Understanding Trap States in InP and GaP Quantum Dots Through Density Functional Theory

Ezra Alexander [email protected]    Matthias Kick    Alexandra R. McIsaac    Troy Van Voorhis Department of Chemistry, Massachusetts Institute of Technology, Cambridge, Massachusetts 02139, USA
(June 26, 2024)
Abstract

The widespread application of III-V colloidal quantum dots (QDs) as non-toxic, highly tunable emitters is stymied by their high density of trap states. Here, we utilize density functional theory (DFT) to investigate trap state formation in a diverse set of realistically passivated core-only InP and GaP QDs. Through orbital localization techniques, we deconvolute the dense manifold of trap states to allow for detailed assignment of surface defects. We find that the three-coordinate species dominate trap** in III-V QDs and identify features in the geometry and charge environment of trap centers capable of deepening, or sometimes passivating, traps. Furthermore, we observe stark differences in surface reconstruction between InP and GaP, where the more labile InP reconstructs to passivate three-coordinate indium at the cost of distortion elsewhere. These results offer explanations for experimentally observed trap** behavior and suggest new avenues for controlling trap states in III-V QDs.

keywords:
quantum dots, indium phosphide, gallium phosphide, cadmium-free, density functional theory, trap state
{tocentry}
[Uncaptioned image]

For Table of Contents Only

1

Colloidal semiconductor nanocrystals, more commonly known as quantum dots (QDs), have attracted considerable attention as solution-processable materials 1, 2 with highly tunable optical properties 3, 4, 5, 6, 7, 8. They have begun to see use in a wide range of applications including photovoltaics 9, 10, photodetectors 2, 11, LEDs 1, 12, lasing 13, 14, drug delivery 15, 16, biological imaging 17, 18, and quantum computing 19, 20. However, the QDs with the best performance to date 21, 22, 23 are those composed of highly toxic and internationally-restricted cadmium or lead chalcogenides 24, 25, 26, making the development of a non-toxic alternative material with equally strong optical properties necessary for safe widespread commercialization. III-V QDs, namely indium phosphide (InP), are promising candidates for this replacement due to their low toxicity 25, 26 and widely tunable emission range 18, 27. Until recently, their implementation has been held back by generally low quantum yields and broad emission line widths relative to their II-VI counterparts 28, 29, 30, 31, 32. These phenomena are often understood to result from a high density of trap states: occupied or virtual electronic states, usually localized on the surface of the QD, with energies between the valence band maximum (VBM) and conduction band minimum (CBM) 31, 33. While recent advances in control over III-V core/shell heterostructures have led to InP QDs with near-unity quantum yields 12, 34, 35, a complete atomistic understanding of the formation and character of surface traps in III-V QDs remains elusive, especially for core-only QDs.

Trap states are not unique to III-V QDs; a vast body of both experimental and theoretical literature discusses trap states in II-VI QDs 33, 36, 37, 38, 39, 40, 41, 42, 43, 44, 45, IV-VI QDs 33, 46, 47, 48, and lead halide perovskite nanocrystals 33, 49, 50, 51. The most widely accepted origin of trap states is under-coordinated surface atoms 31, 38, 50, 51, 52, 53, 54, 55, 56, 57, 42, although in certain systems excess charge 37, 39, imperfect stoichiometry 46, and substitutional defects 58 have also been implicated. For CdSe, several studies employing density functional theory (DFT) have shown that trap states arise primarily from two-coordinate Se atoms (Se-2c) but not Se-3c or any under-coordinated Cd 38, 40, 41. No such consensus has been reached for InP QDs, however. Many studies have implied that hole trap** dominates in InP QDs 36, 53, 56, 57, 58, 59, 60, 61, but there is also considerable evidence for the presence of electron traps, especially in the absence of a core/shell heterostructure 31, 52, 54, 55. This disagreement has been compounded by a relative lack of atomistic ab initio studies of trap states in InP QDs 52, 54, 55, 56, 57, 58, 62, 63, 64, many of which only employ less-accurate GGA functionals. Studies have variously emphasized P-3c traps 57, In-3c traps 55, 64, both In-3c and P-3c traps but with disagreement on their respective depths 52, 54, traps from the two-coordinate species with additional P-3c traps only in tetrahedral geometries 56, as well as studies that find InP QDs with both In-3c and P-3c to be trap free but see traps introduced upon different surface treatments 58, 63. Most of these studies only compute the electronic structure of a single model InP QD, limiting generalizability with respect to shape, size, faceting, and surface passivation. Moreover, very few studies have applied computation to understand trap states in other III-V QDs such as gallium phosphide (GaP) 65, a promising but under-studied emissive material 66, 67.

Here, we use DFT to study a large, diverse set of InP and GaP QDs and draw generalized conclusions on the nature of their trap states and the factors that influence trap depth. The six base QD morphologies studied here are summarized in Figure 1. Several important decisions inform the development of our test set. We focus on core-only QDs, carved from the bulk crystal using a well-established construction procedure 40, 68. We create six starting QDs for both InP and GaP, chosen to represent distinct faceting and synthetically realizable shape69, 70, 71, 72, 73. The four larger models, with diameters of 2-2.5 nm, represent the upper limit of computationally realizable QDs. The two smaller models allow for size extrapolation. Surfaces are passivated with X-type \chF- ligands, representative of the well-established treatment of InP QDs with HF 31, 55. Calculations show that larger halogen ligands create states close to the VBM, potentially interfering with the assignment of trap and bulk states (SI I.I). As all our QDs are cation rich, one can equivalently think of them as stoichiometric \chInP/\chGaP cores passivated by Z-type \chInF3/\chGaF3 ligands74.

Refer to caption
Figure 1: Overview of the six base computational InP QD models used in this study. Each InP structure shown here has a GaP counterpart with the same shape and stoichiometry (SI, Figure S2). The colored inserts to the right of each structure provide a visual guide of the corresponding surface facets. Structures (b,c) and (e,f) can be thought of as extended versions of structures (a) and (d), respectively.

