Startseite Synthesis, characterization and optoelectronic properties of 2D hybrid RPbX4 semiconductors based on an isomer mixture of hexanediamine-based dications
Artikel Öffentlich zugänglich

Synthesis, characterization and optoelectronic properties of 2D hybrid RPbX4 semiconductors based on an isomer mixture of hexanediamine-based dications

  • Anna Ioannou , Ioanna Vareli , Andreas Kaltzoglou und Ioannis Koutselas EMAIL logo
Veröffentlicht/Copyright: 27. September 2021
Veröffentlichen auch Sie bei De Gruyter Brill

Abstract

Three new hybrid two-dimensional (2D) organic–inorganic semiconductors are presented, which contain lead halides and a mixture of hexanediamine-based isomers in the stoichiometry [2,2,4(2,4,4)-trimethyl-1,6-hexanediamine]PbX4 (X = I, Br, Cl). These hexanediamine derivatives, with attached methyl groups at the carbon backbone of both isomers, determine the packing of the organic layers between the inorganic 2D sheets, while the optical absorption and photoluminescence spectra reveal excitonic peaks at T = 77 K and room temperature. The as-synthesized semiconductors were stored for three years in the dark and under low humidity and were examined again and the results were compared to those of the fresh materials. The chloride analogue, after the three year storage, displays white-like luminescence. The use of non-equivalent isomer and racemic mixtures in the organic component to form hybrid organic–inorganic semiconductors is an efficient method to alter the properties of 2D perovskites by tuning the isomers’ chemical functionalities. Finally, a comparison of the observed excitonic absorption and photoluminescence signals to that of analogous 2D compounds is discussed.

1 Introduction

Materials exhibiting the perovskite structure have been known since the 19th century [1]. However, the low-dimensional organic–inorganic semiconductors (LD HOIS) based on metal halide perovskites have redrawn the attention of the research community over the past 10 years, thanks to their tunable optoelectronic properties, their low cost, simple synthesis as well as the plethora of their new applications [2], [3], [4], [5], [6], [7], [8]. Since 2009, when the first perovskite photovoltaic device was reported [9], the photovoltaic power conversion efficiency has reached 25.2%, which is close to the silicon-based single-junction photovoltaics [10], while perovskite/silicon tandem solar cells have shown efficiencies up to 29.1% [1014]. HOIS research has focused on semiconducting applications, such as luminescence, photodetection, chemical sensors, and photocatalysis [15], [16], [17], [18], [19], [20], [21], [22], [23], [24].

The general chemical formula of halide perovskites is R y M x X z , where R is typically a large or small mono/di-valent cation, M is a divalent cation, usually Pb or Sn, while X is Cl, Br or I [25]. This leads to hundreds of chemical compounds, but the main trend [26], [27], [28], [29], [30], [31], [32], [33], [34], [35], [36], [37] is to incorporate either methylammonium or formamidinium ions as the cation, Pb as the metal and iodine or a mixture of halides as the anion, at least for solar cells. Their structure usually consists of face- or corner-sharing MX6 octahedra where the organic cations fill the voids among the octahedra in the 3D systems; in the 2D perovskites the organic molecules stabilize the 2D corner-sharing octahedra-based layers [38]. By adjusting the chemical composition of a variety of perovskites, their optical/electronic properties can be tuned [39, 40]. Moreover, their active inorganic structures can be tuned to behave as 3D, 2D, 1D, 0D or intermediate dimensionalities semiconductors [2, 6, 28, 41, 42].

A simple method to produce LD HOIS is to introduce large organic molecules as cations [43], [44], [45], [46], [47], [48], [49], [50], [51], [52]. In this case the contributing function of the R-site cation, besides leading the 2D (lamellar) formation, is providing anisotropy to a variety of properties as well as to carrier properties and diffusion lengths for which details have been reported numerous times [53], [54], [55], [56]. Also, a recent example is the use of [NH3(CH2)6NH3]PbI4 in perovskites in solar cells, since its 2D structure passivates the humidity sensitive 3D mixed-anion and mixed-cation perovskite, that acts as an efficient light absorber [57] yet allowing the electrical carriers’ drift.

Here, three new 2D HOIS based on the long-chain amine 2,2,4(2,4,4)-trimethyl-1,6-hexanediamine are presented. This amine is composed of a mixture of the 2,2,4- and 2,4,4-isomers, as acquired, both being soluble in water, thus, degrading H2O molecules attacking these compounds could be removed by drying without leading to a phase separation of the organics. It is interesting to study the formation of 2D HOIS where the organic spacer, such as this one, first alters its conformation slightly as a double-protonated salt compared to its free-standing form, and second its conformation with the methyl groups around its carbon backbone presents a packing problem in order to support the naturally forming 2D inorganic structures. In fact, both isomers unfold slightly when protonated, however, their length is slightly different between the two isomers according to the ab initio calculations presented herein, due to the distance of the two methyl groups from the nitrogen atom. It is expected that the final 2D hybrid material could contain a mixture of the isomers as organic spacers. It is also possible that these could aggregate in 2,4,4- and 2,2,4-rich crystal domains. This is particularly favorable for the iodine compound where the large I–Pb–I lengths allow for more packing possibilities of the organic spacer which needs to grasp the inorganic layer at every apical iodine via electrostatic bonds. The synthesis, crystal structure and some optoelectronic properties of [2,2,4(2,4,4)-trimethyl-1,6-hexanediamine]PbX4 (X = I, Br, Cl) are reported and discussed. It is important to note that both the 2,2,4 and 2,4,4 isomers most probably are composed of equal amounts of enantiomers, thus, it is expected that this underlying chirality also plays a crucial role for both structure and optical properties. To our knowledge, there is no report concerning the chirality of this particular amine. Also, related preliminary studies of their optical properties did not reveal strong chiral properties. Finally, the optical properties of the here reported materials display all the characteristic novel phenomena of the 2D hybrid lead halide semiconductors, however, the excitonic peaks observed are broadened and diminished in intensity.

