Next Article in Journal
Crystallization and Structural Properties of Oleogel-Based Margarine
Next Article in Special Issue
Electronic, Optical, Thermoelectric and Elastic Properties of RbxCs1−xPbBr3 Perovskite
Previous Article in Journal
Egg White-Mediated Fabrication of Mg/Al-LDH-Hard Biochar Composite for Phosphate Adsorption
Previous Article in Special Issue
Synthesis, Photoluminescence and Vibrational Properties of Aziridinium Lead Halide Perovskites
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Organic Cation on Optical Properties of [A]Mn(H2POO)3 Hybrid Perovskites

by
Dagmara Stefańska
Institute of Low Temperature and Structure Research, Polish Academy of Sciences, Okólna 2, 50-422 Wroclaw, Poland
Molecules 2022, 27(24), 8953; https://doi.org/10.3390/molecules27248953
Submission received: 14 November 2022 / Revised: 12 December 2022 / Accepted: 13 December 2022 / Published: 15 December 2022
(This article belongs to the Special Issue Recent Developments in Perovskite-Based Functional Materials)

Abstract

:
Hybrid organic–inorganic compounds crystallizing in a three-dimensional (3D) perovskite-type architecture have attracted considerable attention due to their multifunctional properties. One of the most intriguing groups is perovskites with hypophosphite linkers. Herein, the optical properties of six hybrid hypophosphite perovskites containing manganese ions are presented. The band gaps of these compounds, as well as the luminescence properties of the octahedrally coordinated Mn2+ ions associated with the 4T1g(G) → 6A1g(S) transition are shown to be dependent on the organic cation type and Goldschmidt tolerance factor. Thus, a correlation between essential structural features of Mn-based hybrid hypophosphites and their optical properties was observed. Additionally, the broad infrared luminescence of the studied compounds was examined for potential application in an indoor lighting system for plant growth.

1. Introduction

Hybrid organic–inorganic compounds with a three-dimensional (3D) perovskite structure constitute a very large family of materials. Their general formula is ABX3, where A is an alkali metal or an organic cation, B is a divalent metal cation, and X denotes an organic or inorganic anion (halide, HCOO, N3, CN, N(CN)2, or H2PO2). The 3D perovskite structure consists of corner-sharing BX6 (X = Cl, Br, I, O, or N) octahedral units. Metal ions and linkers form a three-dimensional framework, and organic cations are accommodated inside the created voids. Due to their unique physical phenomena, such as ferroelectricity, multiferroicity, and dielectricity, as well as magnetic and luminescent properties, these compounds have been extensively studied in recent years [1,2,3,4,5,6,7,8,9,10,11]. Hybrid organic–inorganic perovskites (HOIPs) are frequently called multifunctional materials because they exhibit several physical phenomena at the same time [12,13,14,15,16,17,18,19]. As a consequence, HOIPs offer a wide range of potential applications for catalysis, gas storage, drug delivery, nonlinear optics, solid-state lighting, luminescent sensors, temperature sensors, biomedical photonics, and numerous other utilizations [12,13,15,19,20,21,22].
Lead halides are one of the most commonly investigated families of HOIPs [4,12,23,24,25]. For instance, lead halides containing methylammonium (MA+) or formamidinium (FA+) cations attract great interest due to their promising application in solar cells [26,27,28,29], whereas methylhydrazinium (MHy+) lead halides have been reported to exhibit multiple nonlinear optical phenomena [4,24,30]. It is worth noting that the band gap value (Eg) and the position of the free exciton (FE) emission depend both on the halide used and the type of organic cation. Generally, the A-site organic cation is often ordered at low temperatures and becomes disordered as the temperature rises. Furthermore, organic cations have different sizes, shapes, and the ability to form hydrogen bonds (HBs) with the metal-ligand framework [31,32,33].
Three-dimensional hypophosphites are a relatively novel family of HOIPs. They were obtained for the first time in 2017 [34]. Unfortunately, the number of hypophosphite perovskites discovered is still low [34,35,36,37,38]. To date, only a few manganese, cadmium, and magnesium hypophosphite analogs have been reported [34,35,36,37,38]. Structural data revealed that metal–hypophosphite frameworks tend to have unconventional tilts, columnar shifts, and off-center positions of the organic cations [33,34], which may induce ferroelectricity or other functional properties [33,34,35]. Furthermore, some exhibit magnetic order, although this feature strictly depends on the type of organic cation, e.g., [MHy]Mn(H2POO)3 has an antiferromagnetic order near 6.5 K, [FA]Mn(H2POO)3 shows an antiferromagnetic order below 2.4 K, and [EA] Mn(H2POO)3 (EA+ = ethylammonium) exhibits a magnetic order at 7.0 K [36,38].
Manganese and cadmium hypophosphites can also show bright and broadband photoluminescence (PL) [38]. The optical properties of divalent manganese strongly depend on the crystal field. Due to the 3d5 configuration, Mn2+ electrons are in the outer orbital and are sensitive to changes in their immediate surroundings. Mn2+ in the tetrahedral coordination exhibits green emission characteristics for the ion in a weak crystal field, whereas in the octahedral site (strong crystal field), it possesses red infrared emission. It is well-known that the type of A-site cation strongly influences the structural and optical properties of hypophosphite perovskites. For this reason, the photoluminescence (PL) properties of six manganese hypophosphites, namely [DMA]Mn(H2POO)3, [EA]Mn(H2POO)3, [MHy]Mn(H2POO)3, [FA]Mn(H2POO)3, [Pyr]Mn(H2POO)3, and [IM]Mn(H2POO)3 (DMA+ = dimethylammonium, Pyr+ = pyrrolidinium, and IM+ = imidazolium) were investigated. Four ammonium cations are aliphatic, one is cyclic, and one is aromatic. Figure 1 presents the structural formulae and ionic radii of chosen ammonium cations. This paper presents some general remarks on the influence of the organic cations on the Eg value, PL band position, and thermal stability of the observed emission.

