Ferdinand
Lédée
ab,
Pierre
Audebert
*b,
Gaëlle
Trippé-Allard
a,
Laurent
Galmiche
b,
Damien
Garrot
c,
Jérôme
Marrot
d,
Jean-Sébastien
Lauret
a,
Emmanuelle
Deleporte
*a,
Claudine
Katan
e,
Jacky
Even
*f and
Claudio
Quarti
*eg
aUniversité Paris-Saclay, ENS Paris-Saclay, CNRS, CentraleSupelec, LuMIn (Laboratoire Lumière, Matière et Interfaces), 91190 Gif-sur-Yvette, France. E-mail: emmanuelle.deleporte@ens-cachan.fr
bUniversité Paris-Saclay, ENS Paris-Saclay, CNRS, PPSM (Laboratoire de Photophysique et Photochimie Supramoléculaires et Macromoléculaires) – UMR 8531, 91190 Gif-sur-Yvette, France. E-mail: pierre.audebert@ppsm.ens-cachan.fr
cUniversité Paris-Saclay, UVSQ, CNRS, GEMaC (Groupe d’étude de la Matière Condensée) – UMR 8635, 78000 Versailles, France
dUniversité Paris-Saclay, UVSQ, CNRS, ILV (Institut Lavoisier de Versailles) – UMR 8180, 78000 Versailles, France
eUniv Rennes, ENSCR, INSA Rennes, CNRS, ISCR (Institut des Sciences Chimiques de Rennes) – UMR 6226, F 35000 Rennes, France. E-mail: claudio.quarti@umons.ac.be
fUniv Rennes, INSA Rennes, CNRS, Institut FOTON (Fonctions Optiques pour les Technologies de l’Information) – UMR 6082, 35000 Rennes, France. E-mail: jacky.even@insa-rennes.fr
gUniversity of Mons, Laboratory for Chemistry of Novel Materials, B-7000 Mons, Belgium
First published on 9th March 2021
Taking advantage of an innovative design concept for layered halide perovskites with active chromophores acting as organic spacers, we present here the synthesis of two novel two-dimensional (2D) hybrid organic–inorganic halide perovskites incorporating for the first time 100% of a photoactive tetrazine derivative as the organic component. Namely, the use of a heterocyclic ring containing a nitrogen proportion imparts a unique electronic structure to the organic component, with the lowest energy optical absorption in the blue region. The present compound, a tetrazine, presents several resonances between the organic and inorganic components, both in terms of single particle electronic levels and exciton states, providing the ideal playground to discuss charge and energy transfer mechanisms at the organic/inorganic interface. Photophysical studies along with hybrid time-dependent DFT simulations demonstrate partial energy transfer and rationalise the suppressed emission from the perovskite frame in terms of different energy-transfer diversion channels, potentially involving both singlet and triplet states of the organic spacer. Periodic DFT simulations also support the feasibility of electron transfer from the conduction band of the inorganic component to the LUMO of the spacer as a potential quenching mechanism, suggesting the coexistence and competition of charge and energy transfer mechanisms in these heterostructures. Our work proves the feasibility of inserting photoactive small rings in a 2D perovskite structure, meanwhile providing a robust frame to rationalize the electronic interactions between the semiconducting inorganic layer and organic chromophores, with the prospects of optimizing the organic moiety according to the envisaged application.