The construction procedure employed for these starting QDs is analogous to the ones used in previous ab initio studies of trap states in III-V QDs (SI I.II), and results in some number of three-coordinate In/Ga and P atoms in all structures 55, 56, 57, 58, 63. The difficulty in creating perfectly four-coordinate III-V model QDs arises from charge-orbital balance, in which the formal charge of each atom must add to the total charge of the system to prevent do** 75. While some experimental evidence suggests QDs must be strictly charge-neutral 76, 77, it has also been shown that strict charge-balance greatly limits possible model III-V QDs, additionally restricting defects one could induce to these QDs to charge neutrality 56. To extend the range of QD shapes and defects available to us without inducing do**, we allow for slight positive charges in our structures 37, 78. The restriction to positive charges serves to avoid exacerbating DFT’s self-interaction error 79 and allows our cation-rich systems to have their charge balanced by fluoride counter-ions in solution 53. We observe no do** in any of our systems and no qualitative difference in geometry or electronic structure between our charged and neutral models (SI I.III).

We diversify our dataset and study the effects of surface reconstruction by creating "defective" QDs out of these starting models. These defective structures are created in a similar manner to previous studies, but our lack of charge neutrality affords us a greater variety of available defects 52, 38, 33, 78, 56, 54, 50, 41. When creating defects in a starting structure, we consider all symmetry-unique removals of a single \chF-, \chP^3-, \chInF_x, and \chInP (SI I.IV). This procedure results in a total library of 160 InP and GaP QDs for consideration.

We then compute the ground state electronic structure of each QD using PBE0, as hybrid functionals are necessary for the accurate reproduction of band gaps 80, 81. Comprehensive identification of trap states requires the prior identification of the first bulk states, i.e. the VBM and CBM. This is made challenging by the dense manifold of intermediately localized states near the band edges (Figure 2). Two techniques are used to visualize the band structure of our models. The first is the projected density of states (PDOS), which visualizes contributions from different atomic species to each band as a function of energy (Figure 2a,c). The second is the participation ratio (PR) (SI II.II), which measures the localization of each electronic state (Figure 2b,d). Analysis of the PDOS and PR alone is insufficient to understand the trap states in the QDs studied here for two reasons. First, the starting structures (Figures 1, S2) have trap states before any defects are induced, arising primarily from pre-existing three-coordinate atoms. Second, the trap states in our QDs are generally shallow because of their high degree of surface reconstruction. Thus, almost all QDs in our dataset display a dense quasi-continuum of trap states at both band edges, which obfuscate both new trap states induced by specific defects and the "true" CBM and VBM.

Refer to caption
Figure 2: (a,c) Projected density of states for two fluorine defects in the truncated InP cuboctahedron. Colored lines indicate contributions from different elements. (b,d) Participation ratio for the same two structures. Blue lines represent KS eigenstates that are clearly localized while purple lines represent intermediately delocalized states that become localized upon Pipek-Mizey localization. Green lines represent the first bulk state identified by our algorithm. The orange line represents the highly delocalized quantum-confined S-like state. (e) Pipek-Mizey orbital localization for the VB edge of the truncated InP cuboctahedron. Three states with intermediate localization are mixed with eight clearly localized states to form eleven localized states on different P-3c. Orbitals shown at an isosurface level of 0.03.

A naive analysis of the PR is problematic because localization is not an intrinsic property of DFT Kohn-Sham eigenstates 82, 83. In fact, any linear combination of degenerate eigenfunctions is also a solution to the Kohn-Sham equations, and in many systems these linear combinations will be more delocalized than what experiment and chemical intuition would suggest, especially when the eigenstate spectrum is particularly dense 84. Confronted with this problem, we can utilize orbital localization methods, such as Foster-Boys 85 or Pipek-Mizey 86, to perform unitary transformations on a selected subset of molecular orbitals to maximize their localization. We find applying Pipek-Mizey localization to the band edges of our QDs reveals that many states with intermediate delocalization reduce to linear combinations of clearly localized trap states (Figure 2e). Combined with the observation that “true” bulk states fail to localize into clear surface states, orbital localization gives us a powerful tool to test for the location of the VBM and CBM. Our procedure is described in detail in the Supporting Information (II.III) alongside data highlighting the consistency of our predictions between related structures. As an example, the VBM is chosen to be the highest energy delocalized occupied state without a disproportionate contribution from under-coordinated P, and all occupied states above it in energy should be localizable into clearly surface-bound trap states. All such higher energy states are then taken to be hole traps, with trap depth equal to the difference between their energy and the energy of the VBM. An analogous definition identifies the CBM and associated electron traps.

A complication arises when considering the conduction band due to the intermittent presence of a low-energy, highly-delocalized state which likely corresponds to the 1Se1subscript𝑆𝑒1S_{e}1 italic_S start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT ground exciton state observed in experiment and the S-like envelope state predicted by "particle-in-a-sphere" theories 3, 87, 88, 89. While delocalized, this quantum-confined state cannot be considered the CBM due to its energetic isolation from the quasi-continuous conduction band. Furthermore, the state is present in less than 30% of our structures. Figures 2b,d display two QDs with identical shape and stoichiometry but where the former displays the quantum-confined state, the latter does not. We have found no chemical justification for this intermittency. Nevertheless, our algorithm produces consistent VBMs, CBMs, and trap depths if we exclude the quantum confined state from consideration (SI II.IV). We observe no general qualitative difference between the trap states in QDs with and without the quantum-confined state.

Recently, Snee et al. found that DFT predicts non-singlet ground states for certain halide-passivated model InP QDs - a situation that would have major implications for the magnetic properties of these structures 90. We were unable to demonstrate a similar effect in the structures studied here. For all 12 starting QDs in Figures 1 and S2, we find the singlet to be the lowest energy spin state, as would naively be expected for charge-balanced QDs. We thus conclude that there must be some important geometric differences between our QDs and those of Snee et al. While it would be interesting to understand the nature of those differences, for now we focus on the non-magnetic ground states of these structures.

The defective QD structures reveal an immediate difference in surface reconstruction between InP and GaP (Figure 3a,b). While GaP is relatively rigid with little reconstruction after defect induction, InP reconstructs heavily both in the immediate vicinity of the defect and further afield. Representative of all QDs studied here, we use the smaller tetrahedral QD in Figure 3 as a particularly clear example of this effect. Degree of reconstruction can be quantitatively measured though reorganization energies, which we find to be over three times larger on average in InP than in GaP (SI III.I). We hypothesize the origin of this contrast to be In’s stronger, more ionic bonds with anionic ligands than Ga, due to In’s lower electronegativity. We note that the observed trends in reconstruction still hold when Cl ligands are employed instead of F (SI III.II). These results are supported by prior experimental findings for InGaP, where Ga is found to reside disproportionately at the surface, and In-to-Ga substitution is found to be thermodynamically favorable, increasing QD stability and narrowing X-ray peaks 91, 92.