2 Experimental section

2.1 Starting materials

The following materials were used as received without any further purification; lead(II) iodide 99% (Sigma-Aldrich 211168), lead(II) bromide 98%+ (Sigma-Aldrich 211141), lead(II) chloride (Alfa 21490), 2-2-4(2,4,4)-trimethyl-1,6-hexanediamine 99% (Sigma-Aldrich 722650), 57% aqueous hydroiodic acid (Sigma-Aldrich 210021), 48% aqueous hydrobromic acid (Alfa-Aesar 14036), 37% aqueous hydrochloric acid (Sigma-Aldrich 320331), anhydrous acetonitrile (Sigma-Aldrich 34851), dimethyl sulfoxide 99.9% (DMSO, Sigma Aldrich 276855) and N,N-dimethylformamide (DMF, Sigma-Aldrich 68-12-2).

2.2 Synthesis of 2,2,4(2,4,4)-trimethyl-1,6-hexanediamine lead halide materials

2.2.1 Synthesis of (2,2,4(2,4,4)-trimethyl-1,6-hexanediamine)PbI4 (m1)

240 μL of 2,2,4(2,4,4)-trimethyl-1,6-hexanediamine (1.3 mmol) was mixed with 300 μL of 57% aqueous hydroiodic acid (2.2 mmol). 0.461 g of PbI2 (1 mmol) was dissolved in a mixture of 1.2 mL CH3CN and 60 μL of HI (0.45 mmol). The two solutions were mixed under stirring at 40 °C until dark orange-reddish crystals were formed by slow solvent evaporation. Other approaches to synthesize variants of this compound (denoted as m1_ns/2/3/4) were investigated and described in detail in the Supplementary Material, which provide similar peaks in the low angle XRD patterns.

2.2.2 Synthesis of (2,2,4(2,4,4)-trimethyl-1,6-hexanediamine)PbBr4 (m2)

362 μL of 2,2,4(2,4,4)-trimethyl-1,6-hexanediamine (2 mmol) and 120 μL of 48% aqueous hydrobromic acid was mixed (1 mmol). 0.734 g of PbBr2 (2 mmol) was dissolved in a mixture of 800 μL CH3CN and 0.5 mL of HBr (4.4 mmol). The two solutions were stirred at 40 °C. From the final solution, the solvent was slowly evaporated until a white powder formed.

2.2.3 Synthesis of (2,2,4(2,4,4)-trimethyl-1,6-hexanediamine)PbCl4 (m3)

181 μL of 2,2,4(2,4,4)-trimethyl-1,6-hexanediamine (1 mmol) and 40 μL of 37% aqueous hydrochloric acid was mixed (0.47 mmol). 0.278 g of PbCl2 (1 mmol) was dissolved in a mixture of 1.2 mL DMSO and 40 μL of HCl (0.47 mmol). The two solutions were mixed under stirring at 40 °C until a white precipitate was formed. The solution was left to dry by slow evaporation of the solvent.

2.3 X-ray powder diffraction

The X-ray powder diffraction data (XRPD) were obtained from polycrystalline samples at room temperature on a Bruker D8 Advance diffractometer equipped with a LynxEye® detector and Ni-filtered Cu radiation. The scanning area covered the 2θ range of 2°–80°, with a scanning angle step size of 0.015° and a time step of 0.161 s. Structural analysis was carried out using the FullProf software [58].

2.4 Scanning electron microscopy and EDX

The SEM images and EDX data were recorded on an EVO-MA 10 Carl Zeiss instrument equipped with a 129 eV resolution INCAx-act Silicon Drift Detector. EDX spectra were acquired at 15 kV and images at 5 kV accelerating voltage. Samples were measured without the application of any conductive coating.

2.5 Optical measurements

The UV/Vis optical absorption (OA) spectra were recorded on a UV-1800 UV Shimadzu spectrophotometer in the range of 200–800 nm, at a sampling step of 0.5 nm using 1.5 nm slits, a combination of halogen and D2 lamps as sources. The samples were measured as thin or thick spin-coated films on quartz or ITO substrates, after subtracting the substrate’s spectra as reference. The photoluminescence (PL) and photoluminescence excitation (PLE) spectra were obtained from thick deposits on quartz plates, mounted in a Hitachi F-2500 FL spectrophotometer employing a xenon 150 W lamp and an R928 photomultiplier. All the above data were obtained at room temperature. For the measurements at T = 77 K liquid N2 was used as a blanket to a sealed quartz tube hosting the materials under argon atmosphere.

3 Results and discussion

3.1 Crystal structures

The Powder XRD patterns of m1, m2 and m3 have been indexed, where all compounds crystallize with monoclinic symmetry with strong resemblance to the 2D hybrid semiconductor [NH3(CH2)6NH3]PbI4 [59] which is presented in Figure 1, and are presented in Figure 2 along with the latter’s compound computed XRD pattern. No peaks were assigned to unreacted PbCl2, PbBr2 or PbI2. However, due to the low symmetry of the perovskite compounds and the lack of high-quality single crystals, the XRPD patterns are not suitable for a full structural Rietveld analysis to determine the atomic positions.

Figure 1: 
Crystal structure of [NH3(CH2)6NH3]PbI4 from ref. [59] (COD ID 7203879). Unit cell edges are denoted with dark blue lines, whereas the corner-sharing [PbI6]4– octahedra edges are denoted with violet lines.
Figure 1:

Crystal structure of [NH3(CH2)6NH3]PbI4 from ref. [59] (COD ID 7203879). Unit cell edges are denoted with dark blue lines, whereas the corner-sharing [PbI6]4– octahedra edges are denoted with violet lines.

Figure 2: 
Powder XRD patterns with normalized diffraction intensities at room temperature for m1 (green), m2 (red) and m3 (blue) and the computed powder XRD of the compound in Figure 1 (magenta).
Figure 2:

Powder XRD patterns with normalized diffraction intensities at room temperature for m1 (green), m2 (red) and m3 (blue) and the computed powder XRD of the compound in Figure 1 (magenta).

Peaks at 2θ of ca. 6° and ca. 12° ± 0.1° are associated with the inorganic interlayer distance of the 2D perovskites according to literature [60], [61], [62] and assigned to the (100) and (200) planes, respectively, in compounds m1, m2 and m3. It is worth observing that the position of these peaks is almost independent from the choice of the halogen atom since the interlayer distance, giving rise to the strongest XRD peak, is mostly affected by the organic spacer and not by the halogen atom. As 2,2,4(2,4,4)-trimethyl-1,6-hexanediamine is significantly bulkier than the 1,6-hexanediamine, larger d-spacings are expected for the (h00) reflections (assuming that a axis is vertical to the inorganic layers), thus, shifting the 2θ angles to lower values than in [NH3(CH2)6NH3]PbI4. Very recently, a similar 2D crystal structure was reported for the 2D-(MBA)2PbI4 (MBA = chiral R/S-methylbenzylammonium) which produces a similar diffraction pattern to the title compounds [63] as well as a similar work by Ma et al. [64] where chirality further induces more exotic light–matter interaction phenomena.