New conceptsLayered metal halide perovskites are natural quantum-well semiconductors, where frontier electrons and holes are often confined on atomically thin inorganic sheets, spaced by insulating organic moieties. At the same time, the inclusion of optically active chromophores as an organic spacer represents an effective strategy to further tune the optoelectronic properties of these systems, opening the way for several charge and energy transfer phenomena at the inorganic/organic interface, like exciton separation or organic-mediated charge transport, potentially useful for optoelectronic applications. Still, thorough understanding of these processes has to be achieved, in order to fully exploit them in real devices. Here, we provide a detailed picture of the charge and energy transfer processes taking place in newly synthesized layered chloride and mixed chloride/bromide lead perovskites, incorporating a tetrazine-based chromophore as an organic spacer with a strong oxidizing character. Various resonances highlighted by optical characterization and hybrid DFT simulations at the level of monoelectronic and exciton states allowed us to discuss several decay channels for the photogenerated carriers, responsible for the observed perovskite emission quenching. This study provides then a robust foundation to discuss charge and energy transfer processes in layered halide perovskites, in the presence of active organic chromophores. |
2D organic–inorganic perovskites are characterized by a usual chemical formula (R-NH3)2MX4, where (R-NH3+) is a bulky organic cation, M is a divalent metal (such as Sn2+ or more commonly Pb2+) and X is a halide anion (Cl−, Br− or I−). To date, most of the studies on 2D perovskites have utilized aliphatic or mono aromatic compounds such as butylammonium or phenylethylammonium which are characterized by large optical band gaps (5–6 eV). It is now known that many of these 2D perovskites form type I multiple quantum well (QW)-like structures17 in which the organic spacers act as an insulating barrier while the inorganic haloplumbate layers act as QWs for both electrons and holes. The classical concept of type I semiconductor QW (Fig. 1a) must be however handled with great care, because several fundamental differences exist between 2D perovskites and classical QW. These include the lack of suitable definition of a transverse effective mass for the barrier, the strong influence of carrier dispersion non-parabolicity in the well and the absence of common Bloch states for the well and the barrier.18 In classical semiconductor heterostructures, this schematic representation is usually associated with a free carrier picture, where a pair of electrons and holes can flow from one region to the other almost independently. In such a case the overlap of wavefunctions between the two semiconductors plays a major role in the transport of particles. The weakness of the excitonic interaction modifies only slightly the free carrier picture, and an indirect exciton may for example appear at the interface between two materials for type II QW (Fig. 1b). On the other hand, the free carrier picture breaks down in 0D semiconductor quantum dots owing to charge localization effects and enhanced quantum confinement. In that case, multiexcitons and charged excitonic states provide a more suitable representation of the crystal optoelectronic properties. For 2D perovskites, additionally to the quantum confinement in the quantum wells, the large difference between the dielectric constants of the inorganic haloplumbate layers and the organic layers leads to a dielectric confinement,19,20 which sizeably enhances the electron–hole pair interaction and results in exciton binding energies in the typical alkylammonium-based 2D organic–inorganic perovskites up to 500 meV.20 As a consequence, these self-assembled films exhibit bright and narrow photoluminescence at room temperature, which makes them ideal candidates for light-emitting applications.21,22 Conceptually although highly localized within the stacking axis, these 2D excitonic states can be considered as Wannier-like delocalized excitations developed on the basis of monoelectronic states20 and free to travel along the inorganic layer.23 They are therefore prone to carrier deconfinement (Fig. 1b) and vertical localization.
