Next Article in Journal
Proteome Analysis of Human Natural Killer Cell Derived Extracellular Vesicles for Identification of Anticancer Effectors
Next Article in Special Issue
Diffuse Reflectance Spectroscopy with Dilution: A Powerful Method for Halide Perovskites Study
Previous Article in Journal
Towards Volatile Organoselenium Compounds with Cost-Effective Synthesis
Previous Article in Special Issue
A Review on Lead-Free Hybrid Halide Perovskites as Light Absorbers for Photovoltaic Applications Based on Their Structural, Optical, and Morphological Properties
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Vibrational Properties and DFT Calculations of Perovskite-Type Methylhydrazinium Manganese Hypophosphite

1
Institute of Low Temperature and Structure Research, Polish Academy of Sciences, 50-422 Wrocław, Poland
2
Department of Bioorganic Chemistry, Faculty of Production Engineering, University of Economics and Business, 118/120 Komandorska str., 53-345 Wrocław, Poland
*
Author to whom correspondence should be addressed.
Submission received: 11 September 2020 / Revised: 23 October 2020 / Accepted: 5 November 2020 / Published: 9 November 2020

Abstract

:
Recently discovered hybrid perovskites based on hypophosphite ligands are a promising class of compounds exhibiting unusual structural properties and providing opportunities for construction of novel functional materials. Here, we report for the first time the detailed studies of phonon properties of manganese hypophosphite templated with methylhydrazinium cations ([CH3NH2NH2][Mn(H2PO2)3]). Its room temperature vibrational spectra were recorded for both polycrystalline sample and a single crystal. The proposed assignment based on Density Functional Theory (DFT) calculations of the observed vibrational modes is also presented. It is worth noting this is first report on polarized Raman measurements in this class of hybrid perovskites.

Graphical Abstract

1. Introduction

In recent years, a lot of attention has been devoted to hybrid organic-inorganic compounds of general formula ABX3 (A = organic cation; B = divalent metal cation; X = halide, formate (HCOO), azide (N3), cyanide (CN), dicyanamide (N(CN)2, dca), hypophosphite (H2PO2)), crystallizing in a perovskite-type architecture due to their interesting physicochemical properties including photovoltaic [1], light-emitting [2,3,4,5], multiferroic [6,7,8], barocaloric [9,10], and non-linear optical (NLO) [11,12,13] properties. The structure-related properties of these compounds depend on flexibility of ligands, size and shape of organic cations as well as their ability to form hydrogen bonds (HBs) with the anionic framework. In this respect, very large flexibility of the dca- ligands and pronounced temperature-induced disorder of organic cations at elevated temperatures is responsible for unusual softness and promising barocaloric properties of dicyanamides [9,10,14,15]. Order-disorder phenomena are also responsible for onset of spontaneous polarization in formates and their multiferroic properties [6,8]. These phenomena can be tuned by proper use of organic cations and we showed for the first time in 2017 that methylhydrazinium cation (CH3NH2NH2+, MHy+) may also be used for the construction of perovskite-type formates [8]. Very recently, we showed that MHy+ also allows extending the formation of three-dimensional (3D) lead halides beyond two already known methylammonium and formamidinium analogues [12,13]. This cation also allows constructing the two-dimensional (2D) perovskite of the formula MHy2PbI4 [5] as well as 3D hypophosphite [MHy][Mn(H2PO2)3] [5].
While perovskite-type halides, formates, azides, cyanides, and dicyanamides have been known for a long time, their hypophosphite analogues were discovered for the first time in 2017 [16]. Except for one magnesium analogue [17], all known compounds contain manganese as divalent cation [5,16,18,19]. Literature data show that although H2POO ligand has a very similar size to the formate ion, the hypophosphite frameworks show much higher flexibility than their formate counterparts [16,19]. As a result, hypophosphites have higher potential than formates for structural distortion, which results in the presence of unconventional tilts of MnO6 octahedra and columnar shifts [16,19]. Furthermore, they often show pronounced off-centering of organic cations away from the cavity center [16,19].
Up until now, Raman and IR studies were reported for only one hypophosphite, i.e., formamidinium manganese hypophosphite [20]. Regarding methylhydrazinium compounds, we reported Raman and IR data for some formates and two halides (MHy2PbI4 and MHyPbBr3). However, deep understanding of vibrational modes of MHy+ cations is still lacking and vibrational properties of recently discovered [MHy][Mn(H2PO2)3] are not known. We decided, therefore, to analyze vibrational properties of [MHy][Mn(H2PO2)3] crystal and propose reliable assignment of vibrational bands by performing Density Functional Theory (DFT) calculations.

2. Experimental Section

2.1. Synthesis of Single Crystals

[MHy][Mn(H2POO)3] crystals were grown by slow evaporation at 50°C of solution prepared by dissolving 5 mmol of MnCO3 (99.9%, Sigma-Aldrich, Saint Louis, MO, USA) and 10 mmol (0.53 mL) of methylhydrazine (98%, Sigma-Aldrich) in 6.46 mL of aqueous solution of hypophosphorous acid (50%, Sigma-Aldrich) (equivalent of 60 mmol of H3PO2). Further details can be found in our previous paper [5].

