Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

This is an open access article published under an ACS AuthorChoice License, which permits

copying and redistribution of the article or any adaptations for non-commercial purposes.

Article

Cite This: ACS Omega 2018, 3, 18917−18924 http://pubs.acs.org/journal/acsodf

Electronic and Optical Properties of Ultrasmall ABX3 (A = Cs,


CH3NH3/B = Ge, Pb, Sn, Ca, Sr/X = Cl, Br, I) Perovskite Quantum Dots
Athanasios Koliogiorgos, Christos S. Garoufalis,* Iosif Galanakis,* and Sotirios Baskoutas*
Department of Materials Science, School of Natural Sciences, University of Patras, 26504 Patras, Greece

ABSTRACT: Perovskite quantum dots (QDs) constitute a novel and rapidly developing
field of nanotechnology with promising potential for optoelectronic applications. However,
few perovskite materials for QDs and other nanostructures have been theoretically explored.
In this study, we present a wide spectrum of different hybrid halide perovskite cuboid-like
QDs with the general formula of ABX3 with varying sizes well below the Bohr exciton radius.
Density functional theory (DFT) and time-dependent DFT calculations were employed to
Downloaded from pubs.acs.org by 46.148.112.39 on 01/01/19. For personal use only.

determine their structural, electronic, and optical properties. Our calculations include both
stoichiometric and nonstoichiometric QDs, and our results reveal several materials with high
optical absorption and application-suitable electronic and optical gaps. Our study highlights
the potential as well as the challenges and issues regarding nanostructured halide perovskite
materials, laying the background for future theoretical and experimental work.
ACS Omega 2018.3:18917-18924.

1. INTRODUCTION properties, such as low cost, low difficulty of synthesis, and


As the quest for clean, source-abundant ways for energy high carrier mobility.
harvesting in the 21st century reaches a peak, perovskite In addition to the vast amount of literature on bulk
materials have played a significant role in the research for solar perovskites, there is a limited yet rapidly growing number of
energy applications, such as photovoltaics and light-emitting experimental studies on halide perovskite nanostructures,
diodes (LEDs).1−3 An enormous amount of theoretical and mainly QDs. Most of these studies regard a few specific
experimental work in the last 2 decades has, and continues to, materials such as cesium lead (CsPbBr3)1−8 and methyl-
explored the potential of these materials. The majority of the ammonium lead perovskites, such as MAPbBr 3 and
research on perovskites has been concentrated on macro- MAPbI3.9−13 On the other hand, very few theoretical studies
structured bulk version of these materials. However, in the last on perovskite QDs with first-principles calculations have been
years, an increasing amount of research has been carried out on conducted so far.14,15 Thus, a thorough, ab initio study of a
the nanostructures of perovskite materials, in the form of great variety of various-sized QDs of different materials of the
quantum dots (QDs) as well as nanowires and nanorods. The MABX3 formula was in place and could be a basis for future in-
size of these materials (below 10 nm, and even as small as 2 depth experimental and theoretical research.
nm) ensures the occurrence of quantum confinement effects, Experimental studies in perovskite QDs share some
which enable the materials under study to be used in fields common features that were used as a guide for our study.
where quantum mechanical properties play a decisive role, Some of these features include the shape of the QDs, which is
such as LED technology and even materials for quantum generally cuboid. Substitution of halide anions (i.e., Br with
computers. Cl) does not change the cubic structure of the QDs.10 In order
A perovskite material has the general formula of ABX3, for the structure to be considered a QD, its average size must
where if A is a monovalent cation, then B is usually a divalent be under 2 × rB, where rB is the exciton Bohr radius. The Bohr
cation of dissimilar size, and X is a monovalent anion to radii for MAPbBr3 and MAPbI3, for instance, are 2 and 2.2 nm,
achieve charge neutrality. In the case of emerging halide respectively.11 Generally, the perovskite QDs studied have a
perovskites, X is a halogen such as Cl, Br, or I. The A cation diameter below 10 nm, with a general average size of 3 nm, and
can be inorganic, such as Cs, or organic, such as a some of them even less than 1.8 nm.4,5,10,12,13,16−18 Another
methylammonium cation (CH3 NH3 +, MA). The basic shared feature of the studied QDs is that, compared to the
structure of ABX3 formula is an octahedron, whereas the corresponding band gap of the bulk material, their highest
final structure of the bulk material can be anything from a high- occupied molecular orbital (HOMO)−lowest unoccupied
symmetry cuboid to a low-symmetry monoclinic structure. The molecular orbital (LUMO) gap is always wider and increases
family of perovskite materials used in this study consists of further as the QD size decreases.10 The gap generally decreases
MABX3 and CsBX3 perovskites with cubic structures, where B
is a cation such as Pb, Sn, Ge, Ca, and Sr and X is a halogen Received: September 26, 2018
such as Cl, Br, and I. These materials have been well studied in Accepted: December 18, 2018
the literature and exhibit an abundance of sought-after Published: December 31, 2018