Refer to caption
Figure 3: InP (a) and GaP (b) converged structures highlighting differences in surface reconstruction after corresponding F and InF defects are created in the smaller tetrahedra. Red colored atoms indicate the metal atom from which a fluorine anion is removed, and cream colored atoms indicate the phosphorus from which InF cations are removed. Distribution of trap depths for all electron (c) and hole (d) traps for all QDs in the dataset combined. Traps from InP and GaP, as well as 3-coordinate and structural traps, are colored separately. Discrete trap depths are broadened by normalized Gaussian functions with RMS width 0.1 eV. Non-trap** 3c atoms are assigned a depth of 0.

The greater reconstruction in InP leads to differences in the depth distribution of electron and hole traps between the two materials, shown in Figures 3c and 3d. Across our data set there are fewer In-3c in InP than there are Ga-3c in GaP, and similar numbers of P-3c in InP and GaP (SI III.I). The cost of the passivation of In-3c is evidenced in the formation of additional trap states localized around distorted In-4c and P-4c atoms, here denoted “structural” traps. While localized with mid-gap energies, these structural traps appear to arise not from under-coordinated surface species but rather from structural deformations caused by extensive surface reconstruction. We do not delve deeply into the nature and origin of these structural traps here; for now, we simply note their presence, even before orbital localization is applied. Despite these additional structural traps, electron traps in InP are generally less deep than those in GaP, with many In-3c being non-trap** or very shallowly trap**. This finding agrees with experimental results for InGaP nanowires, where nonradiative recombination increases with increasing Ga concentration 93. P-based hole traps, on the other hand, have similar depths across InP and GaP QDs (Figure 3d). A structure with each defect clearly labeled can be found in the Supporting Information (III.III).

Close analysis of electron traps in InP and GaP QDs reveals two distinct geometries of the three-coordinate cation, which lead to distinct distributions of trap depths (Figure 4a,b). We designate these geometries as planar, when the cation is coplanar with its coordinated atoms, and pyramidal, when the cation is out of said plane. While both geometries are present in both InP and GaP, planar Ga-3c is far more prevalent relative to pyramidal Ga-3c than planar In-3c is to planar In-3c. This further explains the surface reconstruction in InP, where planar In-3c is more likely to convert to the pyramidal geometry so ligands can bridge to additional In-3c. In both materials, traps arising from the pyramidal geometry are significantly deeper, on average, than those from the planar geometry. In InP this difference is such that most planar In-3c are effectively non-trap**. By separating the two geometries, we see that, averaging over all QDs in our dataset, planar Ga-3c (0.23 eV) forms deeper traps than planar In-3c (0.14 eV), and that pyramidal Ga-3c (0.84 eV) forms deeper traps than pyramidal In-3c (0.52 eV).

Refer to caption
Figure 4: Distribution of trap depths for In-3c (a) and Ga-3c (b) traps for all QDs in the dataset combined, with contributions from planar (blue) and pyramidal (orange) cations displayed separately. Without loss of generality, we only show the depth of the first trap on each 3c cation to better highlight trends. Discrete trap depths are broadened by normalized Gaussian functions with RMS width 0.1 eV. Non-trap** 3c atoms are assigned a depth of 0. (c) Change in LUMO energy when \chInCl3 and \chGaCl3 are interpolated from planar to pyramidal. Plots of initial and final \chInCl3 LUMO orbitals are inset with an isosurface level of 0.06. (d) Cartoon illustrating the difference in electron trap depths between InP and GaP arising from GaP’s larger band gap.

To understand the origin of these differing trap depths, we performed interpolations between pyramidal and planar \chInCl3 and \chGaCl3 (Figure 4c). We find that the differences between pyramidal and planar defects can be understood using simple molecular arguments. As we interpolate from planar to pyramidal (SI IV.I), we observe that the LUMO, which corresponds to the trap state, decreases in energy by around 0.7 eV in both \chInCl3 and \chGaCl3. This is accompanied by a shift of electron density onto the metal in the LUMO, as can be seen in the orbital plots in Figure 4c and through ChELPG charge analysis (SI IV.II) 94. This charge makes the LUMO more lone-pair-like, lowering the energy of the anti-bonding state resulting in a deeper trap. This analysis does not explain the difference between the electron trap depths in InP and GaP, which most likely arise from GaP’s wider band gap 56, 92 where shallow traps in GaP become non-trap** as the CBM decreases in energy (Figure 4d). We find support for this idea through additional interpolations between four-coordinate and three-coordinate \chInCl4 and \chGaCl4 (SI IV.III).

Even accounting for the differing depths of pyramidal and planar traps, the distributions of electron trap depths in Figure 4a,b remain quite broad. These shifts can be explained primarily through two electrostatic effects in the trap centers’ local environment. First, we observe that In-3c and Ga-3c bound to P-3c have shallower traps on average than those not bound to P-3c (Figure 5a,b). This can be understood by recognizing P-3c as having an excess of negative charge, which destabilizes the nearby trap state and raises its energy towards the CBM. A similar effect exists in reverse for hole traps, where P-3c bound to In-3c have shallower traps as the excess positive charge of the In-3c stabilizes the trap state (SI V). The second effect arises from the internal dipole moment of the QD. In both real QDs and our models, any asymmetry will lead to the creation of an internal dipole moment which can shift the depth of trap states, as previously noted for perovskite NCs 51. We show this effect for In-3c and Ga-3c in Figure 5c,d, computing the dipole overlap as the dot product between the dipole moment (pointing toward the positive charge) and the position vector of the trap center. We observe that a positive dipole overlap deepens trap states, whereas a negative dipole overlap almost always makes In-3c and Ga-3c nontrap**. Again, a similar effect exists in reverse for P-3c (SI V).