2.2. Materials and Methods

Polycrystalline IR spectra were measured using a Nicolet iS50 FT-IR spectrometer. The standard KBr pellet method was used for the mid-IR range (4000–400 cm−1) and nujol mull on a polyethylene plate for the far-IR range (600–100 cm−1). Raman spectrum of powdered crystals was measured using a Bruker FT100/S spectrometer (Billerica, MA, USA) with YAG:Nd laser excitation (1064 nm). Polarized and temperature-dependent Raman spectra were measured using a Renishaw InVia Raman spectrometer (Wotton-under-Edge, UK), equipped with confocal DM2500 Leica optical microscope, a thermoelectrically cooled CCD as a detector, and an Ar+ ion laser operating at 488 nm. Spectral resolution of both Raman and IR spectra was set to be 2 cm-1 and temperature was controlled using Linkam THMS600 cryostat cell (Tadworth, UK), equipped with quartz windows.

2.3. Computational Details

The geometry optimization of the molecular structure of the MHy+ cation (Figure 1) was performed with the use of Gaussian 03 program package [21]. All calculations were performed using density functional three-parameters hybrid (B3LYP) methods [22,23,24] with the 6-311G(2d,2p) [25,26] basis set. In the geometry optimization, the structural parameters from X-ray diffraction studies were taken as input data. The comparison of the calculated and experimental bond lengths and bond angles is presented in Table S1. Geometry optimization was followed by the calculations of vibrational frequencies. The calculated harmonic frequencies were scaled using scaling factors (0.96 and 0.98) to correct the evaluated wavenumbers for vibrational anharmonicity and deficiencies inherent to the used computational level. Anharmonic wavenumbers were also calculated using deperturbed vibrational second-order perturbation theory (DVPT2). To assign the calculated modes, the Chemcraft computer program was used [27]. The Potential Energy Distribution (PED) of the normal modes among the respective internal coordinates was calculated for studied compound using the BALGA [28] program.

3. Results and Discussion

3.1. Structure

As follows from Table S1, the calculated C-N and N-N bond lengths (1.506 and 1.447Å, respectively) and C-N-N angle (118.3°) are in very good agreement with the experimental values (1.470, 1.442 Å and 116.3°, respectively [5]). The calculated C-H and N-H bonds are significantly longer than the experimental ones but it should be remembered that X-ray diffraction method is not sensitive for light atoms like hydrogen. The majority of the remaining calculated angles are also in agreement with the experimental ones.