2. Results and Discussion

2.1. Crystal Structure

The difference in the ionic radius, the shape of the amine used, and the number of hydrogen bonds formed with the hypophosphite skeleton affect the crystal structure of the obtained compounds. Five of the six obtained compounds crystallize in the monoclinic system in the P21/n space group. Among these compounds, [FA]Mn(H2POO)3 is unique because it transforms at 178 K into another monoclinic system with the C2/c space group (polymorph I). Manganese ions occupy three non-equivalent positions in the [DMA] Mn(H2POO)3 structure; two positions in the EA, IM, and Pyr analogs; and one position in the FA compound. [MHy] Mn(H2POO)3 is the only hypophosphite that adopts the orthorhombic Pnma symmetry, with the manganese ions occupying only one independent site. Figure 2 presents a visualization of the crystal structures of the investigated hypophosphites. Further details of structural properties can be found in recently published papers [35,37,38], and in the next paragraph, the influence of the distortion of the nearest Mn2+ environment and the number of occupied sites on the PL properties of manganese hypophosphites is discussed. The description of sample synthesis and measurement technique is shown in Supplementary Materials.

2.2. Spectroscopic Properties

The absorption spectra of the investigated manganese hypophosphites were measured at room temperature. As can be seen in Figure 3a, they consist of strong host absorption at around 220 nm and less intense bands associated with the absorption of Mn2+ ions located in octahedral coordination. Thus, in Mn2+ ions, the absorption of the excitation energy occurs between the 6A1g ground state and the 4T1g(P) (275 nm), 4E(D) (339 nm), 4T2g(D) (358 nm), 4A1g and 4Eg(G) (404 nm), 4T2g(G) (435 nm), and 4T1g(G) (524 nm) excited levels [15,17,39]. The position of these bands is clearly visible for all obtained manganese hypophosphite perovskites. The optical band gap (Eg) was estimated using the Kubelka–Munk function. The determined Eg values are presented in the scheme shown in Figure 3b and listed in Table 1. The smallest Eg value of 4.98 eV is observed for [DMA]Mn(H2POO)3 and increases up to 5.32 eV for the [FA]Mn(H2POO)3 compound. At first glance, there is no relationship between the ionic radii of the ammonium cation and the Eg value. However, a more careful analysis shows that some general conclusions can be drawn. First of all, the samples containing linear amines and the samples comprising pyrrolidinium and imidazolium should be analyzed separately. Furthermore, it appears that the ionic radii of the chosen organic cation are insignificant; however, the value of the Goldschmidt tolerance factor (tF) is of major importance. This geometrical parameter estimates the ion size mismatches that the perovskite structure is able to tolerate. To calculate the tF parameters for the investigated manganese hypophosphites, the modified Goldschmidt equation was applied as appropriate for HOIPs [31,32]:
t F = r A e f f + r x e f f 2 r B + 0.5 h x e f f
where rB denotes the ionic radii of cation B in the perovskite structure, and rieff and hieff are the effective ionic radii and the effective height of the used organic amine (A) and ligand (X), respectively. The determined Eg values, together with the ionic radii of ammonium cations and tF values, are given in Table 1. As can be seen, the Eg of [A]Mn(H2POO)3 increases as tF decreases. In other words, the optical band gap is smaller for a more perfect perovskite structure. However, this relation is not valid for the cyclic or aromatic amines, as [Pyr]Mn(H2POO)3 has a slightly higher value of Eg than [IM]Mn(H2POO)3, in spite of its larger tF.
The normalized PL spectra of the prepared hybrid hypophosphites registered at 80 K are presented in Figure 4a. The PL bands are very broad and range from 600 nm to 900 nm. The full width at the half maximum (FWHM) value of these bands is approximately 2200 cm−1. After excitation under 266 nm, the electron moves to the 4T1g(P) excited level (Figure 4b). Then, as a result of non-radiative depopulation, the electron is transferred to the 4T1g(G) state. The radiative recombination of the electron from the 4T1g(G) level to the 6A1g(S) ground state generates intense red PL. The shape of the observed PL bands does not change with the applied organic cation, but the type of organic cation affects the band position (Table 1). The most redshifted PL is observed for the EA and MHy samples, with maxima at 688 nm and 686 nm, respectively. On the other hand, the emissions of the IM and FA compounds are positioned at 646 nm and 656 nm, respectively. PL of the DMA and Pyr analogs is located in the intermediate range, i.e., at 670 nm and 675 nm, respectively. It is worth mentioning that the emission of Mn-based dicyanamide perovskites was observed at lower wavelengths of about 626–629 nm, which is a result of using a shorter H2POO ligand instead of the long dicyanamide linkers [17,40]. The presented redshift of the emission is a natural consequence of the increasing crystal field strength acting on Mn2+ ions. The analysis of the manganese octahedron deviation (δ) from the ideal shape and position of the PL band shows a certain relationship between these parameters for the FA, EA, and MHy compounds (Table 1). In particular, if the Mn2+ ions are located in an almost perfect octahedron, and the strong crystal field causes a large splitting of energy levels and a significant redshift of the emission. [DMA]Mn(H2POO)3 deviates from this trend because the generated PL is a superposition of three emission bands belonging to Mn2+ ions located at three non-equivalent sites in the crystal structure. This trend is also observed for the IM and Pyr analogs. The considerable deviation (δ, 1.35%) of manganese octahedra in [IM]Mn(H2POO)3 shifts the emission band to a lower wavelength (λem = 646 nm), whereas almost perfect octahedra in [Pyr]Mn(H2POO)3 (δ is only 0.18%) lead to a redshifted emission at λem = 675 nm. Regardless of the observed changes in the position of the band maxima, all samples exhibit red emission at 80 K. The chromatic coordinates of the investigated compounds are demonstrated in Figure 4c.