The amine configuration has been optimized with ab initio dynamics (Supplementary Tables 2 and 3) in order to check for self-consistency in the XRD large plane spacings. The computed average protonated amine lengths plus the halogen-Pb-halogen distances amounts to almost the experimentally deduced d values, as derived from the first low angle peak in the Powder XRD patterns.

In the Supplementary Material, the powder XRD patterns (Supplementary Figures S1, S5 and S9) for the iodine synthesis variations are presented. The low angle peaks presented, indicate in all cases a 2D superstructure. Sample m1_ns3 (Supplementary Figure S5, which is similar to m1 besides the extra HI and PbI2 but within the ideal stoichiometry) has two sharp XRD peaks at low angles showing the formation of a unit cell with a smaller large axis by ca. 1.4 Å smaller. It is possible that this formation is due to the existence of two different phases derived from the 2,2,4 and 2,4,4 amine packings, which differ as found in the computed ab initio isomer lengths, described in the Supplementary Material section as the distance between the N atoms on each isomer.

The as prepared powders of the materials were stored in the dark at room temperature for a period of three years. The samples were dried before characterizing them. Compound m1 appeared to decompose at high drying temperatures (∼70 °C). The m2 and m3 materials were dehydrated by partial thermal treatment in a desiccator set at 120 °C, after which these were reversed possibly from hydrates to non-hydrates. The color of the powders did not change after the thermal treatment. Figures 3 and 4 present the XRD patterns for compounds m2 and m3, respectively, as prepared and after the aforementioned storage period and the thermal treatment. It is possible that H2O molecules were partially intercalated in the perovskite structure, so that the organic–inorganic framework was retained, but the X-ray diffraction pattern became more diffuse, as in the case of CH3NH3PbI3 upon hydration [65]. For both compounds the first low angle peak has become more intense and the other peaks seem to be minimized while new peaks have appeared. The normal degradation appears to induce change mostly to the inorganic layers, while the organic spacer is not altered significantly.

Figure 3: 
XRD patterns of m2 (a) as prepared and (b) after three years.
Figure 3:

XRD patterns of m2 (a) as prepared and (b) after three years.

Figure 4: 
XRD patterns of m3 (a) as prepared and (b) after three years.
Figure 4:

XRD patterns of m3 (a) as prepared and (b) after three years.

3.2 Electron microscopy

SEM images of all three compounds as prepared are presented in Supplementary (Figure S13). In particular, Supplementary Figure S13 (a, b), (c, d) and (e, f) show images of the as prepared compounds m1, m2 and m3, respectively, where some pin holes are observed and a flower like structure, while it is also seen that the anion substitution alters the perovskite morphology. Compound m1 forms a grain-like microstructure while its variants m1_ns2 (Supplementary Figure S2) and m1_ns3 (Supplementary Figure S6) form needles on top of the flower-like microstructure. The morphology of the variant m1_ns4 appears to be plate-like (Supplementary Figure S10). On the contrary, compounds m2 and m3 form needle-like microstructures (Supplementary Figure S13). Should a synthesis proceed with a surplus of the amine and smaller quantity of hydroiodic acid, the material assumes a soft-plastic texture.

Comparison of the SEM images of m2 and m3 are presented further below, before and after the three-year storage period except for m1 since it was the only one that degraded beyond characterization capability. In Figure 5, m2 is presented, where after drying the samples a more well-formed cubic shaped microstructure is observed.

Figure 5: 
SEM images of m2 as synthesized (a–c) and dried (d–f).
Figure 5:

SEM images of m2 as synthesized (a–c) and dried (d–f).

A comparison of the SEM images of m3 (Figure 6) allowed us to confirm that the morphology is quite similar except for some tetragonal particles observed in the as-synthesized compound (red circles). After drying and overnight storing of the compound under vacuum, the small tetragonal units disappeared and a needle-like microstructure had formed.

Figure 6: 
SEM images of m3 as synthesized (a–c) and dried (d–f).
Figure 6:

SEM images of m3 as synthesized (a–c) and dried (d–f).

According to EDX measurements (Supplementary Figures S14–S16), all compounds showed a molar ratio Pb:X (X = I, Br, Cl) of 1:4. This is in agreement with the results of the powder XRD measurements (Figure 2) that indicate that the materials form 2D arrangements, where the corner-sharing topology leads to the aforementioned stoichiometry.

3.3 Optical studies

The optoelectronic properties of the compounds were investigated using UV/Vis spectroscopy (Figure 7) at room temperature. The optical absorption (OA) spectra reveal excitonic peaks at 494, 400 and 332 nm for compounds m1, m2 and m3, respectively; for compound m2 the peak is not as strong or sharp. Compound m1 contains iodine which is larger and less electronegative than bromine and chlorine in compounds m2 and m3, respectively, resulting to an excitonic red shift, in agreement with previous studies of similar compounds [59, 66, 67]. In all three compounds, a second peak is observed at higher energies relative to the n = 1 excitonic peak. These low wavelength excitonic peak positions at 318 and 258 nm correspond to the absorption of the PbX64− standalone units of the analogous Br and Cl compounds [2], respectively, which are, however, not completely isolated but slightly interact with the environment, as for example the isolated PbBr64− peak would be located at ca. 308 nm [68].

Figure 7: 
UV/Vis OA spectra of m1, m2, and m3 as prepared.
Figure 7:

UV/Vis OA spectra of m1, m2, and m3 as prepared.

The variant m1_ns2 shows excitonic peaks at 487 nm (Supplementary Figure S3), due to the small amount of lead precursor used for this variant; possibly only small sized platelets formed leading to a considerable blue shift size effect. Sample m1_ns4 exhibits peak at 502 nm (Supplementary Figure S11), that are slightly shifted relative to m1, which probably can be attributed to steric effects of the amine and “loose” hydrogen bonding and their modulation on energy band gaps of iodoplumbates [69, 70]. Variant m1_ns3 exhibits excitonic peak at 363 nm (Supplementary Figure S7), as noted before, as well as a weak 2D excitonic peak at 506 nm. This shows that minute changes, i.e. doubling the precursor HI moles, lead to the formation of 0D entities of the PbI64− type and the minimization of the number of the 2D inorganic sheets.