Refer to caption
Figure 5: (a,b) Trap depth distribution for In-3c and Ga-3c, respectively, across all QDs in the dataset combined with contributions from cations bound to P-3c separated from those not bound to P-3c. (c,d) Trap depth distribution for In-3c and Ga-3c, respectively, across all QDs in the dataset combined with contributions from cations with different dipole overlaps separated. Dipole overlap is calculated as the dot product of the position vector of the trap center with the dipole moment (pointing toward positive charge). A cutoff of (-40,40) is used to define the "near zero" region. Only the depth of the first trap on each 3c cation is shown to highlight trends. Discrete trap depths are broadened by normalized Gaussian functions with RMS width 0.1 eV. Non-trap** 3c atoms are assigned a depth of 0.

Note that we have not included traps from two-coordinate atoms in the above discussion. Such species appear in our dataset, and moreover cause deep traps when they appear (SI VI). However, we find that surface reconstruction in InP is sufficient to passivate most two-coordinate defects when created, and that structures with two-coordinate atoms are highly unstable relative to structures without. We thus conclude that two-coordinate species are at best minor contributors to trap** in real InP and GaP QDs.

In conclusion, we have investigated the prevalence, character, and depth of trap states in InP and GaP QDs using DFT. Our results are kept generalizable through the study of 160 QDs with variable size, shape, and surface defects. Through orbital localization, we deconvolute the dense band edge to identify otherwise evasive trap states including as yet unexplored “structural” traps tied to distorted fully-coordinated atoms. We leverage our dataset to analyze trends in trap depth arising from cation species, cation geometry, and local electrostatic effects which provide avenues for trap control. These results yield important insights into trap states in InP and GaP QDs, as well as informing guidelines for effective simulation of trap states in general QDs. Future directions include detailed investigation of the origin and character of the structural traps observed here, as well as the study of trap states in other III-V materials such as InGaP and III-V core-shell heterostructures. Investigation into the effects of surface oxidation and other impurities on trap state formation, as well as the efficacy of different surface passivation schemes, such as co-passivation with L-type ligands, is necessary to gain a complete picture of trap states in III-V QDs. Finally, it would be instructive to investigate the excited state electronic structure of III-V QDs directly using, for example, time-dependent DFT.

{acknowledgement}

This work was supported by the German Research Foundation (DFG, KI 2558/1-1). We would like to thank the MIT School of Science, MIT Chemistry Department, and the Office of Graduate Education for support of this work through the Dean of Science Fellowship.

{suppinfo}

The Supporting Information is available free of charge at XXX

  • Ligand choice, construction procedure, charge effects, defect induction procedure, computational details, participation ratio, procedure for identification of 1st bulk state, discussion of quantum-confined state, reconstruction energies, reconstruction with Cl ligands, labeled defects, interpolation details, ChELPG charge analysis, interpolation from four-coordinate to three-coordinate, broadening of hole trap depths, two-coordinate traps (PDF)

  • All QD .xyz files (.zip)

  • All trap depths (.xlsx)