On the opposite side of the barricade, donor–acceptor molecular pairs are interacting systems dominated by localized Frenkel pair states. The Förster resonance energy transfer (FRET) from a donor to an acceptor (Fig. 1c) is a perfect example of a non-radiative excitation transfer, not relying on a monoelectronic state picture (Fig. 1a and b). The transfer or optical matching between two optically allowed transitions depends on the overlap integral of the acceptor absorption spectrum with the donor emission spectrum and the mutual transition dipole orientation, and classically exhibits a long-range ∼1/d6 dependence on the distance d between the donor and the acceptor. Let's also mention that the energy matching between the two optically allowed transitions, and the resulting resonance effect, are usually attributed to two singlet excited spin states (S1 on Fig. 1c). The Dexter non-radiative mechanism allows transferring excitations between two optically forbidden states, carrying null transition dipoles, as in the case of spin-forbidden triplet states T1 on Fig. 1c or dipole-forbidden singlet S1 states (Fig. 1c). This short-range interaction requires overlap between the wavefunctions, which in that sense makes this mechanism closer than FRET to the classical semiconductor processes depicted in Fig. 1a and b. Era and coworkers deduced that the observed exponential variation of the energy transfer from the perovskite lattice to naphthalene was more compatible with a Dexter mechanism than FRET.24 Still, it is worth mentioning that the classification of energy transfer processes in lead halide perovskites as singlet–singlet or triplet–triplet is conceptually challenging, as exciton states in these systems are not pure spin-states (either singlet or triplet) but have mixed character, resulting from sizable spin–orbit coupling interaction.25,26 Apart from the optical resonance effect, the overlap between wavefunctions at the heart of the Dexter process led several authors to the empirical conclusion that beside energy matching, band alignments between the well and the barrier in a hybrid perovskite structure will also play an important role.27
From that perspective, 2D hybrid halide perovskites featuring a type II organic/inorganic interface, i.e. one frontier orbital HOMO/LUMO of the organic spacer lies in the energy gap between the CBM and VBM of the perovskite frame, have been seldom reported.28–33 In that case, the band alignment is expected to enhance charge separation at the interface, and subsequently improve out-of-plane conductivity, as well as the photovoltaic performances.29,33 Suitable band alignment may be crucial for optimizing the perovskite material for a given application. Energy matching and band alignment engineering between the organic/inorganic parts can either be carried out by tuning the halide composition,34,35 by modifying the number of haloplumbate layers,36 or by adjusting the HOMO/LUMO gap and/or levels of the organic spacer, or in other words, incorporating optically active molecules in the perovskite structure. Indeed, instead of playing a simple passive barrier role, the organic part is flexible enough in the interstitial dimension (z-axis perpendicular to the perovskite planes), and therefore can be used to engineer the optoelectronic properties to these materials. Depending on the optical matching, defined for hybrid perovskites as the spectral overlap between the perovskite and organic absorber optical transitions, along with the electronic level alignment (type I/II, see Fig. 1a and b) at the interface, charge and/or energy transfers can be induced between the inorganic and organic moieties.
From the chemistry viewpoint, previous research focused on polyaromatic organic cations with reduced band gap such as polythiophene,30,33,37,38 naphthalene derivatives,39,40 anthryl,41 pyrene29,35,42,43 and perylene29 derivatives. One of the first trials was performed by Era et al. who prepared bromide-based perovskites containing a naphthalene luminophore.44 It was later on shown, that these compounds feature extremely efficient (>99%) perovskite-to-luminophore energy transfers.24 Another pioneering attempt to introduce a potent organic absorber was performed by Mitzi et al. who succeeded in preparing iodide-based perovskites featuring quaterthiophenes as the organic part.37 They observed in that case mutual quenching between the optical properties of both partners,45 which was recently associated with a type II interface.31 Recent studies also showed that molecular engineering may quench the perovskite exciton luminescence and lead to efficient formation of mobile free charge carriers allowing a better extraction of charge carriers from the perovskite material, as demonstrated in quasi-2D cesium lead bromide colloidal nanoplatelets, consisting of four layers of lead bromide octahedra (CsPbBr3 NPLs, n = 4) with a reduced (∼80 meV) exciton binding energy.46 A recent attempt by some of us was made to introduce a linear naphthalimide inside the organic part of a monolayered perovskite, in order to induce luminophore-to-perovskite energy transfer.47 A perfect optical matching in the case of a bromide perovskite was obtained, leading to about a four-fold enhancement of the perovskite emission. However, only a limited amount of luminophore (10%) could be introduced while retaining the perovskite structure, probably because of the lattice mismatch between the lead octahedra and the ammonium salt containing the organic luminophore.