Table 1. List of the most important parameters of [A]Mn(H2POO)3 perovskites, such as ionic radius (IR) of the ammonium cations [41], tolerance factor (tF), energy band gap (Eg) [36,37,38], the position of the PL band at 80 K (λmax), octahedral deviation (δ) from the perfect shape [36,37,38], and quenching temperature (T0.5).
Table 1. List of the most important parameters of [A]Mn(H2POO)3 perovskites, such as ionic radius (IR) of the ammonium cations [41], tolerance factor (tF), energy band gap (Eg) [36,37,38], the position of the PL band at 80 K (λmax), octahedral deviation (δ) from the perfect shape [36,37,38], and quenching temperature (T0.5).
DMA+EA+MHy+FA+IM+Pyr+
IR (pm)272344264253320258
tF0.900.9050.8910.8580.8690.90
Eg (eV)4.985.175.205.325.145.18
λmax (nm)670688686656646675
δ (%)0.170.390.4410.61.350.18
T0.5 (K)100130122135135115
The broad emission localized in the near-infrared region can be attractive for plant growth. The absorbed photoenergy can be used by plants in the photosynthesis process. The main photopigments found in plants are chlorophyll A, chlorophyll B, phytochrome PR, and phytochrome PFR [42]. The first two types of photopigments mainly absorb blue light (400–500 nm), whereas PR and PFR are sensitive to red and far-infrared light, respectively [42]. As can be clearly seen in Figure 5, the emission generated from Mn-based hybrid hypophosphite perovskite covers almost the entire absorption range of the PR and PFR phytochromes. Nevertheless, it appears that [DMA]Mn(H2POO)3 emission coincides to the greatest extent with the absorption of both types of phytochrome. The biggest disadvantage is the low thermal stability of [A]Mn(H2POO)3 sample emissions (Figure 6 and Figure 7). At the moment, it seems impossible to construct a device that would keep the hypophosphite crystal at the temperature of 80 K.
The temperature-dependent PL spectra of the investigated samples recorded from 80 K to 320 K every 10 K are presented in Figure 6. Almost all hypophosphites exhibit the highest emission intensity at 80 K. However, the thermal stability of [IM]Mn(H2POO)3 and [EA]Mn(H2POO)3 emissions is quite different. At the beginning, the intensity of Mn2+ emission decreases and then increases. This abnormal thermal behavior can be explained by energy transfer between the Mn2+ ions located at two independent optical sites [36]. A common feature of all investigated perovskites is the asymmetric shape of the PL band, moving the band position to a higher wavelength with sample heating (Tab 7a). This behavior can be associated with the Mn2+ ions occupying more than one optical site and small changes in the crystal field strength with increasing temperature [23,36,43,44]. A careful analysis of the obtained results shows that the emission of the [DMA]Mn(H2POO)3 analog quenches much faster than the remaining samples. The quenching temperature, defined as the temperature at which intensity drops to half the initial value (T0.5), is 100 K for the DMA compound (Figure 7b). [Pyr]Mn(H2POO)3 and [MHy]Mn(H2POO)3 exhibit significantly better thermal stability, with T0.5 values of 115 K and 122 K, respectively. However, the emission of Mn2+ ions in the samples comprising EA+, FA+, and IM+ has the highest thermal stability. The emission of manganese ions is quenched by the crossing of the 4T1g(G) excited state with the parabola of the 6A1(S) ground level (Figure 7c). Nevertheless, temperature-dependent measurements proved that Mn2+ ion emission in the hypophosphite perovskites quenched faster than in the case of the Mn-based dicyanamides.
The presented results explain the influence of some key structural factors on the optical properties of the family of Mn-based hybrid hypophosphites. Figure 8 summarizes the correlation between structural parameters, such as tF, octahedral deviation (δ), the number of Mn2+ sites, and optical features. As can be seen, the compounds comprising linear cations should be analyzed separately from those with cyclic organic cations. Furthermore, there is no doubt that there is a correlation between the Goldschmidt tolerance factor and the energy band gap value. The higher the tF value, the lower the Eg, so the optical band gap decreases for the samples with an almost ideal perovskite structure. For the IM and Pyr analogs, the tF and Eg values are comparable; therefore, it is impossible to define any rules. Analysis of the results allowed for the correlation of the octahedral deviation with the position of the PL band. The shift of the PL band is greater for manganese ions occupying the almost perfect octahedron. The results reported for [DMA]Mn(H2POO)3 do not comply with this rule due to the three independent optical sites of Mn2+ ions.