References

  • Dai et al. 2014 Dai, X.; Zhang, Z.; **, Y.; Niu, Y.; Cao, H.; Liang, X.; Chen, L.; Wang, J.; Peng, X. Solution-processed, high-performance light-emitting diodes based on quantum dots. Nature 2014, 515, 96–99.
  • García de Arquer et al. 2017 García de Arquer, F. P.; Armin, A.; Meredith, P.; Sargent, E. H. Solution-processed semiconductors for next-generation photodetectors. Nature Reviews Materials 2017, 2, 1–17.
  • Bawendi et al. 1990 Bawendi, M. G.; Steigerwald, M. L.; Brus, L. E. The Quantum Mechanics of Larger Semiconductor Clusters ("Quantum Dots"). Annual Review of Physical Chemistry 1990, 41, 477–496.
  • Murray et al. 1993 Murray, C. B.; Norris, D. J.; Bawendi, M. G. Synthesis and characterization of nearly monodisperse CdE (E = sulfur, selenium, tellurium) semiconductor nanocrystallites. Journal of the American Chemical Society 1993, 115, 8706–8715.
  • Alivisatos 1996 Alivisatos, A. P. Semiconductor Clusters, Nanocrystals, and Quantum Dots. Science 1996, 271, 933–937.
  • Talapin et al. 2010 Talapin, D. V.; Lee, J.-S.; Kovalenko, M. V.; Shevchenko, E. V. Prospects of Colloidal Nanocrystals for Electronic and Optoelectronic Applications. Chemical Reviews 2010, 110, 389–458.
  • Boles et al. 2016 Boles, M. A.; Ling, D.; Hyeon, T.; Talapin, D. V. The surface science of nanocrystals. Nature Materials 2016, 15, 141–153.
  • Kagan et al. 2016 Kagan, C. R.; Lifshitz, E.; Sargent, E. H.; Talapin, D. V. Building devices from colloidal quantum dots. Science 2016, 353, aac5523.
  • Semonin et al. 2011 Semonin, O. E.; Luther, J. M.; Choi, S.; Chen, H.-Y.; Gao, J.; Nozik, A. J.; Beard, M. C. Peak External Photocurrent Quantum Efficiency Exceeding 100% via MEG in a Quantum Dot Solar Cell. Science 2011, 334, 1530–1533.
  • Scalise 2019 Scalise, E. Tailoring the electronic properties of semiconducting nanocrystal-solids. Semiconductor Science and Technology 2019, 35, 013001.
  • Livache et al. 2019 Livache, C.; Martinez, B.; Goubet, N.; Gréboval, C.; Qu, J.; Chu, A.; Royer, S.; Ithurria, S.; Silly, M. G.; Dubertret, B.; Lhuillier, E. A colloidal quantum dot infrared photodetector and its use for intraband detection. Nature Communications 2019, 10, 2125.
  • Won et al. 2019 Won, Y.-H.; Cho, O.; Kim, T.; Chung, D.-Y.; Kim, T.; Chung, H.; Jang, H.; Lee, J.; Kim, D.; Jang, E. Highly efficient and stable InP/ZnSe/ZnS quantum dot light-emitting diodes. Nature 2019, 575, 634–638.
  • Fan et al. 2017 Fan, F. et al. Continuous-wave lasing in colloidal quantum dot solids enabled by facet-selective epitaxy. Nature 2017, 544, 75–79.
  • Park et al. 2021 Park, Y.-S.; Roh, J.; Diroll, B. T.; Schaller, R. D.; Klimov, V. I. Colloidal quantum dot lasers. Nature Reviews Materials 2021, 6, 382–401.
  • Mura et al. 2013 Mura, S.; Nicolas, J.; Couvreur, P. Stimuli-responsive nanocarriers for drug delivery. Nature Materials 2013, 12, 991–1003.
  • Patra et al. 2018 Patra, J. K.; Das, G.; Fraceto, L. F.; Campos, E. V. R.; Rodriguez-Torres, M. d. P.; Acosta-Torres, L. S.; Diaz-Torres, L. A.; Grillo, R.; Swamy, M. K.; Sharma, S.; Habtemariam, S.; Shin, H.-S. Nano based drug delivery systems: recent developments and future prospects. Journal of Nanobiotechnology 2018, 16, 1–33.
  • Medintz et al. 2005 Medintz, I. L.; Uyeda, H. T.; Goldman, E. R.; Mattoussi, H. Quantum dot bioconjugates for imaging, labelling and sensing. Nature Materials 2005, 4, 435–446.
  • Saeboe et al. 2021 Saeboe, A. M.; Nikiforov, A. Y.; Toufanian, R.; Kays, J. C.; Chern, M.; Casas, J. P.; Han, K.; Piryatinski, A.; Jones, D.; Dennis, A. M. Extending the Near-Infrared Emission Range of Indium Phosphide Quantum Dots for Multiplexed In Vivo Imaging. Nano Letters 2021, 21, 3271–3279.
  • Ladd et al. 2010 Ladd, T. D.; Jelezko, F.; Laflamme, R.; Nakamura, Y.; Monroe, C.; O’Brien, J. L. Quantum computers. Nature 2010, 464, 45–53.
  • Zajac et al. 2018 Zajac, D. M.; Sigillito, A. J.; Russ, M.; Borjans, F.; Taylor, J. M.; Burkard, G.; Petta, J. R. Resonantly driven CNOT gate for electron spins. Science 2018, 359, 439–442.
  • Dai et al. 2017 Dai, X.; Deng, Y.; Peng, X.; **, Y. Quantum-Dot Light-Emitting Diodes for Large-Area Displays: Towards the Dawn of Commercialization. Advanced Materials 2017, 29, 1607022.
  • Hanifi et al. 2019 Hanifi, D. A.; Bronstein, N. D.; Koscher, B. A.; Nett, Z.; Swabeck, J. K.; Takano, K.; Schwartzberg, A. M.; Maserati, L.; Vandewal, K.; van de Burgt, Y.; Salleo, A.; Alivisatos, A. P. Redefining near-unity luminescence in quantum dots with photothermal threshold quantum yield. Science 2019, 363, 1199–1202.
  • García de Arquer et al. 2021 García de Arquer, F. P.; Talapin, D. V.; Klimov, V. I.; Arakawa, Y.; Bayer, M.; Sargent, E. H. Semiconductor quantum dots: Technological progress and future challenges. Science 2021, 373, eaaz8541.
  • Derfus et al. 2004 Derfus, A. M.; Chan, W. C. W.; Bhatia, S. N. Probing the Cytotoxicity of Semiconductor Quantum Dots. Nano Letters 2004, 4, 11–18.
  • Reiss et al. 2016 Reiss, P.; Carrière, M.; Lincheneau, C.; Vaure, L.; Tamang, S. Synthesis of Semiconductor Nanocrystals, Focusing on Nontoxic and Earth-Abundant Materials. Chemical Reviews 2016, 116, 10731–10819.
  • Allocca et al. 2019 Allocca, M.; Mattera, L.; Bauduin, A.; Miedziak, B.; Moros, M.; De Trizio, L.; Tino, A.; Reiss, P.; Ambrosone, A.; Tortiglione, C. An Integrated Multilevel Analysis Profiling Biosafety and Toxicity Induced by Indium- and Cadmium-Based Quantum Dots in Vivo. Environmental Science & Technology 2019, 53, 3938–3947.