On the other hand, luminophores belonging to the tetrazine (Tz) family have been developed over the last 15 years by some of us, and demonstrated their efficiency as good organic emitters, associated with a very small size and molecular weight, allowing for example unexpected applications like fingerprint revelation.48 Tz derivatives possess quite unique optical and electronic features as regards to others luminophores used as active barriers in 2D perovskites. Most notably, these systems exhibit low energy optical transition in the blue region, which in acenes- and thiophenes-derivatives is usually obtained via the fusion of 3–5 rings. Such low energy transition for a small-size organic molecule is due to its peculiar electronic structure, as the partial substitution of carbon with nitrogen atoms turns the character of the frontier orbitals from π–π*, common to benzene-derivatives, to n–π*, with corresponding narrowing of the HOMO–LUMO band gap.49 A higher energy π–π* transition is also found in these compounds, in the near-UV range. Tetrazine derivatives also show significant fluorescence, with record quantum emission efficiency (ca. 40%) and lifetime (160 ns),50 along with a strong oxidizing character both in the ground and excited states. In particular, Tz is known to undergo oxidative electron transfer in the presence of an electron donor, a process that consists in the oxidation of a nearby molecular species by the photoexcited tetrazine moiety, see Fig. 1d. In light of its intriguing light-emitting properties along with its reduced molecular size, Tz represents the ideal organic dye for incorporation in the lead halide frame, to probe charge and energy transfer processes in hybrid layered perovskite materials. Altogether, the use of nitrogen containing heterocycles as spacers in layered perovskite structures represents a novel strategy for the incorporation of optically active organic components in these systems, at least, compared to the dominant paradigm of extending the π-conjugation of the molecular spacer.30 In this work, we synthesise two novel layered perovskites containing tetrazines as the sole organic spacer and report thorough optical characterization via optical spectroscopy, with theoretical support from Density Functional Theory (DFT). We focused more specifically on the competition between several deactivation channels possibly leading to luminescence quenching of the perovskite frame, with the aim to better rationalize the possible charge and/or energy transfer mechanisms at the interface between the two components.
Unfortunately, only the chloride salt of the tetrazinammonium could be produced, because of the impossibility to obtain hydrobromic acid dry enough, resulting in salt degradation during the formation reaction. This chloride salt, 2-(6-ethoxy-1,2,4,5-tetrazin-3-yl)oxyethylammonium chloride, will be designed by TzCl in the following. This situation limited the relative incorporation of any other ion than chloride into the final perovskite to 50%, as this must be introduced via the lead halide precursor. Since iodide is easily photo-oxidized by the tetrazine ring, the choice was limited to chloride and bromide. Two perovskites were finally investigated, the pure chloride based one Tz2PbCl4 and the mixed bromide–chloride one Tz2PbCl2Br2. They were prepared from the precursors according to the following chemical reactions:
2TzCl + PbCl2 → Tz2PbCl4 |
2TzCl + PbBr2 → Tz2PbCl2Br2 |
Additional XRD measurements performed on single crystals of Tz2PbCl4 definitely demonstrate that this compound crystallizes in the form of a monolayered lead halide perovskite (R-NH3)2MX4, with an inorganic layer thickness consisting of just one PbCl4 octahedron (see the ESI† for details). The resolved structure is monoclinic and belongs to the P21/c space group, with lattice parameters a, b and c, equal to 19.05 Å, 7.75 Å and 8.29 Å, respectively and a β angle of 93.02° (Fig. 2e and Fig. SX1, ESI†). The atomic positions associated with the ammonium bridge are affected by conformational disorder, with two accessible molecular conformations, as shown in Fig. SX1 (ESI†), but with the spacer clearly exhibiting a well-oriented molecular stacking. The newly introduced Tz organic spacer features similar herringbone packing as found in phenylethylammonium lead chloride.52 The long molecular axis does not lie orthogonal to the inorganic plane, as often occurs for short spacers, but is tilted with respect to it by 46 degrees, comparable to the 47 degrees found for the phenylethylammonium spacer (see Fig. SX2, ESI†). The centre of mass distances between nearest neighbour π-conjugate rings correspond to 5.0 Å and 6.4 Å but with the molecules shifted with respect to each other, both along their molecular axis and along the b-axis of the lattice, in a non-optimal configuration for π–π interactions. A very short distance of 3.5 Å is found instead between the centre of the π-conjugated ring of the spacer and the closest apical chlorine (see Fig. SX3, ESI†), which is compatible with energy transfer processes, or with a modification of the electronic structure allowing the separation and transfer of charge carriers. For the sake of comparison, a distance of 4.05 Å was found by Dou and co-workers for a lead-iodide lattice incorporating thiophene derivatives, for which charge transfer from the inorganic lattice to the spacer has been reported (see Fig. SX3, ESI†). Thanks to the single crystal data, Tz2PbCl4 thin film diffraction peaks shown in Fig. 2b (red curve) can be indexed, thus revealing the (h 0 0) family of planes that corresponds to the inorganic haloplumbate layers (see Fig. SX4, ESI†). This shows that Tz2PbCl4 thin-films grow with a preferential in-plane direction with respect to the inorganic layers, which is a typical behaviour observed in 2D hybrid perovskites.