In Figure 8, the UV/Vis spectra of m2 as prepared and aged samples are presented. The excitonic peak at ca. 400 nm is still present but the peak at 310 nm has strongly decreased, which implies that after the thermal dehydration of the stored compound m2, the peak corresponding to the 0D isolated lead bromide octahedra has been degraded, a phenomenon that has been observed before for similar compounds [71]. According to the SEM images (Figure 5), m2 after its long storage has in fact formed better crystals. Thus, the m2 compound has not been degraded by oxygen and humidity, while compound m3 has been affected as discussed further below.

Figure 8: 
UV/VIS OA spectra of m2 as prepared (a) and after three years storage (b).
Figure 8:

UV/VIS OA spectra of m2 as prepared (a) and after three years storage (b).

A peculiar behavior is observed for compound m3 after its prolonged storage. As it can be observed in Figure 9, the peak at 332 nm has been minimized and the peak at 286 nm has been widened and enhanced, both remaining despite the loss of the 257 nm peak. It is possible that the 2D perovskite structure had its superstructure broken down along the long a axis to accommodate both the organic cation and any H2O molecules trapped in between, and this has partially degraded the inorganic layers [72], [73], [74], since the organic spacers here are partially hydrophilic.

Figure 9: 
UV/VIS OA spectra of m3 as prepared (a) and after three years storage (b).
Figure 9:

UV/VIS OA spectra of m3 as prepared (a) and after three years storage (b).

By extrapolating the linear part of the Tauc graphs for the optical absorption, it is possible to get an estimate of the energy band gap values (Supplementary Figure S17) of the as synthesized compounds m1, m2 and m3; these were computed to be 2.82, 3.29 and 3.92 eV, respectively (see Supplementary Material). Obtaining a rough experimental estimate of the energy band gap values, from the first minimum of the OA spectra at higher energies than the excitonic peak, Eg values are estimated at 2.69, 3.28 and 4.0 eV for m1, m2 and m3, respectively, leading to experimental excitonic binding energies of 185, 195 and 253 meV, respectively.

Figures 1012 present the photoluminescence (PL) spectra of m1, m2 and m3, respectively, while Figures 13 and 14 compare the time variation of the PL of m2 and m3, respectively, after their prolonged storage. In all three compounds, broad bands are observed rather than sharp excitonic peaks, before and after their storage. PL emission spectra exhibit excitonic broad bands as a double peak at 512 and 531 nm for m1, 414 nm for m2 and 346 nm for m3. Comparing the OA and PL emission spectra, the excitonic peaks appear to have an average Stokes shift of 15 nm for all compounds, indicative of only few defect crystalline states of all compounds [75]. The double peak for m1 is possibly attributed to two different existing structural variants, exhibiting, however, the same XRD pattern, or even more probably due to defects. It is interesting that the as prepared m3 shows also a sharp peak at 400 nm and a broad one centered at 465 nm, both latter cannot be linked to other experimental observations of similar materials.

Figure 10: 
OA (red) and PL (green) spectra of m1.
Figure 10:

OA (red) and PL (green) spectra of m1.

Figure 11: 
OA (red) and PL (green) spectra of m2.
Figure 11:

OA (red) and PL (green) spectra of m2.

Figure 12: 
OA (red) and PL (green) spectra of m3.
Figure 12:

OA (red) and PL (green) spectra of m3.

Figure 13: 
PL spectra of m2 as prepared (b) and after three years storage (a).
Figure 13:

PL spectra of m2 as prepared (b) and after three years storage (a).

Figure 14: 
PLE (left) and PL (right) spectra of m3 as prepared (a) and after three years (b).
Figure 14:

PLE (left) and PL (right) spectra of m3 as prepared (a) and after three years (b).

As seen in Figure 13, compound m2, presented minor alterations in the photoluminescence peaks after the three years storage period. A more intense and sharper peak has appeared which is slightly blue-shifted to ca. 408 nm in contrast to the as prepared sample, ca. 414 nm, yet both spectra are quite wide. The PL spectra of the as prepared sample has been enhanced in order to be comparable to the aged one, which are comparable due to the similar geometry used for both.

For compound m3, the PL spectra have revealed major differences after the three-year storage period (Figure 14). According to the OA spectra, m3 retains the 2D excitonic peaks, but the as prepared sample exhibits the PL peak at 347 nm, while the stored sample exhibits a broad and red-shifted excitonic PL peak to 363 nm, due to defects, and at the same time the 400 nm PL peak has disappeared. The PLE spectra, however, which should resemble the OA spectra, show that the spectra of both aged and fresh m3 have a strong peak at 369 nm, probably denoting some defect states responsible for the low energy emission at 550 nm.

In Figure 15, 3D PL spectra of compound m2 after being exposed to humidity and oxygen are presented. A blue up to green luminescence is observed that may be attributed to defects in the crystal structure in accordance to the data of Figure 13. According to Figure 6 the SEM images show more well-formed crystals with more sharp edges after the samples were exposed to humidity, as compared to the as synthesized sample which presents a smoother surface.

Figure 15: 
3D PL spectra of m2 at room temperature after three years of storage.
Figure 15:

3D PL spectra of m2 at room temperature after three years of storage.

In Figure 16, 3D PL spectra of m3 after being exposed to humidity and air are presented. A broad blueish-green luminescence is observed for a variety of excitation wavelengths from 260 nm up to 400 nm, evident of white like luminescence.

Figure 16: 
3D PL spectra of m3 at room temperature after three years of storage.
Figure 16:

3D PL spectra of m3 at room temperature after three years of storage.