  • Reiss et al. 2009 Reiss, P.; Protière, M.; Li, L. Core/Shell Semiconductor Nanocrystals. Small 2009, 5, 154–168.
  • Cros-Gagneux et al. 2010 Cros-Gagneux, A.; Delpech, F.; Nayral, C.; Cornejo, A.; Coppel, Y.; Chaudret, B. Surface Chemistry of InP Quantum Dots: A Comprehensive Study. Journal of the American Chemical Society 2010, 132, 18147–18157.
  • Cui et al. 2013 Cui, J.; Beyler, A. P.; Marshall, L. F.; Chen, O.; Harris, D. K.; Wanger, D. D.; Brokmann, X.; Bawendi, M. G. Direct probe of spectral inhomogeneity reveals synthetic tunability of single-nanocrystal spectral linewidths. Nature Chemistry 2013, 5, 602–606.
  • Tamang et al. 2016 Tamang, S.; Lincheneau, C.; Hermans, Y.; Jeong, S.; Reiss, P. Chemistry of InP Nanocrystal Syntheses. Chemistry of Materials 2016, 28, 2491–2506.
  • Hughes et al. 2019 Hughes, K. E.; Stein, J. L.; Friedfeld, M. R.; Cossairt, B. M.; Gamelin, D. R. Effects of Surface Chemistry on the Photophysics of Colloidal InP Nanocrystals. ACS Nano 2019, 13, 14198–14207.
  • Jang et al. 2020 Jang, E.; Kim, Y.; Won, Y.-H.; Jang, H.; Choi, S.-M. Environmentally Friendly InP-Based Quantum Dots for Efficient Wide Color Gamut Displays. ACS Energy Letters 2020, 5, 1316–1327.
  • Giansante and Infante 2017 Giansante, C.; Infante, I. Surface Traps in Colloidal Quantum Dots: A Combined Experimental and Theoretical Perspective. The Journal of Physical Chemistry Letters 2017, 8, 5209–5215.
  • Van Avermaet et al. 2022 Van Avermaet, H.; Schiettecatte, P.; Hinz, S.; Giordano, L.; Ferrari, F.; Nayral, C.; Delpech, F.; Maultzsch, J.; Lange, H.; Hens, Z. Full-Spectrum InP-Based Quantum Dots with Near-Unity Photoluminescence Quantum Efficiency. ACS Nano 2022, 16, 9701–9712.
  • Li et al. 2022 Li, H.; Zhang, W.; Bian, Y.; Ahn, T. K.; Shen, H.; Ji, B. ZnF2-Assisted Synthesis of Highly Luminescent InP/ZnSe/ZnS Quantum Dots for Efficient and Stable Electroluminescence. Nano Letters 2022, 22, 4067–4073.
  • Kirkwood et al. 2018 Kirkwood, N.; Monchen, J. O. V.; Crisp, R. W.; Grimaldi, G.; Bergstein, H. A. C.; du Fossé, I.; van der Stam, W.; Infante, I.; Houtepen, A. J. Finding and Fixing Traps in II–VI and III–V Colloidal Quantum Dots: The Importance of Z-Type Ligand Passivation. Journal of the American Chemical Society 2018, 140, 15712–15723.
  • Voznyy et al. 2013 Voznyy, O.; Thon, S. M.; Ip, A. H.; Sargent, E. H. Dynamic Trap Formation and Elimination in Colloidal Quantum Dots. The Journal of Physical Chemistry Letters 2013, 4, 987–992, Publisher: American Chemical Society.
  • Houtepen et al. 2017 Houtepen, A. J.; Hens, Z.; Owen, J. S.; Infante, I. On the Origin of Surface Traps in Colloidal II–VI Semiconductor Nanocrystals. Chemistry of Materials 2017, 29, 752–761.
  • du Fossé et al. 2019 du Fossé, I.; ten Brinck, S.; Infante, I.; Houtepen, A. J. Role of Surface Reduction in the Formation of Traps in n-Doped II–VI Semiconductor Nanocrystals: How to Charge without Reducing the Surface. Chemistry of Materials 2019, 31, 4575–4583.
  • Goldzak et al. 2021 Goldzak, T.; McIsaac, A. R.; Van Voorhis, T. Colloidal CdSe nanocrystals are inherently defective. Nature Communications 2021, 12, 890.
  • McIsaac et al. 2023 McIsaac, A. R.; Goldzak, T.; Van Voorhis, T. It Is a Trap!: The Effect of Self-Healing of Surface Defects on the Excited States of CdSe Nanocrystals. The Journal of Physical Chemistry Letters 2023, 14, 1174–1181, Publisher: American Chemical Society.
  • Kilina et al. 2009 Kilina, S.; Ivanov, S.; Tretiak, S. Effect of Surface Ligands on Optical and Electronic Spectra of Semiconductor Nanoclusters. Journal of the American Chemical Society 2009, 131, 7717–7726.
  • Wei et al. 2012 Wei, H. H.-Y.; Evans, C. M.; Swartz, B. D.; Neukirch, A. J.; Young, J.; Prezhdo, O. V.; Krauss, T. D. Colloidal Semiconductor Quantum Dots with Tunable Surface Composition. Nano Letters 2012, 12, 4465–4471.
  • Elward and Chakraborty 2013 Elward, J. M.; Chakraborty, A. Effect of Dot Size on Exciton Binding Energy and Electron–Hole Recombination Probability in CdSe Quantum Dots. Journal of Chemical Theory and Computation 2013, 9, 4351–4359.
  • Califano et al. 2005 Califano, M.; Franceschetti, A.; Zunger, A. Temperature Dependence of Excitonic Radiative Decay in CdSe Quantum Dots: The Role of Surface Hole Traps. Nano Letters 2005, 5, 2360–2364.
  • Kim et al. 2013 Kim, D.; Kim, D.-H.; Lee, J.-H.; Grossman, J. C. Impact of Stoichiometry on the Electronic Structure of PbS Quantum Dots. Physical Review Letters 2013, 110, 196802.
  • Chuang et al. 2015 Chuang, C.-H. M.; Maurano, A.; Brandt, R. E.; Hwang, G. W.; Jean, J.; Buonassisi, T.; Bulović, V.; Bawendi, M. G. Open-Circuit Voltage Deficit, Radiative Sub-Bandgap States, and Prospects in Quantum Dot Solar Cells. Nano Letters 2015, 15, 3286–3294.
  • Vörös et al. 2017 Vörös, M.; Brawand, N. P.; Galli, G. Hydrogen Treatment as a Detergent of Electronic Trap States in Lead Chalcogenide Nanoparticles. Chemistry of Materials 2017, 29, 2485–2493.
  • Shao et al. 2014 Shao, Y.; Xiao, Z.; Bi, C.; Yuan, Y.; Huang, J. Origin and elimination of photocurrent hysteresis by fullerene passivation in CH3NH3PbI3 planar heterojunction solar cells. Nature Communications 2014, 5, 5784.