Unfortunately, single crystals of Tz2PbCl2Br2 could not be obtained, and the 2D hybrid perovskite-like diffraction pattern shown in Fig. 2c was not indexed. However, the similarities between both perovskites’ diffraction diagrams and further optical absorption studies allow us to reasonably infer that this compound also presents a (R-NH3)2MX4 layered perovskite structure.
Interestingly, our newly synthesised Tz-based hybrid perovskites show absorption signals resulting from the superimposition of both Tz salt and inorganic PbX4 lattice optical responses. Indeed, both Tz-based perovskites show an intense excitonic peak at ∼335 nm and ∼375 nm for the chloride and mixed halide compound, respectively. The existence of this sharp excitonic peak is additional proof that a perovskite structure has been synthesised. Additionally, the Tz S1 and S2 excitonic resonances, at ∼540 nm and ∼365 nm respectively, are superimposed on the lead-halide layer absorption. These results demonstrate that the Tz moiety has been successfully included in the lattice as an organic spacer. Contrary to the case of PEA-based perovskites, where the PEA plays little role close to the band gap, Tz is actively involved in the optical response of the hybrid perovskite materials. The optical properties of our newly reported compounds are summarized in Table SX1 (ESI†). Notably, both in the case of Tz2PbCl4 and Tz2PbCl2Br2, there is sizable spectral overlap between the S2 state of the organic spacer and the E1s state of the inorganic frame, which is a crucial requirement for energy transfer to take place. Halide composition clearly influences the detailed S2/Es1 alignment, with S2 being more stable by ca. 0.3 eV than the E1s interaction, for the pure chlorine compound, while the reverse order is found for the mixed one. As this order determines the energetic gain/barrier associated with the energy transfer process, halide substitution may be sought as a strategy to tune/detune this mechanism, although partial substitution of chlorine with bromine, as pursued here, was not sufficient to completely switch off this non-radiative mechanism.46
It is important to stress here that the concentration of the precursors was chosen such that the Tz-based perovskites and TzCl thin films present the same optical density at the wavelength of S1, that is to say that the quantity of Tz molecules are the same in the TzCl and Tz-based perovskite layers. This allows direct comparison of the PL intensities between Tz-based perovskites and the TzCl salt. The S1 emission from the Tz molecule in the Tz-based perovskites appears one order of magnitude smaller than in TzCl films (Fig. 3a and b). This observation could tentatively be attributed to the different packing of the Tz moieties in the perovskite in comparison with the TzCl salt. In fact, a decrease of the luminescence efficiency has been already observed in ref. 50 from TzCl in solution and TzCl in the solid phase, and attributed to the increase of non-radiative losses associated with lattice vibrations in the condensed phase. The quantitative decrease of the S1 emission from TzCl to Tz2PbCl4 thin films could then be attributed to the different organizations of the Tz molecules in these two compounds as shown by the XRD experiments. Besides, we note that considering the standard redox potential of the different species (see discussion in the ESI†), it is very unlikely that the partial quenching of the luminescence of the tetrazine moiety is due to oxidative electron transfer (Fig. 1d). In addition, if such a process were to occur, we would observe fast degradation, not only of the perovskite, but also already of the generic tetrazine-ammonium TzCl salt, accompanied by halogen formation, under UV light (the intermediately produced tetrazine anion-radical being efficiently restored by oxygen, a catalytic oxidation would take place).56 This is not the case, with the tetrazine contained in the perovskite matrix appearing even a little more stable than the generic TzCl salt (see the ESI,† Fig. SX6).