3. Conclusions

This paper presents the results of an investigation on the effect of organic cations on the PL properties of Mn2+ ions. Mn-based hybrid hypophosphites were synthesized by crystallization from solution. Five of the six obtained compounds crystallize in the monoclinic system in the P21/n space group. The only exception is [MHy]Mn(H2POO)3, which adopts an orthorhombic structure (Pnma space group). The obtained results indicate that the type of ammonium cation has a significant impact on the Goldschmidt tolerance factor value and the octahedral deformation parameter. As a consequence, the size of the energy band gap and the position of the emission band are tuned. It was found that the samples with a higher tolerance factor have a smaller energy band gap. Moreover, the reported redshift of the PL band position of the investigated compounds is associated with a smaller octahedral deformation. However, the temperature-dependent measurements prove that Mn2+ emission in the hypophosphite perovskites quenches faster than the PL of the Mn-based dicyanamides. The broad red PL of Mn2+ ions assigned to the 4T1g(G) → 6A1g(S) transition overlaps with the absorption of the phytochrome PR and the phytochrome PFR. Therefore, the investigated hypophosphites may be applied in an indoor lighting system for plant growth. However, a considerably disadvantage of these compounds is the low thermal stability of their PL.

Supplementary Materials

The following supporting information can be downloaded at: www.mdpi.com/article/10.3390/molecules27248953/s1. Description of sample synthesis and measurement technique.

Funding

This work was supported by the Polish National Science Centre under Grant no. 2018/31/B/ST5/00455 as part of the OPUS research project.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would also like to thank M. Mączka for samples preparation and B. Macalik for the absorption measurements.

Conflicts of Interest

The author declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the author.

