the best of our knowledge, there are no

15
experimental data published in the literature concerning on the optical properties of metaboric acid in the solid
state.

The results are shown in Fig. 9. The real and imaginary parts of the complex dielectric function for the
orthorhombic metaboric acid crystal α (Fig. 9, left) reveal a remarkable asymmetry with respect to incident
light polarization. For light polarized along the 001 plane (perpendicular to the metaboric acid molecular
layers), we have important maxima of 𝜖𝜖1 and 𝜖𝜖2 near 15.3 eV (𝜖𝜖1 = 3.5), but both curves have much less
features than the observed for the 100 and 010 directions (parallel to the molecular layers), which exhibit
pronounced peaks near 8 eV (𝜖𝜖1 = 8 for 010 polarization and 12 for 100 polarization) related to valence to
conduction band transitions involving O 2p valence and H 1s, B 2s conduction levels with high density of
states (see top of Fig. 4, the DOS maximum at nearly 8 eV). The polycrystalline case resembles closely the
curves for the 010 case.

Figure 9. Calculated complex dielectric function 𝜖𝜖(𝜔𝜔) = 𝜖𝜖1 (𝜔𝜔) + 𝑖𝑖𝜖𝜖2 (𝜔𝜔) for the α-, β-, and γ-(BOH)3O3

polymorphs considering incident polarized light along the 001, 010, and 100 crystal planes and a

polycrystalline (POLY) sample. The red curves depict the real part 𝜖𝜖1 (𝜔𝜔), while the black curves depict the

imaginary part 𝜖𝜖2 (𝜔𝜔).

For the β polymorph, the dielectric function curves along the three polarizations investigated (100,
010, and 001, Fig. 9, middle) have some interesting features near 8.0 eV, also due to electron transitions
involving O 2p-like valence and H 1s, B 2s conduction levels, as occurred for the α system. However, the
16
maxima are less intense, which is probably due to the smaller density of states of the conduction bands near
8.0 eV in comparison to the orthorhombic phase (Fig. 5, middle). Above 10.0 eV, the 𝜖𝜖1 and 𝜖𝜖2 curves exhibit
very small variation. For energies below 10.0 eV, different polarization states for the incident light lead to
distinct dielectric function profiles, with the 001 curve exhibiting well defined maxima at 7.9 eV, while the
010 curve has a broad structure of peaks in the same region and the 100 curves are smoother. The simulated
polycrystalline sample has maxima near 8.0 eV as well. Lastly, the plots of 𝜖𝜖1 and 𝜖𝜖2 for cubic γ metaboric
acid are very isotropic, practically independent from incident light polarization as an effect of the isotropy of
this crystalline phase. The real part of the dielectric function increases slightly between 7 and 10 eV and
decreases between 10 and 12.5 eV., but well defined maxima or minima are absent, as the electron densities
of states vary smoothly and their corresponding bands have more intense dispersion (see Fig. 6). The optical
absorption curves shown in Fig. 10 are proportional to the imaginary part 𝜖𝜖2 of the complex dielectric function,
so most of their features were already discussed. As practical UV-vis spectroscopy has an effective work range
for wavelengths above 200 nm and the typical gaps predicted involve wavelengths of about 200 nm or smaller,
we could not perform comparisons with experimental data with the equipment available to our research group.

Figure 10. DFT-calculated GGA+TS830 optical absorption for the α-, β-, and γ-(BOH)3O3 polymorphs

considering incident polarized light along the 001, 010, and 100 crystal planes and a polycrystalline (POLY)

sample.