Fig. 4 Singlet and triplet exciton states in tetrazine and related exciton diversion channels for the hybrid halide perovskite. (a) Singlet and triplet excited state energies computed with TD-DFT for the Tz molecule, compared with previous estimates for non-functionalised tetrazine, from wavefunction-based approaches (Equation Of Motion Coupled Cluster EOM-CC,58 and Complete-Active-Space 2nd order Perturbation Theory, CAS-PT2);59 (b) diversion channels for the radiative emission from the exciton E1s of the inorganic PbX4 frame, involving exciton transfer (via FRET and/or Dexter mechanisms) either to the S2 singlet or the T3 triplet excited state of Tz. Continuous and dashed lines represent light absorption and radiative recombination, respectively, while curved lines represent excitation transfer (FRET/Dexter) and non-radiative recombination mechanisms. |
Moreover, we have seen in Fig. 3a and b that the emission of the Tz molecules is one order of magnitude lower in the perovskite, as compared to the TzCl salt. As mentioned earlier this observation could be related to the enhancement of the non-radiative exciton relaxation, as a result of the different packing of the Tz molecules in the salt and in the perovskite frame. However, another possibility is related to an efficient exciton transfer from the inorganic E1s exciton to a triplet excited state of Tz, via a Dexter transfer mechanism.24,60 The latter is traditionally associated with the quenching of the emission due to the spin–forbidden radiative relaxation from the molecular excited triplet to the singlet ground state. This is justified by the non-pure spin character of the excited state E1s of the perovskite (as induced by large spin–orbit coupling25) which makes possible both singlet–singlet and triplet–triplet transfer mechanisms. Most notably, this exciton transfer requires the existence of a triplet excited state in Tz with energetics close to E1s. Here, our calculations show that a triplet state T3 exists, with energy close to the E1s state of the perovskite PbX4 frame, hence potentially supporting the transfer from E1s to the triplet excited state manifold of Tz as a potential mechanism of the emission decrease of the organic molecules in the perovskite structure. The diversion channel via transfer from E1s to the triplet T3 state of Tz is also schematized in Fig. 4b.
Fig. 5 DFT simulations of the electron energy level alignment at the organic/inorganic interface. (a) Derivation of the models for mixed chloride/bromide perovskites (bromine in the apical position, Br-ap, in the equatorial position, Br-eq, or randomly distributed, Br-rand) from the pure chloride model in Fig. 2e. The color-scale is: black = lead, orange = chlorine, red = bromine; (b) DFT band gap variation for the different models and composition of Tz-based halide perovskites; (c) projected Density of State computed for the pure chloride Tz2PbCl4 compound. The positions of the frontier orbitals of the organic and inorganic components are indicated, for clarity; (d) density isosurfaces corresponding to the valence band maximum and conduction band minimum/LUMO level (VBM, CBM/LUMO); (e) projected Density of State of the various mixed halides Tz2PbCl2Br2 and pure halide Tz2PbBr4. |
We then simulated the single particle electronic structure of these crystalline models, with specific interest in the alignment between VBM/HOMO and CBM/LUMO levels. It is worth mentioning in this respect that the quantitative estimates of single particle energetics for lead-based halide perovskites is a challenging task, as it requires the inclusion of both relativistic, spin–orbit coupling effects25 and the accurate treatment of exchange correlation interaction. The latter is accounted for in the literature either via many-body approaches based on the GW approximation62 or via the use of hybrid DFT functionals.26,63 We followed here this second approach, by adopting the PBE0 hybrid exchange–correlation functional64 with the exact exchange contribution increased to 30%, as this was shown to provide accurate estimates of the electronic band gap of layered lead–iodide perovskites.26 The band gaps computed for pure and mixed chloride–bromide systems are reported in Fig. 5b.