© 2018 American Chemical Society 18917 DOI: 10.1021/acsomega.8b02525


ACS Omega 2018, 3, 18917−18924
ACS Omega Article

Figure 1. (a) ST CsPbX3 QDs, (b) hydrogen-passivated NST CsPbX3 QDs, and (c) hydrogen-passivated NST MAPbX3 QDs.

exponentially as the QD size increases, converging to the bulk energy of the lowest spin allowed transition regardless of its
value when the QD size surpasses the 2 × rB limit.5 Similarly, oscillator strength. This was necessary because all the alkaline
the exciton binding energy, which is significantly enhanced in earth metal dots exhibit vanishing oscillator strengths.
small QDs, is a decreasing function of size. For example, the Confusion is avoided by noticing the strength values which
bulk value of Ex for MAPbBr3 is 65 meV, whereas in the small are explicitly presented in all absorption spectra plots.
QD regime, it can be as large as ∼300−400 meV.9,10,12 As the aim of the project is to extend to larger
The ability of many of the experimentally studied perovskite nanostructures, we chose a basis set of double zeta quality
QDs to exhibit strong optical absorption and emission in the (namely, the def2-SVP basis sets of the Karlsruhe group25,26),
visible optical regime makes them suitable candidates for which has been extensively tested to numerous cases and in the
optoelectronic applications such as LEDs. In the present study, same time is small enough to allow for calculations in systems
we focus on the electronic and optical properties of QDs of the with hundreds of atoms. All calculations were performed with
halide perovskites having the chemical formula ABX3, where A the ORCA suite of programs.27,28
is either Cs or the methylammonium cation CH3NH3+ (widely
denoted in the literature with the acronym MA), B is a divalent 3. ST AND NST-QDs
cation (Ge, Sr, Pb, Ca, or Sr), and X is a halogen atom (Cl, Br,
or I). These materials in the form of QDs will exhibit The construction of NP models can be easily achieved by
electronic and optical properties which differ substantially from suitably repeating a fundamental building block along the tree
their bulk properties studied before.19,20 Note that the halide spatial directions. If the building block is the unit cell of the
perovskites with MA are often called hybrid or organometallic material (i.e., A1B1X3), then the atoms of the resulting
halide perovskites in literature. The hybrid halide perovskites structures satisfy the same ratios (i.e., 1/1/3) with the
with Ge, Sn, and especially Pb are among the most studied corresponding bulk material. In this context, the NPs derived
perovskites in literature both experimentally and theoretically by repeating this building block two times in each direction are
because of their potential use in photovoltaic applications.3 denoted as 2 × 2 × 2, whereas when repeated three times, they
The article is organized as follows: in Section 2, we present are labeled 3 × 3 × 3. The size of the resulting nanocrystals is
the computational details. In Section 3, we analyze the approximately ∼1.0 and 1.5 nm. This type of QDs (which in
structure of both the stoichiometric (ST) and nonstoichio- the present work are labeled ST), although build upon the
metric (NST) QDs studied. In Section 4, the electronic and same principle as the bulk, exhibits a surface local environment
optical properties of the systems under study are presented, which is structurally very different from the bulk. In particular,
and finally, in Section 5, we summarize our conclusions. the BX6 octahedra, which constitute the backbone of the
material, are truncated, exposing either the A or the B cations
in three out of six facets of the cuboid QD. This observation
2. OUTLINE OF CALCULATIONS led us to an alternative route of QD construction, based on the
All calculations were performed in the framework of density principle that the BX6 octahedra should not be truncated.
functional theory (DFT) using the generalized gradient Apparently, this leads to violation of the 1/1/3 ratio of the
approximation (GGA) functional of Perdew−Burke−Ernzer- bulk, and consequently, it results in charged NPs (which in the
hof21 (PBE) and its hybrid analogue, to account in a more present work are labeled NST). The charge problem can be
accurate way for the exchange energy, PBE0.22 The initial tackled in two different ways. The first is based on starting with
structures were optimized at the PBE level of theory, which is NPs with full surface coverage by A cations and removing a
known to predict accurately the structural properties, but suitable number of them (i.e., create vacancies) in order to
significantly underestimates the band gap. The resulting achieve neutrality. The second way is based on excluding all
geometries were then used to calculate the reliable values of the surface cations as it might be argued that they are relatively
the single-particle HOMO−LUMO gap with the help of the loosely bound to the BX6 octahedra skeleton and as a result
PBE0 functional. The optical properties were calculated on top they are easier to desorb. In this case, the excessive charge can
of the PBE0 results, using the time-dependent DFT be compensated by a suitable surface passivation. In the
(TDDFT). In particular, we performed TDDFT calculations present manuscript, we have considered only the second type
in both random-phase approximation and Tamm−Dancoff of NST QDs. Their passivation is achieved by pseudohy-
approximations, while we also employed the simplified drogens (in order to maintain neutrality) and the surface A
TDDFT (sTDDFT) approximation of Grimme,23,24 which is cations are considered desorbed. The size of the resulting 2 ×
most suitable for larger systems where traditional TDDFT 2 × 2 and 3 × 3 × 3 NST QDs is approximately ∼1.2 and 1.9
becomes prohibitive. It should be clarified that in the context nm, respectively. The two distinct types of QDs are graphically
of the present manuscript, we consider as optical gap, the presented in Figure 1.
18918 DOI: 10.1021/acsomega.8b02525
ACS Omega 2018, 3, 18917−18924
ACS Omega Article