In the Supplementary Material, PL and PLE spectra of m1_ns2, m2 and m3 are provided, measured at 77 K for the as prepared compounds. These measurements were performed to investigate phenomena responsible for the low PL intensities observed compared to similar 2D lead halide perovskites. The PL spectra at 77 K of some of m1_ns2 show a single peak at 509 nm (Supplementary Figure S18), slightly blue-shifted in comparison to the RT spectra (Supplementary Figure S8). These may be due to some structural variation, while the PLE spectra agree with the RT OA spectra as far as the exciton is concerned. The 77 K spectra of m1_ns2 also display a shoulder at ca. 600 nm, probably due to complex defect states linked to the Pb+2 ion in a halogen environment, as for example seen in ref. [76]. Compound m2 shows a peak at ca. 410 nm in the RT spectra (Figure 11) as well as at 77 K, however, the PLE spectra at 77 K reveals a characteristic absorption at 300 nm (Supplementary Figure S19) as well as a 368 nm PLE peak. The 300 nm PLE peak was used to collect a PL spectra with λexc = 300 nm, that shows that at 77 K this excitation yields a broad exciton peak centered at 407 nm as observed in Supplementary Figure S19. Exciting m2 with higher or lower energies, i.e. 280 or 320 nm, provides PL spectra with less intensity than using 300 nm excitation. It is possible that the degraded m2 has quantum confined sheets that instead of showing absorption at ca. 330 nm they have blue shifted at lower wavelengths, thus, using 300 nm radiation allows the broad 2D excitonic emission by virtue energy transfer. It is suggested that 300 nm radiation induces photoluminescence at high energies which is transferred to the remaining un-degraded 2D sheets to be emitted mainly as 407 nm radiation. In this process, all other radiative processes yield the broad PL peak of 407 nm. All these elucidate the complexity and role of the defects and multi-structural diversity of the lead halide perovskites. It appears that this perovskite could emit broad UV-blue light if excited with a GaN LED or electrically activated.

Finally, m3 shows an excitation dependent PL spectrum at 77 K (Supplementary Figure S20), centered far from the exciton peak at 500–560 nm, while the PLE spectrum reveals that absorptions at 303, 333, 368 nm appear to be responsible for this wide low temperature PL peak. It is conjectured that this PL behavior is due to a mechanism with which the exciton formed by the high energy excitation transfers its energy to defects of lower energy or transfer it towards the organic–inorganic artifacts which could form by lattice interruption introduced by the cooling process. Overall, it can be assessed that these types of materials, which deviate from the standard LD HOIS perovskites, have a quite complex behavior. Finally, it is important to note the similarity of the room temperature absorption spectra for the 2D materials in this work and those in refs. [63] and [64] based on RMX4 which are broad rather than sharp as it usual observed in 2D lattices. It appears, however, that the use of the mixture of 2,2,4 and 2,4,4 isomers as well as including their enantiomers provide a sharper exciton peak at 494 nm.

4 Conclusions

In the present paper, the synthesis of 2D lead halide perovskites using an isomer mixture of a diamine is presented. The as-synthesized materials were characterized in terms of their XRD patterns and their optical spectra. All materials appear to have less pronounced PL signals than the commonly known 2D perovskites, while in the case of the iodide, structural variants can be found. The compounds were stored in the dark under low humidity and re-examined after three years, where the iodine compound partially decomposed without retaining its properties. The bromide and chloride aged analogues needed to be thermally treated in order to recover some of their optical properties, probably due to the inclusion and subsequent removal of H2O molecules within the hydrophilic amine framework which takes place without separating the organic component. The aged chloride compound displays white-like luminescence at room temperature which is desired for the use in OLEDs. It is suggested that the packing of the structurally slightly different amines can in fact create variants in the iodine analogue, due to the larger number of packing possibilities while in the case of the bromine and chlorine the isomers build a 2D organic barrier which is the average of the two isomers. It is possible that this poor stacking as well as the bulky cation may create variations in the inorganic layer, partially responsible for the decreased PL intensities. Some discussion also includes the comparison of the results to other based on 2D chiral lead-iodide-based hybrid organic inorganic semiconductors.

5 Supporting information

Additional synthesis information, powder X-ray diffraction patterns, SEM images, EDX spectra and other supporting data associated with this article can be found as supplementary material in the online version (https://doi.org/10.1515/znb-2021-0090).


Corresponding author: Ioannis Koutselas, Materials Science Department, School of Natural Sciences,University of Patras, Patras, 26504, Greece, E-mail:

Funding source: European Union’s Horizon 2020

Award Identifier / Grant number: 861985

  1. Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: This research received partly funding from the European Union’s Horizon 2020 program PeroCUBE through an Innovation Action under Grant Agreement No. 861985.

  3. Conflict of interest statement: The authors declare that they have no conflicts of interest.

References

1. Wells, H. L. Z. Anorg. Allg. Chem. 1893, 3, 195–210; https://doi.org/10.1002/zaac.18930030124.Suche in Google Scholar

2. Papavassiliou, G. C. Prog. Solid State Chem. 1997, 25, 125–270; https://doi.org/10.1016/s0079-6786(97)80886-2.Suche in Google Scholar

3. Stranks, S. D., Snaith, H. J. Nat. Nanotechnol. 2015, 10, 391–402; https://doi.org/10.1038/nnano.2015.90.Suche in Google Scholar PubMed

4. Chen, J., Du, W. N., Shi, J. W., Li, M. L., Wang, Y., Zhang, Q., Liu, X. F. InfoMat 2019, 2, 170–183; https://doi.org/10.1002/inf2.12051.Suche in Google Scholar

5. Wang, R., Mujahid, M., Duan, Y., Wang, Z., Xue, J., Yang, Y. Adv. Funct. Mater. 2019, 29, 1808843; https://doi.org/10.1002/adfm.201808843.Suche in Google Scholar

6. Zhang, J., Yang, X., Deng, H., Qiao, K., Farooq, U., Ishaq, M., Yi, F., Liu, H., Tang, J., Song, H. Nano-Micro Lett. 2017, 9, 36; https://doi.org/10.1007/s40820-017-0137-5.Suche in Google Scholar PubMed PubMed Central

7. Hwang, J., Rao, R. R., Giordano, L., Katayama, Y., Yu, Y., Shao-Horn, Y. Science 2017, 358, 751–756; https://doi.org/10.1126/science.aam7092.Suche in Google Scholar PubMed

8. Royer, S., Duprez, D., Can, F., Courtois, X., Batiot-Dupeyrat, C., Laassiri, S., Alamdari, H. Chem. Rev. 2014, 20, 10292–10368; https://doi.org/10.1021/cr500032a.Suche in Google Scholar PubMed

9. Kojima, A., Teshima, K., Shirai, Y., Miyasaka, T. J. Am. Chem. Soc. 2009, 17, 6050–6051; https://doi.org/10.1021/ja809598r.Suche in Google Scholar PubMed