References

  1. Zhang, L.; Zhang, X.; Yu, Y.; Xu, X.; Tang, J.; He, X.; Wu, J.; Lan, Z. Efficient Planar Perovskite Solar Cells Based on High-Quality Perovskite Films with Smooth Surface and Large Crystal Grains Fabricated in Ambient Air Conditions. Sol. Energy 2017, 155, 942–950. [Google Scholar] [CrossRef]
  2. Ptak, M.; Sieradzki, A.; Šimėnas, M.; Maczka, M. Molecular Spectroscopy of Hybrid Organic–Inorganic Perovskites and Related Compounds. Coord. Chem. Rev. 2021, 448, 214180. [Google Scholar] [CrossRef]
  3. Ptak, M.; Gągor, A.; Sieradzki, A.; Bondzior, B.; Dereń, P.; Ciupa, A.; Trzebiatowska, M.; Mączka, M. The Effect of K+ Cations on the Phase Transitions, and Structural, Dielectric and Luminescence Properties of [Cat][K0.5Cr0.5(HCOO)3], Where Cat Is Protonated Dimethylamine or Ethylamine. Phys. Chem. Chem. Phys. 2017, 19, 12156–12166. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Mączka, M.; Gągor, A.; Zaręba, J.K.; Stefańska, D.; Drozd, M.; Balciunas, S.; Šimenas, M.; Banys, J.; Sieradzki, A. Three-Dimensional Perovskite Methylhydrazinium Lead Chloride with Two Polar Phases and Unusual Second-Harmonic Generation Bistability above Room Temperature. Chem. Mater. 2020, 32, 4072–4082. [Google Scholar] [CrossRef]
  5. Gao, H.Q.; Wei, W.J.; Tan, Y.H.; Tang, Y.Z. Phase Transition and Negative Thermal Expansion in Guanidinium Magnesium-Hypophosphite Hybrid Perovskite. Chem. Mater. 2020, 32, 6886–6891. [Google Scholar] [CrossRef]
  6. Saparov, B.; Mitzi, D.B. Organic-Inorganic Perovskites: Structural Versatility for Functional Materials Design. Chem. Rev. 2016, 116, 4558–4596. [Google Scholar] [CrossRef]
  7. Trzebiatowska, M.; Mączka, M.; Ptak, M.; Giriunas, L.; Balciunas, S.; Simenas, M.; Klose, D.; Banys, J. Spectroscopic Study of Structural Phase Transition and Dynamic Effects in a [(CH3)2NH2][Cd(N3)3] Hybrid Perovskite Framework. J. Phys. Chem. C 2019, 123, 11840–11849. [Google Scholar] [CrossRef]
  8. Lee, S.; Park, J.H.; Lee, B.R.; Jung, E.D.; Yu, J.C.; Di Nuzzo, D.; Friend, R.H.; Song, M.H. Amine-Based Passivating Materials for Enhanced Optical Properties and Performance of Organic-Inorganic Perovskites in Light-Emitting Diodes. J. Phys. Chem. Lett. 2017, 8, 1784–1792. [Google Scholar] [CrossRef]
  9. Bermúdez-García, J.M.; Sánchez-Andújar, M.; Señarís-Rodríguez, M.A. A New Playground for Organic-Inorganic Hybrids: Barocaloric Materials for Pressure-Induced Solid-State Cooling. J. Phys. Chem. Lett. 2017, 8, 4419–4423. [Google Scholar] [CrossRef]
  10. Xu, W.J.; Du, Z.Y.; Zhang, W.X.; Chen, X.M. Structural Phase Transitions in Perovskite Compounds Based on Diatomic or Multiatomic Bridges. CrystEngComm 2016, 18, 7915–7928. [Google Scholar] [CrossRef]
  11. Boström, H.L.B.; Hill, J.A.; Goodwin, A.L. Columnar Shifts as Symmetry-Breaking Degrees of Freedom in Molecular Perovskites. Phys. Chem. Chem. Phys. 2016, 18, 31881–31894. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Chang, Q.; Zhou, X.; Jiang, S.; Xiang, G.; Li, L.; Li, Y.; Jing, C.; Ling, F.; Wang, Y.; Xiao, P. Dual-Mode Luminescence Temperature Sensing Performance of Manganese (II) Doped CsPbCl3 Perovskite Quantum Dots. Ceram. Int. 2022, 48, 33645–33652. [Google Scholar] [CrossRef]
  13. Wang, M.; Zang, Z.; Yang, B.; Hu, X.; Sun, K.; Sun, L. Performance Improvement of Perovskite Solar Cells through Enhanced Hole Extraction: The Role of Iodide Concentration Gradient. Sol. Energy Mater. Sol. Cells 2018, 185, 117–123. [Google Scholar] [CrossRef]
  14. Ptak, M.; Mączka, M. Phonon Properties and Mechanism of Order-Disorder Phase Transition in Formamidinium Manganese Hypophosphite Single Crystal. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2020, 230, 118010. [Google Scholar] [CrossRef]