4. CONCLUSIONS

17
Density functional theory calculations have been performed in order to investigate the structural,
electronic and optical properties of three metaboric acid crystal polymorphs, α, β, and γ. The calculations
employed three distinct exchange-correlation energy models: local density approximation (LDA), generalized
gradient approximation (GGA), and generalized gradient approximation including a dispersion energy term
(GGA+TS). The structural properties obtained (lattice parameters, unit cell angle) for each crystal – after a
careful convergence study – were compared with experimental data, revealing that the inclusion of dispersion
corrections is essential to find an accurate picture of these systems, reducing the error from 7% (LDA, α-
polymorph, c parameter) to less than 1% (GGA+TS). This error reduction was more pronounced for the
molecular (α-) and polymeric (β-) polymorphs. A study of the unit cell energy as a function of the unit cell
parameters revealed that the α-polymorph is softer along the stacking direction of its metaboric acid planes,
while the β-polymorph is mechanically more anisotropic. Analysis of the Kohn-Sham band structures reveal
that all crystals are insulators, with the following band gaps: 6.26 eV, direct (Γ→Γ), for the α-polymorph; 6.56
eV, indirect (U→Γ and α→Γ), for the β-polymorph; and 7.22 eV, direct (Γ→Γ), for the γ-polymorph. An
indirect to direct gap transition can be produced, therefore, by heating a indirect gap 2A boric acid sample
above 100 oC, inducing a phase change to the direct gap metaboric α structure. The uppermost valence band
states of the metaboric acid polymorphs are O 2p in character while the lowest conduction band originates
mostly from O 2p levels but with some relevant contributions from H 1s and B 2s, and B 2p orbitals. The
HOMO and LUMO orbitals of metaboric acid and the corresponding HOVB and LUCB orbitals at the Γ point
for the polymorphs were also determined. The HOMO-LUMO transition involves the transfer of electron
charge from the periphery of the molecule towards its B3O3 ring in a π-like configuration. The HOVB and
LUCB orbitals of the α- and β-polymorphs share many features, while for the γ-polymorph the LUCB orbital
displays a certain overlap of s orbitals involving two hydrogen atoms in distinct molecular units and the
absence of significant contributions from the boron atoms. Lastly, the complex dielectric function and optical
absorption related to electron transitions were obtained for each polymorph considering different
configurations of polarized incident light and a polycrystalline sample. The α system exhibited some optical
anisotropy between light polarizations parallel and perpendicular to the metaboric acid planes, while the β(γ)-
polymorph is optically anisotropic (isotropic) along the three crystalline axes. The most relevant maxima and
minima of the optical properties occur near to 8.0 eV, being due to O 2p → H 1s B 2s transitions and with
intensity depending on the electron density of states of the conduction band at 8.0 eV.

CONFLICTS OF INTEREST

There are no conflicts of interest to declare.

ACKNOWLEDGEMENTS

V. N. F., and A. Valentini are senior researchers from the Brazilian National Research Council (CNPq),
and would like to acknowledge the financial support received during the development of this work. E. W. S.

18
C. and M. B. da Silva received financial support from CNPq projects 304781/2016-9 and 140898/2016-6,
respectively.

REFERENCES

1 F. H. Nielsen, J. Trace Elem. Med. Biol., 2014, 28, 383–387.

2 E. D. Farfán-García, N. T. Castillo-Mendieta, F. J. Ciprés-Flores, I. I. Padilla-Martínez, J. G. Trujillo-


Ferrara and M. A. Soriano-Ursúa, Toxicol. Lett., 2016, 258, 115–125.

3 L. C. Bhami and S. S. M. Das, J. Vector Borne Dis., 2015, 52, 147–152.

4 W. A. Qualls, G. C. Müller, S. F. Traore, M. M. Traore, K. L. Arheart, S. Doumbia, Y. Schlein, V. D.


Kravchenko, R.-D. Xue and J. C. Beier, Malar. J., 2015, 14, 301.

5 G. Benelli and H. Mehlhorn, Parasitol. Res., 2016, 115, 1747–1754.

6 N. E. A. Murray, M. B. Quam and A. Wilder-Smith, Clin. Epidemiol., 2013, 5, 299–309.

7 D. J. G. Didier Musso, Clin. Microbiol. Rev., 2016, 29, 487–524.

8 M. B. da Silva, R. C. R. dos Santos, A. M. da Cunha, A. Valentini, O. D. L. Pessoa, E. W. S. Caetano


and V. N. Freire, Cryst. Growth Des., 2016, 16, 6631–6640.

9 M. Elango, R. Parthasarathi, V. Subramanian and N. Sathyamurthy, J. Phys. Chem. A, 2005, 109,


8587–8593.

10 A. N. Enyashin and A. L. Ivanovskii, Chem. Phys. Lett., 2005, 411, 186–191.

11 W. Gao, Y. Jing, J. Yang, Z. Zhou, D. Yang, J. Sun, J. Lin, R. Cong and T. Yang, Inorg. Chem.,
2014, 53, 2364–2366.

12 D. L. Dorset, Acta Crystallogr. Sect. A Found. Crystallogr., 1992, 48, 568–574.

13 R. R. Shuvalov and P. C. Burns, Acta Crystallogr. Sect. C Cryst. Struct. Commun., 2003, 59, 0–3.

14 K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, V. V Khotkevich, S. V Morozov and A. K. Geim,


Proc. Natl. Acad. Sci. U. S. A., 2005, 102, 10451–10453.

15 C. R. Peters and M. E. Milberg, Acta Crystallogr., 1964, 17, 229–234.

16 W. H. Zachariasen, Acta Crystallogr., 1963, 16, 385–389.

17 W. H. Zachariasen, Acta Crystallogr., 1963, 16, 380–384.

18 E. W. S. Caetano, J. R. Pinheiro, M. Zimmer, V. N. Freire, G. A. Farias, G. A. Bezerra, B. S. Cavada,


J. R. L. Fernandez, J. R. Leite, M. C. F. De Oliveira, J. A. Pinheiro, J. L. De Lima Filho and H. W.

19
Leite Alves, AIP Conf. Proc., 2005, 772, 1095–1096.

19 M. Z. S. Flores, V. N. Freire, R. P. Dos Santos, G. A. Farias, E. W. S. Caetano, M. C. F. De Oliveira,


J. R. L. Fernandez, L. M. R. Scolfaro, M. J. B. Bezerra, T. M. Oliveira, G. A. Bezerra, B. S. Cavada
and H. W. Leite Alves, Phys. Rev. B - Condens. Matter Mater. Phys., 2008, 77, 115104.