Table 1. HOMO−LUMO and Optical Gap of CsPbX3 (X = Cl, Br, I) NPsa


cuboid 2 × 2 × 2 cuboid 3 × 3 × 3
PBE PBE0 PBE PBE0
system H−L gap (eV) optical gap (eV) H−L gap (eV) optical gap (eV) H−L gap (eV) optical gap (eV) H−L gap (eV) optical gap (eV)
ST
Cs8Pb8X24 Cs27Pb27X81
X = Cl 2.67 2.67 4.23 3.94 2.30 2.30 3.81 3.60
X = Br 2.56 2.56 4.09 3.70 2.13 2.13 3.60 3.42
X=I 1.92 1.96 3.28 2.88 1.42 1.42 2.64 2.42
NST
Cs1Pb8X36:H24 Cs8Pb27X108:H54
X = Cl 3.90 3.97 5.80 5.18 3.49 3.49 5.06 4.60
X = Br 2.78 2.80 4.66 4.05 2.51 2.51 4.25 3.78
X=I 1.70 1.72 3.20 2.66 1.20 1.20 2.72 2.45
a
The calculations have been performed with the PBE and PBE0 functionals and the def2-SVP basis set. Bulk values (experimental and theoretical)
can be found in ref 20.

The first part of calculations was focused on CsPbX3, (X = long-chain molecules. As a result, the variation of the angles
Cl, Br, I) NPs, which consist of a much easier system to study B−X−R is more or less hindered because of steric interactions
(compared to MA analogues) and the relevant literature is and the shape of the QDs is stabilized and maintained. This
more abundant. In this way, the theoretical approximations combination of hydrogen passivation and angle freezing
and methodological choices could be rigorously evaluated, somehow imitates the passivating and stabilizing effect that
offering a useful insight with regard to the study of the more the molecular ligands have on the structure, while in the same
complicated case of MA containing NPs. Moreover, it was a time makes the computational cost of the calculations
first step toward examining different key morphological affordable.
characteristics of the QDs such as composition and surface Right after the evaluation of the CsPbX3, (X = Cl, Br, I)
passivation and allowed us to consider explicitly the two results, we turned our focus on the ABX3 systems with A =
aforementioned categories of QDs, avoiding the computational CH3NH3+ obeying the same building principles as above. This
difficulties introduced by the presence of methylammonium as time, the B cation was set to B = Sn, Pb, Ge, Ca, Sr, while X =
cation A. In particular, the NPs are free of the extra degrees of Cl, Br, I. The NST structures have a total of 52 (+24
freedom, which are introduced by the presence of CH3NH3+ passivating hydrogens) atoms (size 2 × 2 × 2) and 199 (+54
(giving huge optimization difficulties). Especially for the case passivating hydrogens) atoms (size 3 × 3 × 3). The ST NPs
of ST QDs, where the cations A are on the surface of the dot, if consist of 96 and 324 atoms, respectively. It should be
A = CH3NH3+, then the convergence of geometric relaxation is emphasized that because of the very flat potential energy
almost impossible. This is a manifestation of a very flat surface, which is a consequence of the loose binding of surface
potential energy surface because of the fact that CH3NH3+ is CH3NH3+ cations, the convergence of geometry optimizations
weakly bound to the backbone structure and it can almost is very difficult. For this reason, it was considered necessary to
freely rotate relative to it. On the contrary, if cation A is a terminate the optimization procedure when the energy reached
single atom, these problems are vanished and the geometric a very flat and horizontal plateau. Apparently, this practice
relaxation is straightforward. affects the quality of specific results by introducing some small
As already mentioned, the NST QDs violate the bulk atomic error bars, which are estimated to be smaller than 0.1 eV. This
ratios and require suitable surface passivation in order to estimation is consistent with the molecular dynamics (MD)
remain neutral. If this step is omitted, then the resulting NPs calculations of ref 14 in which the sampling of HOMO−
exhibit self-consistent field convergence problems because of LUMO gap during the MD steps led to this value.
the existence of multiple-gap states, giving a false impression of The sizes of the NPs considered are approximately ∼0.9−1.0
metallic behavior. The passivation is achieved by capping the and ∼1.4−1.6 nm for the 2 × 2 × 2 and 3 × 3 × 3 ST QDs,
halide edges of the cubes with pseudohydrogen atoms (Figure respectively, and ∼1.1−1.3 and ∼1.8−2.0 nm for the
1b), leading to chemically and structurally stable nanostruc- corresponding NST.
tures. Experimentally, this is usually accomplished with long
molecular species such as octylamine.10,13 Including such large
4. RESULTS AND DISCUSSION
molecules in our calculations would make the computational
cost of the passivants even larger than the one for the actual 4.1. CsPbX3 (X = Cl, Br, I) NPs. The comparison of the
QD. As a result, we adopted the solution of passivating results between ST and hydrogen-passivated NST CsPbX3
pseudohydrogen atoms, with their atomic numbers set to 0.792 QDs gives some interesting insight with regard to two
and 0.85185 for the 2 × 2 × 2 (24 hydrogen atoms) and 3 × 3 interconnected but distinct aspects of the subject. Primarily,
× 3 (54 hydrogen atoms) QDs, respectively. On top of this, an it gives evidence on whether the adopted hydrogen passivation
additional geometric constrain was imposed on them. scheme is reliable, when the focus is on the size dependence of
Although their bond lengths were allowed to change on the the HOMO−LUMO and optical gaps of the related materials,
course of geometric relaxation, the bond angles were kept and secondarily, it serves as an initial step toward under-
fixed. This can be rationalized if one considers that in reality, standing possible deviations from the expected size depend-
the passivation is achieved by ligands which, in most cases, are ence due to surface manipulation of the QDs.
18919 DOI: 10.1021/acsomega.8b02525
ACS Omega 2018, 3, 18917−18924
ACS Omega Article