10. NREL. https://www.nrel.gov/pv/assets/pdfs/best-research-cell-efficiencies-rev210726.pdf (accessed Sep 2021).Suche in Google Scholar

11. Eperon, G. E., Leijtens, T., Bush, K. A., Prasanna, R., Green, T., Wang, J. T.-W., McMeekin, D. P., Volonakis, G., Milot, R. L., May, R., Palmstrom, A., Slotcavage, D. J., Belisle, R. A., Patel, J. B., Parrott, E. S., Sutton, R. J., Ma, W., Moghadam, F., Conings, B., Babayigit, A., Boyen, H. -G., Bent, S., Giustino, F., Herz, L. M., Johnston, M. B., McGehee, M. D., Snaith, H. J. Sci. 2016, 6314, 861–865; https://doi.org/10.1126/science.aaf9717.Suche in Google Scholar PubMed

12. McMeekin, D. P., Sadoughi, G., Rehman, W., Eperon, G. E., Saliba, M., Horantner, M. T., Haghighirad, A., Sakai, N., Korte, L., Rech, B., Johnston, M. B., Herz, L. M., Snaith, H. J. Sci. 2016, 351, 151–155; https://doi.org/10.1126/science.aad5845.Suche in Google Scholar PubMed

13. Karani, A., Yang, L., Bai, S., Futscher, M. H., Snaith, H. J., Ehrler, B., Greenham, N. C., Di, D. ACS Energy Lett. 2018, 3, 869–874; https://doi.org/10.1021/acsenergylett.8b00207.Suche in Google Scholar

14. Mazzarella, L., Lin, Y., Kirner, S., Morales-Vilches, A. B., Korte, L., Albrecht, S., Crossland, E., Stannowski, B., Case, C., Snaith, H. J., Schlatmann, R. Adv. Energy Mater. 2019, 14, 1803241; https://doi.org/10.1002/aenm.201803241.Suche in Google Scholar

15. Young, K., Yang, T. Y., Suhonen, R., Kemppainen, A., Hwang, K., Jeon, N. J., Seo, J. Nat. Commun. 2020, 11, 5146–5156; https://doi.org/10.1038/s41467-020-18940-5.Suche in Google Scholar PubMed PubMed Central

16. Li, H., Zuo, C., Scully, A. D., Angmo, D., Yang, J., Gao, M. Flex. Print. Electron. 2020, 5, 014006; https://doi.org/10.1088/2058-8585/ab639e.Suche in Google Scholar

17. Gusain, A., Thankappan, A., Thomas, S. J. Mater. Sci. 2020, 55, 13490–13542; https://doi.org/10.1007/s10853-020-04883-1.Suche in Google Scholar

18. Qiu, L., Deng, J., Lu, X., Yang, Z., Peng, H. Angew. Chem. Int. Ed. 2014, 53, 1–5; https://doi.org/10.1002/anie.201404973.Suche in Google Scholar PubMed

19. Lee, M., Ko, Y., Jun, Y. J. Mater. Chem. A 2015, 3, 19310–19313; https://doi.org/10.1039/c5ta02779a.Suche in Google Scholar

20. Huang, X., Huang, X., Guo, Q., Yang, D., Xiao, X., Liu, X., Xia, Z., Fan, F., Qiu, J., Dong, G. Nat. Photonics 2020, 14, 82–88; https://doi.org/10.1038/s41566-019-0538-8.Suche in Google Scholar

21. Hu, J., Wang, C., Qiu, S., Zhao, Y., Gu, E., Zeng, L., Yang, Y., Li, C., Liu, X., Forberich, K., Brabec, C. J., Nazeeruddin, M. K., Mai, Y., Guo, F. Adv. Energy Mater. 2020, 10, 2000173; https://doi.org/10.1002/aenm.202000173.Suche in Google Scholar

22. Li, Z., Li, Z., Li, P., Chen, G., Cheng, Y., Pi, X., Yu, X., Yang, D., Han, L., Zhang, Y., Song, Y. ACS Appl. Mater. Interfaces 2020, 12, 39082–39091; https://doi.org/10.1021/acsami.0c09485.Suche in Google Scholar PubMed

23. Nikolaou, P., Vareli, I., Deskoulidis, E., Matsoukas, J., Vassilakopoulou, A., Koutselas, I., Topoglidis, E. J. Solid State Chem. 2019, 273, 17–24; https://doi.org/10.1016/j.jssc.2019.02.018.Suche in Google Scholar

24. Liu, Y., Zheng, Y., Zhu, Y., Ma, F., Zheng, X., Yang, K., Zheng, X., Xu, Z., Ju, S., Zheng, Y., Guo, T., Qian, L., Li, F. ACS Appl. Mater. Interfaces 2020, 12, 39649–39656; https://doi.org/10.1021/acsami.0c11103.Suche in Google Scholar PubMed

25. Tiley, R. J. D. Perovskites Structure-Property Relationships; John Wiley & Sons: Hoboken, 2016.10.1002/9781118935651Suche in Google Scholar

26. Eperon, G. E., Stranks, S. D., Menelaou, C., Johnston, M. B., Herz, L. M., Snaith, H. J. Energy Environ. Sci. 2014, 7, 982–988; https://doi.org/10.1039/c3ee43822h.Suche in Google Scholar

27. Li, Y., Liu, F. Z., Waqas, M., Leung, T. L., Tam, H. W., Lan, X. Q., Tu, B., Chen, W., Djurisic, A. B., He, Z. B. Small Methods 2018, 2, 1700387; https://doi.org/10.1002/smtd.201700387.Suche in Google Scholar

28. Hong, K., Le, Q. V., Kim, S. Y., Jang, H. W. J. Mater. Chem. C 2018, 6, 2189–2209; https://doi.org/10.1039/c7tc05658c.Suche in Google Scholar

29. Mohd Yusoff, A. R. B., Gao, P., Nazeeruddin, M. K. Coord. Chem. Rev. 2018, 373, 258–294; https://doi.org/10.1016/j.ccr.2017.10.021.Suche in Google Scholar

30. Ogomi, Y., Morita, A., Tsukamoto, S., Saitho, T., Fujikawa, N., Shen, Q., Toyoda, T., Yoshino, K., Pandey, S. S., Ma, T., Hayase, S. J. Phys. Chem. Lett. 2014, 5, 1004–1011; https://doi.org/10.1021/jz5002117.Suche in Google Scholar PubMed