  • Nenon et al. 2018 Nenon, D. P.; Pressler, K.; Kang, J.; Koscher, B. A.; Olshansky, J. H.; Osowiecki, W. T.; Koc, M. A.; Wang, L.-W.; Alivisatos, A. P. Design Principles for Trap-Free CsPbX3 Nanocrystals: Enumerating and Eliminating Surface Halide Vacancies with Softer Lewis Bases. Journal of the American Chemical Society 2018, 140, 17760–17772.
  • Du Fossé et al. 2022 Du Fossé, I.; Mulder, J. T.; Almeida, G.; Spruit, A. G. M.; Infante, I.; Grozema, F. C.; Houtepen, A. J. Limits of Defect Tolerance in Perovskite Nanocrystals: Effect of Local Electrostatic Potential on Trap States. Journal of the American Chemical Society 2022, 144, 11059–11063.
  • Fu and Zunger 1997 Fu, H.; Zunger, A. InP quantum dots: Electronic structure, surface effects, and the redshifted emission. Physical Review B 1997, 56, 1496–1508.
  • Stein et al. 2016 Stein, J. L.; Mader, E. A.; Cossairt, B. M. Luminescent InP Quantum Dots with Tunable Emission by Post-Synthetic Modification with Lewis Acids. The Journal of Physical Chemistry Letters 2016, 7, 1315–1320.
  • Cho et al. 2018 Cho, E.; Kim, T.; Choi, S.-m.; Jang, H.; Min, K.; Jang, E. Optical Characteristics of the Surface Defects in InP Colloidal Quantum Dots for Highly Efficient Light-Emitting Applications. ACS Applied Nano Materials 2018, 1, 7106–7114.
  • Kim et al. 2018 Kim, T.-G.; Zherebetskyy, D.; Bekenstein, Y.; Oh, M. H.; Wang, L.-W.; Jang, E.; Alivisatos, A. P. Trap Passivation in Indium-Based Quantum Dots through Surface Fluorination: Mechanism and Applications. ACS Nano 2018, 12, 11529–11540.
  • Dümbgen et al. 2021 Dümbgen, K. C.; Zito, J.; Infante, I.; Hens, Z. Shape, Electronic Structure, and Trap States in Indium Phosphide Quantum Dots. Chemistry of Materials 2021, 33, 6885–6896.
  • Dümbgen et al. 2023 Dümbgen, K. C.; Leemans, J.; De Roo, V.; Minjauw, M.; Detavernier, C.; Hens, Z. Surface Chemistry of InP Quantum Dots, Amine–Halide Co-Passivation, and Binding of Z-Type Ligands. Chemistry of Materials 2023, 35, 1037–1046.
  • Janke et al. 2018 Janke, E. M.; Williams, N. E.; She, C.; Zherebetskyy, D.; Hudson, M. H.; Wang, L.; Gosztola, D. J.; Schaller, R. D.; Lee, B.; Sun, C.; Engel, G. S.; Talapin, D. V. Origin of Broad Emission Spectra in InP Quantum Dots: Contributions from Structural and Electronic Disorder. Journal of the American Chemical Society 2018, 140, 15791–15803.
  • Calvin et al. 2020 Calvin, J. J.; Swabeck, J. K.; Sedlak, A. B.; Kim, Y.; Jang, E.; Alivisatos, A. P. Thermodynamic Investigation of Increased Luminescence in Indium Phosphide Quantum Dots by Treatment with Metal Halide Salts. Journal of the American Chemical Society 2020, 142, 18897–18906.
  • Richter et al. 2019 Richter, A. F.; Binder, M.; Bohn, B. J.; Grumbach, N.; Neyshtadt, S.; Urban, A. S.; Feldmann, J. Fast Electron and Slow Hole Relaxation in InP-Based Colloidal Quantum Dots. ACS Nano 2019, 13, 14408–14415.
  • Enright et al. 2022 Enright, M. J. et al. Role of Atomic Structure on Exciton Dynamics and Photoluminescence in NIR Emissive InAs/InP/ZnSe Quantum Dots. The Journal of Physical Chemistry C 2022, 126, 7576–7587.
  • Park et al. 2021 Park, N.; Eagle, F. W.; DeLarme, A. J.; Monahan, M.; LoCurto, T.; Beck, R.; Li, X.; Cossairt, B. M. Tuning the interfacial stoichiometry of InP core and InP/ZnSe core/shell quantum dots. The Journal of Chemical Physics 2021, 155, 084701.
  • Ubbink et al. 2022 Ubbink, R. F.; Almeida, G.; Iziyi, H.; du Fossé, I.; Verkleij, R.; Ganapathy, S.; van Eck, E. R. H.; Houtepen, A. J. A Water-Free In Situ HF Treatment for Ultrabright InP Quantum Dots. Chemistry of Materials 2022, 34, 10093–10103.
  • Hassan et al. 2018 Hassan, A.; Zhang, X.; Liu, C.; Snee, P. T. Electronic Structure and Dynamics of Copper-Doped Indium Phosphide Nanocrystals Studied with Time-Resolved X-ray Absorption and Large-Scale DFT Calculations. The Journal of Physical Chemistry C 2018, 122, 11145–11151.
  • Zhu et al. 2023 Zhu, D.; Bahmani Jalali, H.; Saleh, G.; Di Stasio, F.; Prato, M.; Polykarpou, N.; Othonos, A.; Christodoulou, S.; Ivanov, Y. P.; Divitini, G.; Infante, I.; De Trizio, L.; Manna, L. Boosting the Photoluminescence Efficiency of InAs Nanocrystals Synthesized with Aminoarsine via a ZnSe Thick-Shell Overgrowth. Advanced Materials 2023, 35, 2303621.
  • Kim et al. 2014 Kim, S.; Lee, K.; Kim, S.; Kwon, O.-P.; Heo, J. H.; Im, S. H.; Jeong, S.; Lee, D. C.; Kim, S.-W. Origin of photoluminescence from colloidal gallium phosphide nanocrystals synthesized via a hot-injection method. RSC Advances 2014, 5, 2466–2469.
  • Choi et al. 2023 Choi, Y.; Choi, C.; Bae, J.; Park, J.; Shin, K. Synthesis of gallium phosphide quantum dots with high photoluminescence quantum yield and their application as color converters for LEDs. Journal of Industrial and Engineering Chemistry 2023, 123, 509–516.
  • Geva et al. 2018 Geva, N.; Shepherd, J. J.; Nienhaus, L.; Bawendi, M. G.; Van Voorhis, T. Morphology of Passivating Organic Ligands around a Nanocrystal. The Journal of Physical Chemistry C 2018, 122, 26267–26274.
  • Kim et al. 2016 Kim, K.; Yoo, D.; Choi, H.; Tamang, S.; Ko, J.-H.; Kim, S.; Kim, Y.-H.; Jeong, S. Halide–Amine Co-Passivated Indium Phosphide Colloidal Quantum Dots in Tetrahedral Shape. Angewandte Chemie International Edition 2016, 55, 3714–3718.
  • Stein et al. 2018 Stein, J. L.; Holden, W. M.; Venkatesh, A.; Mundy, M. E.; Rossini, A. J.; Seidler, G. T.; Cossairt, B. M. Probing Surface Defects of InP Quantum Dots Using Phosphorus Kα𝛼\alphaitalic_α and Kβ𝛽\betaitalic_β X-ray Emission Spectroscopy. Chemistry of Materials 2018, 30, 6377–6388.