For the pure chloride compound, our approach predicts a band gap of ∼4.0 eV, which nicely parallels the exciton resonance from UV-vis measurements in Fig. 2f at 3.68 V, assuming additional exciton binding energy of ∼300 meV.20 For mixed halide perovskites, the resulting band gap depends on the distribution of chlorine/bromine atoms in apical/equatorial distribution, as expected both from physical intuition and symmetry arguments.23 Namely, with bromine in an equatorial position (Br-eq), the band gap is very close to that of the pure bromine case, consistent with the fact that the electronic dispersion in the inorganic plane is associated with the electronic interaction between lead and the atoms in an equatorial position. Actually, in-deep symmetry analyses of the layered perovskite structure reveal that apical halides still take part to the valence band edge and can therefore affect the band gap of the system.65 This fact can explain the unexpected result for the band gap computed for the Br-ap model, in Fig. 5d. In spite of having all equatorial atoms corresponding to chlorides, this system presents in fact an intermediate band gap between pure bromide and chloride phases. In this sense, symmetry analysis justifies the trend in the band gaps depicted in Fig. 5b either in terms of indirect perturbation, due to not only different dielectric environments or structural distortions, but also as direct perturbation on the atomic hybridization pattern, as induced to the apical bromides.65 Finally, random distribution of the halides in apical/equatorial positions (Br-rand) results in an intermediate band gap, compared with the pure case, consistent with the monotonic dependence of the band gap on the bromide/iodide ratio reported in the literature.23 Most notably, the 250 meV band gap closing compared to pure chloride nicely parallels the 300 meV reduction of the exciton E1s resonance from Fig. 2f and g, suggesting that the halides are randomly distributed in the apical/equatorial position, in real samples. This conclusion is further supported by our DFT total energy estimates, which suggest that Br-rand and Br-eq are almost iso-energetic at zero Kelvin, while Br-ap is less stable by ∼50 meV per chemical unit. In other words, there is no energy gain going from bromines randomly distributed in apical/equatorial positions to all bromines located in the equatorial position, with hence the former configuration favoured for entropic reasons.
VBM/HOMO and CBM/LUMO alignment is first evaluated for the pure chloride model, considering the atomic projected Density of States (pDOS) in Fig. 5c. The VBM and CBM of the PbCl4 frame nicely correspond to 3pCl–6sPb and 3sCl–6pPb orbitals, respectively, as widely reported in the literature.26,28,66 The HOMO level of Tz lies 0.4 eV below the VBM of the inorganic lattice, making the hole injection process from the PbCl4 frame to the Tz molecule thermodynamically non-spontaneous. In the energy region of the CBM/LUMO orbitals, instead, both the organic and inorganic components similarly contribute to the electronic states, hence complicating the rigorous assignment of the CBM and LUMO levels (magnification of the pDOS in the CBM/LUMO region is reported in the ESI,† Fig. SX8). To clarify such an assignment, we therefore analysed the surface isodensities associated with the wavefunction of the last occupied and first unoccupied state, shown in Fig. 5d. These results clearly indicate that the highest energy occupied electronic level is a pure state of the inorganic PbCl4 lattice, consistent with the pDOS in Fig. 5c. In contrast, the first unoccupied state cannot be assigned solely to the inorganic/organic component, as it delocalises over both the Pb-atoms and on the π-conjugated system of Tz. This result is likely strongly influenced by the close energy resonance of the inorganic CBM and organic LUMO levels and should therefore be considered with care, in light of the possible perturbations to the organic/inorganic level alignment. In particular, thermally induced atomic vibrations (phonons) are expected to alter the energetics of these states. On the other hand, the electronic configuration reported in Fig. 5d opens the prospect for potential ionization of the exciton E1s photogenerated on the PbCl4 site and subsequent charge transfer of the electron into the organic Tz spacer.