Figure 2. Comparative DOS plot of ST and NST CsPbCl3 and MAPbCl3 cuboids of size (a) 2 × 2 × 2 and (b) 3 × 3 × 3.

The first straightforward observation which can be made by 2.18 + 1.55/x0.82. For the NST CsPbBr3 dots, application of
a simple inspection of the HOMO−LUMO and optical gap this formula to sizes relevant to our present calculations (2 × 2
data compiled in Table 1 is related to the performance of the × 2 and 3 × 3 × 3) gives gap values of 3.49 and 3.11 eV,
two different functionals. As expected, there is a systematic respectively. Similarly, the corresponding values for the ST
difference between the PBE and PBE0 values, which can be dots (2 × 2 × 2 and 3 × 3 × 3) are 3.83 and 3.28 eV. These
readily attributed to the well-known tendency of the pure GGA extrapolated values, although rather uncertain, are consistent
functionals to underestimate the gaps. Moreover, it can be seen with the data of ref 17 and indicate that the PBE results appear
that, in almost all cases (considered here), the PBE HOMO− to significantly underestimate the value of the optical gap (in
LUMO and optical gaps are practically identical. It is worth both ST and NST cases). On the contrary, the predictions of
noting that this picture is somehow unrealistic as it does not the PBE0 calculations, especially for the case of ST QDs, and
account for the exciton binding energy which is defined as the the aforementioned extrapolated values appear to be in very
difference between the single particle and the optical gaps. On good agreement ∼0.15 eV. This is strong evidence that the
the other hand, the PBE0 results appear to be consistent in this adopted computational methodology is capable of producing
respect, predicting exciton binding energies ranging from 210 reliable results. Moreover, it hints that despite the presence of
to 600 meV. truncated octahedra, the ST configurations offer a more
As for the difference between ST and NST QDs, it is evident realistic description of the QDs.
that the results are strongly dependent on the type of halogen. In all CsPbX3 NPs studied, the PBE0 exciton binding energy
For X = Cl, the NST NPs exhibit a significantly larger optical (which is considered as the difference between the single-
gap, while as we go to heavier halogen, the difference gets particle gap and the optical gap) exhibits a theoretically
smaller (or even reversed for X = I). A qualitative consistent trend. For the 2 × 2 × 2 ST QDs, the corresponding
interpretation can be obtained with the help of the X
values are BECsPbCl X
= 290 meV , BECsPbBr = 390 meV , and
3 3
corresponding density of states (DOS) diagrams presented in X
Figure 2a,b. On the left panel of these figures, there are BECsPbI 3
= 400 meV and they shift to smaller values as the size
comparative DOS plots for the ST and NST NPs. In all cases, of the dots increases. In particular, for the 3 × 3 × 3 QDs, the
the presence of passivants results in a shift of valence states red shift is as large as 80, 210, and 180 meV, respectively. A
which are close to HOMO, deeper into the valence band. similar picture (but with larger values) is also observed for the
Taking into consideration that these states are mostly halogen- c a s e o n N S T d o t s , BEMAPbCl X
= 620 meV ,
3
related, it might be inferred that this is the result of a strong X X
X−H bond. The implication of the presence of such strong BEMAPbBr 3
= 610 meV , and BEMAPbI 3
= 540 meV with the
bonds between the surface halogens and the passivating corresponding red shifts of 160, 140, and 270 meV.
pseudohydrogens indirectly affects the position of the low- 4.2. MABX3 (X = Cl, Br, I) NPs. The case of MABX3 (B =
energy conduction band (CB) states. As these CB states are Sn, Pb, Ge, Ca, Sr/X = Cl, Br, I) NPs was a much more
mostly related to the B cation through the antibonding challenging task mainly because of the difficulties described in
interaction B−X, their shift toward lower energies is indicative the previous paragraph. Contrary to what happened in cesium-
of a weaker B−X bonding (which is the consequence of the containing dots, the relaxation procedure (especially for ST
strong X−H bond). These qualitative arguments are also NPs) leads to significant distortions of the initial geometries,
supported by a Mayer bond order analysis. which are more pronounced in the smaller QDs. As expected,
As the size of the studied QDs is quite small, comparison the BX6 octahedra, apart from being truncated in three out of