  • Kim et al. 2021 Kim, Y.; Choi, H.; Lee, Y.; Koh, W.-k.; Cho, E.; Kim, T.; Kim, H.; Kim, Y.-H.; Jeong, H. Y.; Jeong, S. Tailored growth of single-crystalline InP tetrapods. Nature Communications 2021, 12, 4454.
  • Zhao et al. 2022 Zhao, T.; Zhao, Q.; Lee, J.; Yang, S.; Wang, H.; Chuang, M.-Y.; He, Y.; Thompson, S. M.; Liu, G.; Oh, N.; Murray, C. B.; Kagan, C. R. Engineering the Surface Chemistry of Colloidal InP Quantum Dots for Charge Transport. Chemistry of Materials 2022, 34, 8306–8315.
  • Kim et al. 2022 Kim, T.; Kim, Y.; Park, S.; Park, K.; Wang, Z.; Oh, S. H.; Jeong, S.; Kim, D. Shape-Tuned Multiphoton-Emitting InP Nanotetrapods. Advanced Materials 2022, 34, 2110665.
  • Owen 2015 Owen, J. The coordination chemistry of nanocrystal surfaces. Science 2015, 347, 615–616.
  • Voznyy et al. 2012 Voznyy, O.; Zhitomirsky, D.; Stadler, P.; Ning, Z.; Hoogland, S.; Sargent, E. H. A Charge-Orbital Balance Picture of Do** in Colloidal Quantum Dot Solids. ACS Nano 2012, 6, 8448–8455.
  • Fritzinger et al. 2010 Fritzinger, B.; Capek, R. K.; Lambert, K.; Martins, J. C.; Hens, Z. Utilizing Self-Exchange To Address the Binding of Carboxylic Acid Ligands to CdSe Quantum Dots. Journal of the American Chemical Society 2010, 132, 10195–10201.
  • Moreels et al. 2011 Moreels, I.; Justo, Y.; De Geyter, B.; Haustraete, K.; Martins, J. C.; Hens, Z. Size-Tunable, Bright, and Stable PbS Quantum Dots: A Surface Chemistry Study. ACS Nano 2011, 5, 2004–2012.
  • Zherebetskyy et al. 2015 Zherebetskyy, D.; Zhang, Y.; Salmeron, M.; Wang, L.-W. Tolerance of Intrinsic Defects in PbS Quantum Dots. The Journal of Physical Chemistry Letters 2015, 6, 4711–4716.
  • Vydrov et al. 2007 Vydrov, O. A.; Scuseria, G. E.; Perdew, J. P. Tests of functionals for systems with fractional electron number. The Journal of Chemical Physics 2007, 126, 154109.
  • Kurth et al. 1999 Kurth, S.; Perdew, J. P.; Blaha, P. Molecular and solid-state tests of density functional approximations: LSD, GGAs, and meta-GGAs. International Journal of Quantum Chemistry 1999, 75, 889–909.
  • Azpiroz et al. 2014 Azpiroz, J. M.; Ugalde, J. M.; Infante, I. Benchmark Assessment of Density Functional Methods on Group II–VI MX (M = Zn, Cd; X = S, Se, Te) Quantum Dots. Journal of Chemical Theory and Computation 2014, 10, 76–89.
  • Boys 1960 Boys, S. F. Construction of Some Molecular Orbitals to Be Approximately Invariant for Changes from One Molecule to Another. Reviews of Modern Physics 1960, 32, 296–299, Publisher: American Physical Society.
  • Lehtola and Jónsson 2013 Lehtola, S.; Jónsson, H. Unitary Optimization of Localized Molecular Orbitals. Journal of Chemical Theory and Computation 2013, 9, 5365–5372.
  • Truhlar 2012 Truhlar, D. G. Are Molecular Orbitals Delocalized? Journal of Chemical Education 2012, 89, 573–574.
  • Foster and Boys 1960 Foster, J. M.; Boys, S. F. Canonical Configurational Interaction Procedure. Reviews of Modern Physics 1960, 32, 300–302, Publisher: American Physical Society.
  • Pipek and Mezey 1989 Pipek, J.; Mezey, P. G. A fast intrinsic localization procedure applicable for ab initio and semiempirical linear combination of atomic orbital wave functions. The Journal of Chemical Physics 1989, 90, 4916–4926.
  • Norris and Bawendi 1996 Norris, D. J.; Bawendi, M. G. Measurement and assignment of the size-dependent optical spectrum in CdSe quantum dots. Physical Review B 1996, 53, 16338–16346.
  • Ellingson et al. 2003 Ellingson, R. J.; Blackburn, J. L.; Nedeljkovic, J.; Rumbles, G.; Jones, M.; Fu, H.; Nozik, A. J. Theoretical and experimental investigation of electronic structure and relaxation of colloidal nanocrystalline indium phosphide quantum dots. Physical Review B 2003, 67, 075308.
  • Shulenberger et al. 2021 Shulenberger, K. E.; Coppieters ‘t Wallant, S. C.; Klein, M. D.; McIsaac, A. R.; Goldzak, T.; Berkinsky, D. B.; Utzat, H.; Barotov, U.; Van Voorhis, T.; Bawendi, M. G. Resolving the Triexciton Recombination Pathway in CdSe/CdS Nanocrystals through State-Specific Correlation Measurements. Nano Letters 2021, 21, 7457–7464.
  • Snee 2021 Snee, P. T. DFT Calculations of InP Quantum Dots: Model Chemistries, Surface Passivation, and Open-Shell Singlet Ground States. The Journal of Physical Chemistry C 2021, 125, 11765–11772.
  • Lebedev et al. 2016 Lebedev, M. V.; Kalyuzhnyy, N. A.; Mintairov, S. A.; Calvet, W.; Kaiser, B.; Jaegermann, W. Comparison of wet chemical treatment and Ar-ion sputtering for GaInP2(100) surface preparation. Materials Science in Semiconductor Processing 2016, 51, 81–88.
  • Hudson et al. 2022 Hudson, M. H.; Gupta, A.; Srivastava, V.; Janke, E. M.; Talapin, D. V. Synthesis of In1–xGaxP Quantum Dots in Lewis Basic Molten Salts: The Effects of Surface Chemistry, Reaction Conditions, and Molten Salt Composition. The Journal of Physical Chemistry C 2022, 126, 1564–1580.
  • Zhang et al. 2017 Zhang, W.; Zeng, X.; Su, X.; Zou, X.; Mante, P.-A.; Borgström, M. T.; Yartsev, A. Carrier Recombination Processes in Gallium Indium Phosphide Nanowires. Nano Letters 2017, 17, 4248–4254.
  • Breneman and Wiberg 1990 Breneman, C. M.; Wiberg, K. B. Determining atom-centered monopoles from molecular electrostatic potentials. The need for high sampling density in formamide conformational analysis. Journal of Computational Chemistry 1990, 11, 361–373, _eprint: https://onlinelibrary.wiley.com/doi/pdf/10.1002/jcc.540110311.