Similar analysis of the VBM/HOMO–CBM/LUMO alignment is performed in Fig. 5e for all the mixed halide models. Notice that the pDOS of all the mixed chlorides/bromides and of the pure bromide compound have been aligned with respect to the 5d3/2 Pb states of the chloride compound (see the ESI,† Fig. SX9), so as to provide a common energy reference for all the models. The band gap closing from Tz2PbCl4 to Tz2PbCl2Br2, and finally to pure Tz2PbBr4 is mainly associated with a destabilization of the valence band, as shown in Fig. 5e, consistent with the fact that the valence is mainly composed by the outer p-shell of the halide. As a result, the VBM/HOMO energy mismatch increases from Tz2PbCl4 to Tz2PbCl2Br2, therefore enhancing the energy barrier for the hole injection process from the inorganic to the organic Tz molecule. In contrast, the CBM/LUMO levels of the Br-rand and pure Br models remain very close in energy, hence resulting in null (or small) energy barrier for the electron transfer from the PbX4 frame to the Tz dye. For the Br-rand model, unoccupied electronic states lie very close in energy (within 40 meV) and result in similar mixing of states from the organic/inorganic components as found for the pure chloride compound in Fig. 5d. In the pure Tz2PbBr4 model, the states are slightly more spaced (ca. 60 meV) hence reducing the mixing between inorganic and organic electronic states, which is however still visible (see the ESI,† Fig. SX12). For the Br-eq model, instead, the energy spacing among non-occupied states of ca. 165 meV results in clear separation of the organic and inorganic states, with a more stable LUMO, compared to the CBM, still allowing for spontaneous electron injection from the PbX4 frame into Tz. Only in the case of Br-ap, we found the LUMO to be less stable and well separated from the CBM (ca. 250 meV), in a final alignment that therefore does not allow for the spontaneous injection of the electron. This model however was found to be significantly less stable than the others, hence representing a less reliable model for the investigation of the electronic interface between the PbX4 lattice and the organic Tz spacer.
Finally, while the tetrazines were chosen for their perfect steric hindrance matching with the phenyl ring, they are not the only photoactive small rings to possibly adapt and insert into a 2D perovskite structure. Several polyaza aromatic heterocycles are eligible (and among them especially a-triazines) to this characteristic (see Fig. 6), the only mandatory condition being the possibility to bear an aminoethyl chain collinear with the aromatic ring, with in addition some flexibility linked to the possible introduction of small substituents (e.g. halogens) on the phenyl ring to tune the orbital levels. Suitable molecular engineering could lead to e.g. total quenching or enhancement of the luminescence of the two parts, depending on the target application. The recent work of H. Karunadasa et al. also showed that partial insertion of reduced pyrazinium heterocycles in expanded 3D perovskite analogues offers a new route for chemical doping of the semiconductor.68 Building a 2D perovskite with a full load of active organic moieties, tuned with the perovskite emission, is therefore an emerging field allowing hope in the building of new and better performing materials in this family.
Fig. 6 Molecular design. Possible aromatic heterocyclic luminophore candidates for functionalization of the organic moiety in 2D perovskite. |
Footnote |
† Electronic supplementary information (ESI) available: Details on the synthesis and XRD characterization (cif file deposited at the CCDC with reference 2038924); photoluminescence for pristine PEA2PbCl4 perovskite; detailed optical properties of the newly synthesized compounds; details on oxidative electron transfer; details on the computational parameters for periodic DFT and molecular TD-DFT calculations; detailed analysis of the atomic projected Density of States from DFT and related isosurfaces of frontier electronic levels. For ESI and crystallographic data in CIF or other electronic format see DOI: 10.1039/d0mh01904f |
This journal is © The Royal Society of Chemistry 2021 |