31. Stoumpos, C. C., Malliakas, C. D., Kanatzidis, M. G. Inorg. Chem. 2013, 52, 9019–9038; https://doi.org/10.1021/ic401215x.Suche in Google Scholar PubMed

32. Shamsi, J., Urban, A. S., Imran, M., Trizio, L. D., Manna, L. Chem. Rev. 2019, 119, 3296–3348; https://doi.org/10.1021/acs.chemrev.8b00644.Suche in Google Scholar PubMed PubMed Central

33. Dai, X., Xu, K., Wei, F. Beilstein J. Nanotechnol. 2020, 11, 51–60; https://doi.org/10.3762/bjnano.11.5.Suche in Google Scholar PubMed PubMed Central

34. Li, X., Zhang, F., He, H., Berry, J. J., Zhu, K., Xu, T. Nature 2020, 578, 555–558; https://doi.org/10.1038/s41586-020-2001-x.Suche in Google Scholar PubMed

35. Gao, P., Grätzel, M., M. K. Nazeeruddin Energy Environ. Sci. 2014, 7, 2448–2463; https://doi.org/10.1039/c4ee00942h.Suche in Google Scholar

36. Yin, W., Yang, J., Kang, J., Yan, Y., Wei, S. J. Mater. Chem. A 2015, 3, 8926–8942; https://doi.org/10.1039/c4ta05033a.Suche in Google Scholar

37. Chen, Q., Marco, N. D., Yang, Y. M., Song, T. B., Chen, C. C., Zhao, H., Hong, Z., Zhou, H., Yang, Y. Nano Today 2015, 10, 355–396; https://doi.org/10.1016/j.nantod.2015.04.009.Suche in Google Scholar

38. Mercier, N. Angew. Chem. Int. Ed. 2019, 58, 2–8; https://doi.org/10.1002/anie.201980511.Suche in Google Scholar

39. Saparov, B., Mitzi, D. B. Chem. Rev. 2016, 116, 4558–4596; https://doi.org/10.1021/acs.chemrev.5b00715.Suche in Google Scholar PubMed

40. Lin, H., Zhou, C., Tian, Y., Siegrist, T., Ma, B. ACS Energy Lett. 2018, 3, 54–62; https://doi.org/10.1021/acsenergylett.7b00926.Suche in Google Scholar

41. Kahwagi, R. T., Thornton, S. T., Smith, B., Koleilat, G. I. Front. Optoelectron. 2020, 13, 196–224; https://doi.org/10.1007/s12200-020-1039-6.Suche in Google Scholar

42. Papavassiliou, G. C., Koutselas, I. B. Synth. Met. 1995, 71, 1713–1714; https://doi.org/10.1016/0379-6779(94)03017-z.Suche in Google Scholar

43. Abbas, M. S., Hussain, S., Zhang, J., Wang, B., Yang, C., Wang, Z., Wei, Z., Ahmad, R. Sustain Energy Fuels 2020, 4, 324–330; https://doi.org/10.1039/c9se00817a.Suche in Google Scholar

44. Kim, B., Seok, S. I. Energy Environ. Sci. 2020, 13, 805–820; https://doi.org/10.1039/C9EE03473K.Suche in Google Scholar

45. Hautzinger, M. P., Pan, D., Pigg, A. K., Fu, Y., Morrow, D. J., Leng, M., Kuo, M. Y., Spitha, N., Lafayette, D. P., Kohler, D. D., Wright, J. C., Jin, S. ACS Energy Lett. 2020, 5, 1430–1437; https://doi.org/10.1021/acsenergylett.0c00450.Suche in Google Scholar

46. Li, X., Fu, Y., Pedesseau, L., Guo, P., Cuthriell, S., Hadar, I., Even, J., Katan, C., Stoumpos, C. C., Schaller, R. D., Harel, E., Kanatzidis, M. G. J. Am. Chem. Soc. 2020, 142, 11486–11496; https://doi.org/10.1021/jacs.0c03860.Suche in Google Scholar PubMed

47. Masada, S., Yamada, T., Tahara, H., Hirori, H., Saruyama, M., Kawasaki, T., Sato, R., Teranishi, T., Kanemitsu, Y. Nano Lett. 2020, 5, 4022–4028; https://doi.org/10.1021/acs.nanolett.0c01417.Suche in Google Scholar PubMed

48. Chu, Z., Zhao, Y., Ma, F., Zhang, C. X., Deng, H., Gao, F., Ye, Q., Meng, J., Yin, Z., Zhang, X., You, J. Nat. Commun. 2020, 11, 4165; https://doi.org/10.1038/s41467-020-17943-6.Suche in Google Scholar PubMed PubMed Central

49. Fu, Y., Jiang, X., Li, X., Traore, B., Spanopoulos, I., Katan, C., Even, J., Kanatzidis, M. G., Harel, E. J. Am. Chem. Soc. 2020, 142, 4008–4021; https://doi.org/10.1021/jacs.9b13587.Suche in Google Scholar PubMed

50. Smith, M. D., Connor, B. A., Karunadasa, H. I. Chem. Rev. 2019, 119, 3104–3139; https://doi.org/10.1021/acs.chemrev.8b00477.Suche in Google Scholar PubMed

51. Mao, L., Stoumpos, C. C., Kanatzidis, M. G. J. Am. Chem. Soc. 2019, 141, 1171–1190; https://doi.org/10.1021/jacs.8b10851.Suche in Google Scholar PubMed

52. Leblanc, A., Mercier, N., Allain, M., Dittmer, J., Pauporté, T., Fernandez, V., Boucher, F., Kepenekian, M., Katan, C. ACS Appl. Mater. Interfaces 2019, 11, 20743–20751; https://doi.org/10.1021/acsami.9b00210.Suche in Google Scholar PubMed

53. Miyata, K., Atallah, T. L., Zhu, X. Y. Sci. Adv. 2017, 3, e1701469; https://doi.org/10.1126/sciadv.1701469.Suche in Google Scholar PubMed PubMed Central

54. Zhu, H., Miyata, K., Fu, Y., Wang, J., Joshi, P. P., Niesner, D., Williams, K. W., Jin, S., Zhu, X. Y. Science 2016, 353, 1409–1413; https://doi.org/10.1126/science.aaf9570.Suche in Google Scholar PubMed