20 F. F. Maia, V. N. Freire, E. W. S. Caetano, D. L. Azevedo, F. A. M. Sales and E. L. Albuquerque, J.


Chem. Phys., 2011, 134, 175101.

21 A. M. Silva, B. P. Silva, F. A. M. Sales, V. N. Freire, E. Moreira, U. L. Fulco, E. L. Albuquerque, F.


F. Maia and E. W. S. Caetano, Phys. Rev. B, 2012, 86, 195201.

22 J. G. da Silva Filho, V. N. Freire, E. W. S. Caetano, L. O. Ladeira, U. L. Fulco and E. L.


Albuquerque, Chem. Phys. Lett., 2013, 587, 20–24.

23 A. M. Silva, S. N. Costa, B. P. Silva, V. N. Freire, U. L. Fulco, E. L. Albuquerque, E. W. S. Caetano


and F. F. Maia, Cryst. Growth Des., 2013, 13, 4844–4851.

24 S. N. Costa, F. A. M. Sales, V. N. Freire, F. F. Maia, E. W. S. Caetano, L. O. Ladeira, E. L.


Albuquerque and U. L. Fulco, Cryst. Growth Des., 2013, 13, 2793–2802.

25 G. Zanatta, C. Gottfried, A. M. Silva, E. W. S. Caetano, F. A. M. Sales and V. N. Freire, J. Chem.


Phys., 2014, 140, 124511.

26 A. M. Silva, S. N. Costa, F. A. M. Sales, V. N. Freire, E. M. Bezerra, R. P. Santos, U. L. Fulco, E. L.


Albuquerque and E. W. S. Caetano, J. Phys. Chem. A, 2015, 119, 11791–11803.

27 K. Berland, V. R. Cooper, K. Lee, E. Schröder, T. Thonhauser, P. Hyldgaard and B. I. Lundqvist,


Reports Prog. Phys., 2015, 78, 66501.

28 J. Klime and A. Michaelides, J. Chem. Phys., , DOI:10.1063/1.4754130.

29 G. J. O. Beran, Chem. Rev., 2016, 116, 567 − 561.

30 A. Tkatchenko and M. Scheffler, Phys. Rev. Lett., 2009, 102, 6–9.

31 C. C. Freyhardt, M. Wiebcke and J. Felsche, Acta Crystallogr. C., 2000, 56 (Pt 3), 276–278.

32 P. Hohenberg and W. Kohn, Phys. Rev., 1964, 136, B864–B871.

33 W. Kohn and L. J. Sham, Phys. Rev., 1965, 140, A1133–A1138.

34 R. O. Jones, Rev. Mod. Phys., 2015, 87, 897–923.

35 M. D. Segall, P. J. D. Lindan, M. J. Probert, C. J. Pickard, P. J. Hasnip, S. J. Clark and M. C. Payne,


J. Phys. Condens. Matter, 2002, 14, 2717–2744.

20
36 S. J. Clark, M. D. Segall, C. J. Pickard, P. J. Hasnip, M. I. J. Probert, K. Refson and M. C. Payne,
Zeitschrift fur Krist., 2005, 220, 567–570.

37 D. M. Ceperley and B. J. Alder, Phys. Rev. Lett., 1980, 45, 566–569.

38 J. P. Perdew and Alex Zunger, Phys. Rev. B, 1981, 23, 5048.

39 J. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865–3868.

40 A. Tkatchenko, Adv. Funct. Mater., 2015, 25, 2054–2061.

41 S. Grimme, A. Hansen, J. G. Brandenburg and C. Bannwarth, Chem. Rev., 2016, 116, 5105–5154.

42 J. Hoja, A. M. Reilly and A. Tkatchenko, Wiley Interdiscip. Rev. Comput. Mol. Sci., 2017, 7, e1294.

43 M. C. P. J. S. Lin, A. Qteish, Phys. Rev. B, 1993, 47, 4174.

44 B. G. Pfrommer, M. Cote, S. G. Louie and M. L. Cohen, J. Comput. Phys., 1997, 131, 233–240.

45 H. J. Monkhorst and J. D. Pack, Phys. Rev. B, 1976, 13, 5188–5192.

46 B. Delley, J. Chem. Phys., 2000, 113, 7756.

47 S.-P. Gao, Comput. Mater. Sci., 2012, 61, 266–269.

21

Das könnte Ihnen auch gefallen