  15. Li, M.; Smetana, V.; Wilk-Kozubek, M.; Mudryk, Y.; Alammar, T.; Pecharsky, V.K.; Mudring, A.V. Open-Framework Manganese(II) and Cobalt(II) Borophosphates with Helical Chains: Structures, Magnetic, and Luminescent Properties. Inorg. Chem. 2017, 56, 11104–11112. [Google Scholar] [CrossRef]
  16. Bermúdez-García, J.M.; Yáñez-Vilar, S.; García-Fernández, A.; Sánchez-Andújar, M.; Castro-García, S.; López-Beceiro, J.; Artiaga, R.; Dilshad, M.; Moya, X.; Señarís-Rodríguez, M.A. Giant Barocaloric Tunability in [(CH3CH2CH2)4N]Cd[N(CN)2]3 Hybrid Perovskite. J. Mater. Chem. C 2018, 6, 9867–9874. [Google Scholar] [CrossRef] [Green Version]
  17. Mączka, M.; Gągor, A.; Ptak, M.; Stefańska, D.; Macalik, L.; Pikul, A.; Sieradzki, A. Structural, Phonon, Magnetic and Optical Properties of Novel Perovskite-like Frameworks of TriBuMe[M(Dca)3] (TriBuMe = Tributylmethylammonium; Dca = Dicyanamide; M = Mn2+, Fe2+, Co2+, Ni2+). Dalt. Trans. 2019, 48, 13006–13016. [Google Scholar] [CrossRef]
  18. Trzebiatowska, M.; Gągor, A.; Macalik, L.; Peksa, P.; Sieradzki, A. Phase Transition in the Extreme: A Cubic-to-Triclinic Symmetry Change in Dielectrically Switchable Cyanide Perovskites. Dalt. Trans. 2019, 48, 15830–15840. [Google Scholar] [CrossRef]
  19. Xu, W.J.; Li, P.F.; Tang, Y.Y.; Zhang, W.X.; Xiong, R.G.; Chen, X.M. A Molecular Perovskite with Switchable Coordination Bonds for High-Temperature Multiaxial Ferroelectrics. J. Am. Chem. Soc. 2017, 139, 6369–6375. [Google Scholar] [CrossRef]
  20. Lu, M.; Zhang, X.; Bai, X.; Wu, H.; Shen, X.; Zhang, Y.; Zhang, W.; Zheng, W.; Song, H.; Yu, W.W.; et al. Spontaneous Silver Doping and Surface Passivation of CsPbI3 Perovskite Active Layer Enable Light-Emitting Devices with an External Quantum Efficiency of 11.2%. ACS Energy Lett. 2018, 3, 1571–1577. [Google Scholar] [CrossRef]
  21. Zhao, X.; Ng, J.D.A.; Friend, R.H.; Tan, Z.K. Opportunities and Challenges in Perovskite Light-Emitting Devices. ACS Photonics 2018, 5, 3866–3875. [Google Scholar] [CrossRef]
  22. Zeng, X.; Zhou, T.; Leng, C.; Zang, Z.; Wang, M.; Hu, W.; Tang, X.; Lu, S.; Fang, L.; Zhou, M. Performance Improvement of Perovskite Solar Cells by Employing a CdSe Quantum Dot/PCBM Composite as an Electron Transport Layer. J. Mater. Chem. A 2017, 5, 17499–17505. [Google Scholar] [CrossRef]
  23. Kore, B.P.; Das, S.; Sarma, D.D. Temperature-Dependent Anomalous Mn2+ Emission and Excited State Dynamics in Mn2+-Doped MAPbCl3-XBrx Nanocrystals. J. Chem. Sci. 2021, 133, 64. [Google Scholar] [CrossRef]
  24. Mączka, M.; Ptak, M.; Gągor, A.; Stefańska, D.; Sieradzki, A. Layered Lead Iodide of [Methylhydrazinium]2PbI4 with a Reduced Band Gap: Thermochromic Luminescence and Switchable Dielectric Properties Triggered by Structural Phase Transitions. Chem. Mater. 2019, 31, 8563–8575. [Google Scholar] [CrossRef]
  25. Mączka, M.; Ptak, M.; Gągor, A.; Stefańska, D.; Zarȩba, J.K.; Sieradzki, A. Methylhydrazinium Lead Bromide: Noncentrosymmetric Three-Dimensional Perovskite with Exceptionally Large Framework Distortion and Green Photoluminescence. Chem. Mater. 2020, 32, 1667–1673. [Google Scholar] [CrossRef]
  26. Snaith, H.J. Perovskites: The Emergence of a New Era for Low-Cost, High-Efficiency Solar Cells. J. Phys. Chem. Lett. 2013, 4, 3623–3630. [Google Scholar] [CrossRef]
  27. Grätzel, M. The Light and Shade of Perovskite Solar Cells. Nat. Mater. 2014, 13, 838–842. [Google Scholar] [CrossRef]
  28. Jeong, J.; Kim, M.; Seo, J.; Lu, H.; Ahlawat, P.; Mishra, A.; Yang, Y.; Hope, M.A.; Eickemeyer, F.T.; Kim, M.; et al. Pseudo-Halide Anion Engineering for α-FAPbI3 Perovskite Solar Cells. Nature 2020, 592, 381–385. [Google Scholar] [CrossRef]