with experimental data can be attempted by extrapolating six facets of the QD, they also deviate from the ideal symmetric
existing experimental data to the current sizes. In particular, if structure. The CH3NH3+ species exhibit a wide range of
the experimental data of Figure 3b of ref 5 (for CsPbBr3) are orientations, which depend heavily on their relative position
fitted in a function of the form y = a + b/xc, the outcome is y = within the dot. A general trend which (more or less) appears to
18920 DOI: 10.1021/acsomega.8b02525
ACS Omega 2018, 3, 18917−18924
ACS Omega Article

be present in all cases is that the surface CH3NH3+ moieties Sn, Ge QDs and becomes stronger as the halogen type gets
seem to adopt an orientation with their NH3 part toward the heavier. For the case of B = Ca, Sr dots, the picture changes
surface of the cube and the CH3 group directed outward. drastically. Although the tendency of the band edges to shift to
The electronic structure of ST and NST QDs of the most lower energies as the halogen type gets heavier remains, the
commonly used MAPbX3 materials is presented with the help ordering with regard to the position of the ST band edges is
of DOS diagrams, in the right panes of Figure 2a,b (for the two reversed. The complete picture of the gap variations for both
different QD sizes). Once again, the presence of pseudohy- the ST and NST QDs can be conveniently realized with the
drogen passivation seems to shift the band edges toward lower help of Figure 4, which makes the existing trends apparent (the
energies. The source of this shift appears to be the same as the data are taken from Tables 2 and 3).
one described for the case of CsPbX3 NPs. The related gaps The absorption spectra presented in Figure 5 have been
are presented in Tables 2 and 3. Similar diagrams for all the calculated in two steps. At first, the lowest few spin allowed
transitions are calculated at the PBE0/TDDFT/def2-SVP level
Table 2. HOMO−LUMO and Optical Gap Values for NST of theory giving the reported optical gap. Then, a few hundreds
Perovskite QDsa of states are calculated with Grimme’s sTDDFT method.24
The absorption spectrum is created by introducing a suitable
2 × 2 × 2 QD 3 × 3 × 3 QD
Lorentzian broadening to all excited states, while in the same
PBE PBE0 PBE PBE0 time, the sTDDFT results are properly shifted so that the
H−L H−L optical H−L H−L optical lowest transition coincides with the corresponding PBE0/
gap gap gap gap gap gap TDDFT/def2-SVP one.
material (eV) (eV) (eV) (eV) (eV) (eV)
If one attempts to compare with experimental results in a
MAPbCl3 3.69 5.64 4.94 3.00 4.51 4.16 fashion similar to the one adopted for the case of Cs containing
MAPbBr3 2.77 4.57 3.87 2.35 3.96 3.57 NPs (i.e., fit the experimental data and extrapolate to current
MAPbI3 1.71 3.23 2.62 1.46 2.93 2.52 sizes), several inconsistencies may be observed. For example, it
MASnCl3 3.38 4.87 4.35 2.21 3.52 3.17 is not uncommon that the reported gaps of similar sizes, but
MASnBr3 2.69 4.27 3.72 1.93 3.14 2.79 from different studies, do not suit well in a common graph. If,
MASnI3 1.55 3.05 2.46 1.19 2.62 2.29 however, we use a single manuscript data (e.g., the
MAGeCl3 3.15 4.74 4.18 2.13 3.50 3.06 experimental results from Zhenfu et al.29) and fit them to
MAGeBr3 2.10 3.86 3.16 1.59 3.02 2.66 the usual formula y = a + b/xc, the outcome is y = 2.61 + 2.36/
MAGeI3 1.10 2.50 1.88 0.72 2.04 1.69
x1.82. If this is extrapolated to current sizes, then for the case of
MACaCl3 4.39 6.48 5.90 2.79 4.84 4.60
ST 3 × 3 × 3 MAPbBr3 NPs, one gets the value of ∼3.74 eV,
MACaBr3 3.69 5.61 5.04 2.38 4.25 4.02
whereas for the NST ones, the value is ∼3.34 eV. These
MACaI3 2.59 4.27 3.71 1.64 3.26 3.04
extrapolated estimations are consistent with both the ST and
MASrCl3 4.49 6.58 6.01 2.87 4.94 4.71
NST theoretical predictions (3.71 and 3.57 eV). However, the
MASrBr3 3.78 5.72 5.16 2.51 4.43 4.19
comparison highly favors the case of ST QDs. Moreover, if one
MASrI3 2.80 4.52 3.98 2.00 3.69 3.48
a
focuses on the PL energy differences observed29 in halogen
Bulk values (experimental and theoretical) can be found in ref 19. replacement (i.e., Cl → Br → I), they are completely
consistent with the present theoretical values which are
Table 3. HOMO−LUMO and Optical Gap Values for ST graphically presented in Figure 4. In particular, the
Perovskite QDsa experimental values are ΔE(PL)exp Cl→Br = 0.63 eV, ΔE(PL)Br→I
exp