55. Hutter, E. M., Gélvez-Rueda, M. C., Osherov, A., Bulović, V., Grozema, F. C., Stranks, S. D., Savenije, T. J. Nat. Mater. 2017, 16, 115–120; https://doi.org/10.1038/nmat4765.Suche in Google Scholar PubMed

56. Zhu, H., Trinh, M. T., Wang, J., Fu, Y., Joshi, P. P., Miyata, K., Jin, S., Zhu, X. Y. Adv. Mater. 2017, 29, 1603072; https://doi.org/10.1002/adma.201603072.Suche in Google Scholar PubMed

57. Lv, Y., Ma, H., Yin, Y., Dong, Q., Zhao, W., Jinc, S., Shi, Y. J. Mater. Chem. A 2020, 8, 10283–10290; https://doi.org/10.1039/d0ta02437f.Suche in Google Scholar

58. Rodríguez-Carvajal, J. FullProf (Version 5.30); Institut Laue-Langevin: Grenoble, France, 2012.Suche in Google Scholar

59. Mousdis, G. A., Papavassiliou, G. C., Raptopoulou, C. P., Terzis, A. J. Mater. Chem. 2000, 10, 515–518; https://doi.org/10.1039/a906161d.Suche in Google Scholar

60. Jia, G., Shi, Z. J., Xia, Y. D., Wei, Q., Chen, Y. H., Xing, G. C., Huang, W. Opt Express 2018, 26, A66–A74; https://doi.org/10.1364/oe.26.000a66.Suche in Google Scholar PubMed

61. Cortecchia, D., Neutzner, S., Kandada, A. R. S., Mosconi, E., Meggiolaro, D., De Angelis, F., Soci, C., Petrozza, A. J. Am. Chem. Soc. 2017, 139, 39–42; https://doi.org/10.1021/jacs.6b10390.Suche in Google Scholar PubMed

62. Lin, C. W., Liu, F., Chen, T. Y., Lee, K. H., Chang, C. K., He, Y., Leung, T. L., Ng, A. M. C., Hsu, C. H., Popović, J., Djurišić, A., Ahn, H. ACS Appl. Mater. Interfaces 2020, 12, 5008–5016; https://doi.org/10.1021/acsami.9b17881.Suche in Google Scholar PubMed

63. Lu, Y., Wang, Q., Chen, R., Qiao, L., Zhou, F., Yang, X., Wang, D., Cao, H., He, W., Pan, F., Yang, Z., Song, C. Adv. Funct. Mater. 2021, 3, 2104605; https://doi.org/10.1002/adfm.202104605.Suche in Google Scholar

64. Ma, J., Fang, C., Chen, C., Jin, L., Wang, J., Wang, S., Tang, J., Li, D. ACS Nano 2019, 13, 3659–3665; https://doi.org/10.1021/acsnano.9b00302.Suche in Google Scholar

65. Arakcheeva, A., Chernyshov, D., Spina, M., Forro, L., Horvath, E. Acta Crystallogr. B: Struct. Sci., Cryst. Eng. Mater. 2016, 72, 716–722; https://doi.org/10.1107/s2052520616010428.Suche in Google Scholar

66. Papavassiliou, G. C., Mousdis, G. A., Koutselas, I. B. Adv. Mater. Opt. Electron. 1999, 9, 265–271; https://doi.org/10.1002/1099-0712(199911/12)9:6<265::aid-amo390>3.0.co;2-6.10.1002/1099-0712(199911/12)9:6<265::AID-AMO390>3.0.CO;2-6Suche in Google Scholar

67. Mousdis, G. A., Papavassiliou, G. C., Terzis, A., Raptopoulou, C. P. Z. Naturforscher. 1998, 53b, 927–931; https://doi.org/10.1515/znb-1998-0825.Suche in Google Scholar

68. Zhang, L., Ju, M. G., Liang, W. Phys. Chem. Chem. Phys. 2016, 18, 23174–23183; https://doi.org/10.1039/c6cp01994c.Suche in Google Scholar

69. Nagasaka, H., Yoshizawa-Fujita, M., Takeoka, Y., Rikukawa, M. ACS Omega 2018, 3, 18925–18929; https://doi.org/10.1021/acsomega.8b02399.Suche in Google Scholar

70. Yu, T., Zhang, L., Shen, J., Fu, Y., Fu, Y. Dalton Trans. 2014, 43, 13115–13121; https://doi.org/10.1039/c4dt01181c.Suche in Google Scholar

71. Van Gompel, W. T. M., Herckens, R., Van Hecke, K., Ruttens, B., D’Haen, J., Lutsen, L., Vanderzande, D. ChemNanoMat 2019, 5, 323–327.10.1002/cnma.201800561Suche in Google Scholar

72. Mosconi, E., Azpiroz, J. M., De Angelis, F. Chem. Mater. 2015, 27, 4885–4892; https://doi.org/10.1021/acs.chemmater.5b01991.Suche in Google Scholar

73. Christians, J. A., Herrera, P. A. M., Kamat, P. V. J. Am. Chem. Soc. 2015, 137, 1530–1538; https://doi.org/10.1021/ja511132a.Suche in Google Scholar PubMed

74. Bonadio, A., de Oliveira, L. S., Polo, A. S., Souza, J. A. CrystEngComm 2019, 21, 7365–7372; https://doi.org/10.1039/c9ce01275c.Suche in Google Scholar

75. Smith, M. D., Watson, B. L., Dauskardt, R. H., Karunadasa, H. I. Chem. Mater. 2017, 29, 7083–7087; https://doi.org/10.1021/acs.chemmater.7b02594.Suche in Google Scholar

76. Novosad, S. S. Opt Spectrosc. 2000, 88, 522–526; https://doi.org/10.1134/1.626830.Suche in Google Scholar


Supplementary Material

The online version of this article offers supplementary material (https://doi.org/10.1515/znb-2021-0090).


Received: 2021-07-01
Accepted: 2021-08-18
Published Online: 2021-09-27
Published in Print: 2021-10-26

© 2021 Walter de Gruyter GmbH, Berlin/Boston

Heruntergeladen am 26.9.2025 von https://www.degruyterbrill.com/document/doi/10.1515/znb-2021-0090/html
Button zum nach oben scrollen