  29. Yoo, J.J.; Seo, G.; Chua, M.R.; Park, T.G.; Lu, Y.; Rotermund, F.; Kim, Y.K.; Moon, C.S.; Jeon, N.J.; Correa-Baena, J.P.; et al. Efficient Perovskite Solar Cells via Improved Carrier Management. Nature 2021, 590, 587–593. [Google Scholar] [CrossRef]
  30. Drozdowski, D.; Gągor, A.; Stefańska, D.; Zarȩba, J.K.; Fedoruk, K.; Mączka, M.; Sieradzki, A. Three-Dimensional Methylhydrazinium Lead Halide Perovskites: Structural Changes and Effects on Dielectric, Linear, and Nonlinear Optical Properties Entailed by the Halide Tuning. J. Phys. Chem. C 2022, 126, 1600–1610. [Google Scholar] [CrossRef]
  31. Kieslich, G.; Sun, S.; Cheetham, A.K. An Extended Tolerance Factor Approach for Organic-Inorganic Perovskites. Chem. Sci. 2015, 6, 3430–3433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  32. Kieslich, G.; Sun, S.; Cheetham, A.K. Solid-State Principles Applied to Organic–Inorganic Perovskites: New Tricks for an Old Dog. Chem. Sci. 2014, 5, 4712–4715. [Google Scholar] [CrossRef]
  33. Boström, H.L.B. Tilts and Shifts in Molecular Perovskites. CrystEngComm 2020, 22, 961–968. [Google Scholar] [CrossRef] [Green Version]
  34. Wu, Y.; Shaker, S.; Brivio, F.; Murugavel, R.; Bristowe, P.D.; Cheetham, A.K. Mn(H2POO)3: A New Family of Hybrid Perovskites Based on the Hypophosphite Ligand. J. Am. Chem. Soc. 2017, 139, 16999–17002. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Mączka, M.; Gągor, A.; Stefańska, D.; Zaręba, J.K.; Pikul, A. Structural, Magnetic and Photoluminescence Properties of New Hybrid Hypophosphites: Discovery of the First Noncentrosymmetric and Two Cobalt-Based Members. Dalt. Trans. 2022, 51, 9094–9102. [Google Scholar] [CrossRef]
  36. Mączka, M.; Stefańska, D.; Gągor, A.; Pikul, A. The Cation-Dependent Structural, Magnetic and Optical Properties of a Family of Hypophosphite Hybrid Perovskites. Dalt. Trans. 2022, 51, 352–360. [Google Scholar] [CrossRef]
  37. Mączka, M.; Gągor, A.; Pikul, A.; Stefańska, D. Novel Hypophosphite Hybrid Perovskites of [CH3NH2NH2][Mn(H2POO)3] and [CH3NH2NH2][Mn(H2POO)2.83(HCOO)0.17] Exhibiting Antiferromagnetic Order and Red Photoluminescence. RSC Adv. 2020, 10, 19020–19026. [Google Scholar] [CrossRef]
  38. Mączka, M.; Stefańska, D.; Ptak, M.; Gągor, A.; Pikul, A.; Sieradzki, A. Cadmium and Manganese Hypophosphite Perovskites Templated by Formamidinium Cations: Dielectric, Optical and Magnetic Properties. Dalt. Trans. 2021, 50, 2639–2647. [Google Scholar] [CrossRef]
  39. Xue, J.; Li, F.; Liu, F.; Noh, H.M.; Lee, B.R.; Choi, B.C.; Park, S.H.; Jeong, J.H.; Du, P. Designing Ultra-Highly Efficient Mn2+-Activated Zn2GeO4 Green-Emitting Persistent Phosphors toward Versatile Applications. Mater. Today Chem. 2022, 23, 100693. [Google Scholar] [CrossRef]
  40. Mączka, M.; Gągor, A.; Ptak, M.; Stefańska, D.; Sieradzki, A. Temperature-Dependent Studies of a New Two-Dimensional Cadmium Dicyanamide Framework Exhibiting an Unusual Temperature-Induced Irreversible Phase Transition into a Three-Dimensional Perovskite-like Framework. Phys. Chem. Chem. Phys. 2018, 20, 29951–29958. [Google Scholar] [CrossRef]
  41. Shannon, R.D.; Prewitt, C.T. Effective Ionic Radii in Oxides and Fluorides. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1969, 25, 925–946. [Google Scholar] [CrossRef]
  42. Dai, H.; Li, S.; Li, Z.; Li, J.; Xin, S.; Wang, C.; Zhu, G.; Dong, B. Novel Rare Earth Free Phosphors CsMg2P3O10: Mn2+ with Efficient and Ultra-Broadband Red Emission for Plant Growth LEDs. J. Am. Ceram. Soc. 2022, 105, 4719–4730. [Google Scholar] [CrossRef]
  43. Song, E.; Chen, M.; Chen, Z.; Zhou, Y.; Zhou, W.; Sun, H.T.; Yang, X.; Gan, J.; Ye, S.; Zhang, Q. Mn2+-Activated Dual-Wavelength Emitting Materials toward Wearable Optical Fibre Temperature Sensor. Nat. Commun. 2022, 13, 2166. [Google Scholar] [CrossRef]