2 × 2 × 2 QD 3 × 3 × 3 QD = 0.66 eV, and ΔE(PL)Cl→I = 1.29 eV, where the current


exp

theoretical predictions for the ST QDs are ΔEth Cl→Br = 0.57 eV,
PBE0 PBE0
ΔEthBr→I = 0.74 eV, and ΔECl→I = 1.31 eV. The corresponding
th
H−L gap optical gap H−L gap optical gap data of the NST NPs exhibit significant deviations from the
material (eV) (eV) (eV) (eV)
experimental measurement. At this point, it should be clarified
MAPbCl3 4.94 4.34 4.59 4.28 that our theoretical prediction corresponds to absorption
MAPbBr3 4.20 3.67 4.04 3.71 energies (not PL). However, it is experimentally verified that
MAPbI3 3.32 2.86 3.22 2.97 the Stokes shift between absorption and emission is quite small
MASnCl3 4.08 3.61 3.55 3.28 (due to direct transitions). As a result, the aforementioned
MASnBr3 3.44 2.97 3.12 2.83 comparison of energy differences might be considered valid. It
MASnI3 2.47 2.06 2.08 1.86 is worth noting that once again the comparison with
MAGeCl3 3.00 2.54 2.03 1.76
experimental data reveals that the ST NPs produce results
MAGeBr3 3.50 2.95 1.45 1.24
which are closer to experiments.
MAGeI3 3.54 2.90 0.90 0.68
For the cases of MAPbX3 and MASnX3 NPs, there is a clear
MACaCl3 5.86 5.44 5.68 5.47
trend that the NST QDs have stronger transitions. The
MACaBr3 5.36 4.92 4.46 4.17
MACaX3 and MASrX3 QDs exhibit practically vanishing
MACaI3 4.78 4.30 3.11 2.91
oscillator strengths. At this point, it should be emphasized
MASrCl3 5.74 5.19 4.91 4.67
that in energies well above the optical gap, the height of the
MASrBr3 5.15 4.67 4.47 4.23
peaks is dictated by the DOS rather than the strength of
MASrI3 4.58 4.14 4.05 3.82
a
individual transitions. As a means to shed light on this
Bulk values (experimental and theoretical) can be found in ref 19. difference, we have plotted the orbital-resolved projected DOS
(PDOS) diagrams for two representative NPs along with a
NPs considered (but only for the 3 × 3 × 3 size) are plotted in graphic representation of their HOMO and LUMO orbitals in
Figure 3. The red shift of the band edges is clear in all B = Pb, Figure 6. The HOMO, HOMO − 1, and so forth states of
18921 DOI: 10.1021/acsomega.8b02525
ACS Omega 2018, 3, 18917−18924
ACS Omega Article

Figure 3. Comparative DOS plot of an ST and a NST CH3NH3AX3 (A = Sn, Ge, Ca, Sr and X = Cl, Br, I) cuboid of size 3 × 3 × 3.

Figure 4. Variation of the HOMO−LUMO and optical gaps as a function of the anion atom type of NST QDs. The dotted lines correspond to the
optical gap. The gray circle indicates unstable NPs which were completely deformed by the relaxation procedure.

MAPbCl3 are mainly Cl(p) states while the LUMO, LUMO + ultrasmall perovskite QDs having the chemical formula ABX3,
1, and so forth are Pb(p). Additionally, a composition analysis where A is either Cs or the methylammonium cation CH3NH3
with the help of Multiwfn code30 revealed that the specific (MA), B is a divalent cation (Ge, Sr, Pb, Ca, or Sr), and X is a
states are quite extended, as is evident in their three- halogen atom (Cl, Br, or I). For every different material and
dimensional plots of Figure 6. On the contrary, the size examined, two types of dots were considered (ST and
corresponding states of the MACaCl3 QDs are Cl(p) and pseudohydrogen-passivated NST). In all cases, the single-
Ca(s), whereas on the same time, the wave functions exhibit a particle gap, the optical gap, and the absorption spectra were
significant localization on specific atoms (see Figure 6). This systematically calculated and compared. It is found that the
completely different picture is reflected on the shape and gaps are always a decreasing function of the atomic number of
details of the corresponding absorption spectra. As expected, in the halogen. The surface passivation by capping the halogen
all ST QDs, the frontier orbitals are mainly distributed on the atoms results in a red shift of both the valence and CB edges
surfaces of the cuboids. This allows for property manipulation (HOMO and LUMO orbitals) as compared to unpassivated
through appropriate surface treatment. ST dots for all QDs (although NPs with Ge as a B cation seem
to suffer from stability issues). The NPs which contain alkaline
5. SUMMARY AND CONCLUSIONS earth metals exhibit a different behavior with respect to the
In conclusion, we have performed extensive ground-state DFT ones containing one of the isovalent Ge, Pb, and Sn. The
and excited-state TDDFT calculations on a large number of absorption spectra of Pb and Sn containing QDs exhibit strong
18922 DOI: 10.1021/acsomega.8b02525
ACS Omega 2018, 3, 18917−18924
ACS Omega Article

Figure 5. Absorption spectra at the PBE0/sTDDFT/def2-SVP level of theory. The results have been properly shifted so that the lowest transition
coincides with the corresponding PBE0/TDDFT/def2-SVP result.

Our study highlights the potential as well as the challenges


and issues regarding nanostructured halide perovskite materials
and paves the way for future theoretical and experimental work
on these materials.

■ AUTHOR INFORMATION
Corresponding Authors
*E-mail: garoufal@upatras.gr (C.S.G.).
*E-mail: galanakis@upatras.gr (I.G.).
*E-mail: bask@upatras.gr (S.B.).
ORCID
Sotirios Baskoutas: 0000-0003-2782-3501
Notes
The authors declare no competing financial interest.

■ ACKNOWLEDGMENTS
The authors would like to thank Dr. Th. Dimopoulos for
fruitful discussions. A.K., I.G., and S.B. acknowledge the
financial support from the project PERMASOL (FFG project
number: 848929).