  44. Bermúdez-García, J.M.; Sánchez-Andújar, M.; Castro-García, S.; López-Beceiro, J.; Artiaga, R.; Señarís-Rodríguez, M.A. Giant Barocaloric Effect in the Ferroic Organic-Inorganic Hybrid [TPrA][Mn(Dca)3] Perovskite under Easily Accessible Pressures. Nat. Commun. 2017, 8, 15715. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Chemical structures of organic cations used for the synthesis of [A]Mn(H2POO)3 hypophosphite perovskites.
Figure 1. Chemical structures of organic cations used for the synthesis of [A]Mn(H2POO)3 hypophosphite perovskites.
Molecules 27 08953 g001
Figure 2. Crystal structures of the investigated manganese hypophosphites.
Figure 2. Crystal structures of the investigated manganese hypophosphites.
Molecules 27 08953 g002
Figure 3. (a) Diffuse absorption spectra of the investigated samples and (b) scheme of the energy band gap estimated by the Kubelka–Munk function.
Figure 3. (a) Diffuse absorption spectra of the investigated samples and (b) scheme of the energy band gap estimated by the Kubelka–Munk function.
Molecules 27 08953 g003
Figure 4. (a) Low-temperature PL spectra of [A]Mn(H2POO)3; (b) energy-level diagram of Mn2+ ions; (c) CIE coordinates of the investigated samples.
Figure 4. (a) Low-temperature PL spectra of [A]Mn(H2POO)3; (b) energy-level diagram of Mn2+ ions; (c) CIE coordinates of the investigated samples.
Molecules 27 08953 g004
Figure 5. Collation of PL spectra of the investigated manganese hypophosphite and absorption of the PR and PFR phytochromes.
Figure 5. Collation of PL spectra of the investigated manganese hypophosphite and absorption of the PR and PFR phytochromes.
Molecules 27 08953 g005
Figure 6. (af) Temperature-dependent emission spectra of the investigated manganese hypophosphite perovskites.
Figure 6. (af) Temperature-dependent emission spectra of the investigated manganese hypophosphite perovskites.
Molecules 27 08953 g006
Figure 7. (a) Position of the PL band as a function of temperature up to 220 K; (b) integrated emission intensity of [A]Mn(H2POO)3 as a function of temperature; and (c) energy-level diagram of Mn2+ presenting the mechanism of PL quenching.
Figure 7. (a) Position of the PL band as a function of temperature up to 220 K; (b) integrated emission intensity of [A]Mn(H2POO)3 as a function of temperature; and (c) energy-level diagram of Mn2+ presenting the mechanism of PL quenching.
Molecules 27 08953 g007
Figure 8. Influence of the amine applied in [A]Mn(H2POO)3 compounds on structural and optical features. The scheme shows the correlation between the structural parameters (such as tF, octahedral deviation (δ) from the perfect cubic shape, and the number of Mn2+ sites) and the optical properties (energy band gap and PL position).
Figure 8. Influence of the amine applied in [A]Mn(H2POO)3 compounds on structural and optical features. The scheme shows the correlation between the structural parameters (such as tF, octahedral deviation (δ) from the perfect cubic shape, and the number of Mn2+ sites) and the optical properties (energy band gap and PL position).
Molecules 27 08953 g008
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Stefańska, D. Effect of Organic Cation on Optical Properties of [A]Mn(H2POO)3 Hybrid Perovskites. Molecules 2022, 27, 8953. https://doi.org/10.3390/molecules27248953

AMA Style

Stefańska D. Effect of Organic Cation on Optical Properties of [A]Mn(H2POO)3 Hybrid Perovskites. Molecules. 2022; 27(24):8953. https://doi.org/10.3390/molecules27248953

Chicago/Turabian Style

Stefańska, Dagmara. 2022. "Effect of Organic Cation on Optical Properties of [A]Mn(H2POO)3 Hybrid Perovskites" Molecules 27, no. 24: 8953. https://doi.org/10.3390/molecules27248953

Article Metrics

Back to TopTop