■ REFERENCES
(1) Yang, L.; Barrows, A. T.; Lidzey, D. G.; Wang, T. Rep. Prog. Phys.
2016, 79, 026501.
(2) Chen, Q.; De Marco, N.; Yang, Y.; Song, T.-B.; Chen, C.-C.;
Zhao, H.; Hong, Z.; Zhou, H.; Yang, Y. Nano Today 2015, 10, 355−
Figure 6. Atom orbital-resolved PDOS of two representative QDs. 396.
The HOMO and LUMO orbitals are also plotted as a means to (3) Hoefler, S. F.; Trimmel, G.; Rath, T. Monatsh. Chem. 2017, 148,
qualitatively point the differences of the two systems. 795−826.
(4) Shekhirev, M.; Goza, J.; Teeter, J. D.; Lipatov, A.; Sinitskii, A. J.
Chem. Educ. 2017, 94, 1150−1156.
transitions, which are even more enhanced for the case of (5) Hou, J.; Cao, S.; Wu, Y.; Gao, Z.; Liang, F.; Sun, Y.; Lin, Z.; Sun,
pseudohydrogen-passivated NST QDs. On the contrary, Ca L. Chem.Eur. J. 2017, 23, 9481−9485.
and Sr NPs exhibit vanishing oscillator strengths, which are (6) Bohn, B. J.; Tong, Y.; Gramlich, M.; Lai, M. L.; Döblinger, M.;
Wang, K.; Hoye, R. L. Z.; Müller-Buschbaum, P.; Stranks, S. D.;
related to the nature of their frontier orbitals which appear to Urban, A. S.; Polavarapu, L.; Feldmann, J. Nano Lett. 2018, 18, 5231−
be extremely localized. Finally, although the NST QDs with 5238.
the complete octahedra possess a structural configuration (7) Yang, D.; Zou, Y.; Li, P.; Liu, Q.; Wu, L.; Hu, H.; Xu, Y.; Sun, B.;
which is more appealing to the eye, the comparison with Zhang, Q.; Lee, S.-T. Nano Energy 2018, 47, 235−242.
experimental data reveals that the ST QDs exhibit gaps closer (8) Tong, Y.; Fu, M.; Bladt, E.; Huang, H.; Richter, A. F.; Wang, K.;
to the experimental ones. Muller-Buschbaum, P.; Bals, S.; Tamarat, P.; Lounis, B.; Feldmann, J.;

18923 DOI: 10.1021/acsomega.8b02525


ACS Omega 2018, 3, 18917−18924
ACS Omega Article

Polavarapu, L. Angew. Chem., Int. Ed. 57, 16094−16098.


DOI: 10.1002/anie.201810110
(9) Bai, Z.; Zhong, H. Sci. Bull. 2015, 60, 1622−1624.
(10) Zhang, F.; Zhong, H.; Chen, C.; Wu, X.-g.; Hu, X.; Huang, H.;
Han, J.; Zou, B.; Dong, Y. ACS Nano 2015, 9, 4533−4542.
(11) Tanaka, K.; Takahashi, T.; Ban, T.; Kondo, T.; Uchida, K.;
Miura, N. Solid State Commun. 2003, 127, 619−623.
(12) Yang, G.-L.; Zhong, H.-Z. Chin. Chem. Lett. 2016, 27, 1124.
(13) Mali, S. S.; Shim, C. S.; Hong, C. K. NPG Asia Mater. 2015, 7,
No. e208.
(14) Buin, A.; Comin, R.; Ip, A. H.; Sargent, E. H. J. Phys. Chem. C
2015, 119, 13965−13971.
(15) He, J.; Vasenko, A. S.; Long, R.; Prezhdo, O. V. J. Phys. Chem.
Lett. 2018, 9, 1872−1879.
(16) Castañeda, J. A.; Nagamine, G.; Yassitepe, E.; Bonato, L. G.;
Voznyy, O.; Hoogland, S.; Nogueira, A. F.; Sargent, E. H.; Cruz, C. H.
B.; Padilha, L. A. ACS Nano 2016, 10, 8603−8609.
(17) Peng, L.; Geng, J.; Ai, L.; Zhang, Y.; Xie, R.; Yang, W.
Nanotechnology 2016, 27, 335604.
(18) Chen, X.; Peng, L.; Huang, K.; Shi, Z.; Xie, R.; Yang, W. Nano
Res. 2016, 9, 1994−2006.
(19) Koliogiorgos, A.; Baskoutas, S.; Galanakis, I. Comput. Mater. Sci.
2017, 138, 92−98.
(20) Moschou, G.; Koliogiorgos, A.; Galanakis, I. Phys. Status Solidi
A 2018, 215, 1700941.
(21) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1996, 77,
3865−3868.
(22) Adamo, C.; Barone, V. J. Chem. Phys. 1999, 110, 6158−6170.
(23) Grimme, S. J. Chem. Phys. 2013, 138, 244104.
(24) Bannwarth, C.; Grimme, S. Comput. Theoret. Chem. 2014,
1040−1041, 45−53.
(25) Weigend, F.; Ahlrichs, R. Phys. Chem. Chem. Phys. 2005, 7,
3297.
(26) Weigend, F. Phys. Chem. Chem. Phys. 2006, 8, 1057.
(27) Neese, F. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2011, 2, 73−
78.
(28) Neese, F. Wiley Interdiscip. Rev.: Comput. Mol. Sci. 2017, 8,
No. e1327.
(29) Zhenfu, Z.; Liang, J.; Zhihai, W.; Jiong, C.; Miaomiao, Z.; Yafei,
H. J. Mater. Sci. 2018, 53, 15430−15441.
(30) Lu, T.; Chen, F. J. Comp. Chem. 2011, 33, 580−592.

18924 DOI: 10.1021/acsomega.8b02525


ACS Omega 2018, 3, 18917−18924

You might also like