3.2. Vibrational Properties

Experimental IR and Raman spectra of powdered crystals are presented in Figure 2, while polarized Raman spectra are depicted in Figure 3. Table 1 and Table S2 list the calculated and observed experimental wavenumbers as well as the proposed assignment.
Internal vibrations of MHy+ cation. Free MHy+ cation has 24 vibrational degrees of freedom that can be subdivided into symmetric stretching (2νsNH2), antisymmetric stretching (2νasNH2), scissoring (2δNH2), rocking (2ρNH2), wagging (2ωNH2), and torsion or twisting (2τNH2) modes of the NH2 and NH2+ groups; symmetric stretching (νsCNN), antisymmetric stretching (νasCNN) and bending (δCNN) modes of the CNN skeletal group, and symmetric stretching (νsCH3), antisymmetric stretching (2νasCH3), symmetric bending (δsCH3), antisymmetric bending (2δasCH3), rocking (2ρCH3) and torsion (τCH3) modes of the methyl group.
DFT calculations show that the N-H stretching modes of the terminal NH2 (middle NH2+) groups are expected at 3388 + 3306 cm−1 (3292 + 3289 cm−1) (Table 1). The former modes are observed in the experimental spectra as relatively narrow bands shifted to lower wavenumbers by 30–107 cm−1, compared to the calculated ones (Table 1 and Table S2). Small bandwidths and shifts are consistent with involvement of these NH2 groups in formation of weak HBs (according to the X-ray diffraction, the NO distance is 3.001 Å [5]). It is worth noting that the 3281 cm−1 band is significantly weaker in the Raman spectrum than that observed at 3176 cm−1, but opposite behavior is observed in the IR spectrum (Figure 2). Therefore, we assign the 3281 and 3176 cm−1 bands to antisymmetric and symmetric stretching, respectively. The modes calculated at 3292 + 3289 cm−1 are observed as broad and medium intensity IR bands at 3118 + 3085 cm−1 (Figure 2, Table S2). This behavior confirms involvement of the NH2+ groups in formation of medium strong HBs (the shortest NO distance is 2.875 Å [5]).
The C-H stretching modes of the methyl group are observed as the strongest Raman bands in the 3300–2700 cm−1 range. These bands are narrow and those at 3028 and 3021 cm−1 have significantly smaller intensity than that at 2958 cm−1, allowing their assignment to antisymmetric and symmetric modes, respectively. Raman and IR spectra show also a few bands in the 2930–2700 cm−1 range that are not predicted by DFT calculations (Table 1 and Table S2). These bands correspond, therefore, to some combination modes.
The calculated scissoring modes of the NH2 and NH2+ groups (1679 and 1607 cm−1, respectively) are in good agreement with the observed ones (1632 and 1574 cm−1, Table S2). Good agreement is also observed for the C-H bending modes that are predicted (observed) in the 1472–1439 cm−1 (1474–1428 cm−1) range.
According to the DFT calculations, τNH2+, ωNH2, ωNH2+, ρNH2 and ρCH3 modes should be observed in the 1449–1031 cm−1 range. The mode calculated at 1031 cm−1 cannot be observed in our spectra due to overlapping with strong bands of the hypophosphite ion. The remaining modes are, however, observed in the 1403–1136 cm−1 range (Table S2).
In the 1020–770 cm−1 range, three bands can be identified as arising from vibrations of MHy+ cation. The most intense Raman band at 878 cm−1 can be unambiguously assigned to the νsCNN vibration calculated at 811 cm−1. DFT predicts that the ρNH2 mode should be observed at slightly higher wavenumber (821 cm−1) than the νsCNN one. We locate, therefore, this mode in the 915–910 cm−1 range. The band observed at 1010 cm−1 (calculated at 907 cm−1) can be assigned to the νasCNN mode.
DFT calculations predict the presence of three internal modes of MHy+ in the region below 500 cm−1. We could identify two of these modes around 440 and 210 cm−1 and assign them to δCNN and τCH3, respectively (Table 1 and Table S2).
Unit cell of [MHy][Mn(H2PO2)3] contains four MHy+ cations located at sites of Cs symmetry. Therefore, all internal modes of this cation should split into four components (see Table S3). Raman spectra clearly show that many Ag modes are observed at slightly different wavenumbers than their B1g, B2g or B3g counterparts, i.e., the significant Davydov splitting is observed (Table S2). The largest differences are observed for the δNH2+ mode (1574 and 1568 cm−1 for the Ag and B2g mode, respectively) and δasCH3 mode (1451/1452 cm−1 and 1457 cm−1 for the Ag and B1g mode, respectively). In other cases, the differences between wavenumbers of Raman components are less than 1–3 cm−1. For instance, the symmetric stretching mode of the CH3 group (νsCH3) is visible at 2959 cm−1 for Ag and B2g modes and at 2958 cm−1 for B1g and B3g modes. Similar slight changes in the positions of Davydov components are observed for its antisymmetric counterpart (νasCH3), at 2959 cm−1 for Ag and B2g modes and at 2958 cm−1 for B1g and B3g modes. It is worth mentioning that the size of Davydov splitting depends on the strength of intermolecular interactions occurring in the studied compound, namely, stronger interactions lead to larger Davydov splitting.
It is worth making a short comparison of the MHy+ modes observed in this hypophosphite and related formate ([MHy][Mn(HCOO)3]) reported in the literature. According to the X-ray diffraction data, MHy+ cations are ordered in the hypophosphite and the terminal NH2 groups are involved in weak HBs, whereas in the formate these cations are disordered and the terminal groups form no HBs [5,8]. Furthermore, MHy+ cation shifts away from the cavity center in the hypophosphite but sits in the center in the formate. This difference is related to more flexible P-O-Mn bindings compared to the C-O-Mn ones [5]. Raman and IR spectra reflect these differences very clearly. Firstly, stretching modes of the terminal NH2 groups are shifted to higher wavenumbers for the formate (3317 and 3195 cm−1 [8]) due to lack of HBs interactions. The skeletal modes are, however, observed at very similar wavenumbers for the both analogues (the largest difference is only 3 cm−1), reflecting similar geometry of MHy+ cations. Secondly, stretching modes of the methyl groups in the formate are shifted up to 9 cm−1 due to larger constraint imposed on the MHy+ cations. Thirdly, IR band corresponding to the ρNH2+ mode is very broad for the formate (about 30 cm−1 at RT) compared to the hypophosphite (12.2 cm−1), reflecting disorder in the former case.
Internal vibrations of hypophosphite anion. The free hypophosphite ion has C2v symmetry and 9 fundamental vibrations, i.e., symmetric stretching (νs(PH2), A1), antisymmetric stretching (νas(PH2), B1), scissoring (δ(PH2), A1), rocking (ρ(PH2), B1), twisting (τ(PH2), A2) and wagging (ω(PH2), B2) modes of the PH2 groups as well as symmetric stretching (νs(PO2), A1), antisymmetric stretching (νas(PO2), B2) and scissoring (δ(PO2), A1) modes of the PO2 moiety [29].
Former experimental and theoretical studies showed that five of the mentioned above vibrations of H2PO2 anion are observed in well-separated regions. That is, νs(PH2) at 2345–2424 cm−1, νas(PH2) at 2305–2368 cm−1, τ(PH2) at 899–945 cm−1, ρ(PH2) at 792–834 cm−1 and δ(PO2) at 456–518 cm−1 [20,30,31,32]. For [MHy][Mn(H2PO2)], these modes are observed at 2333–2379, 2297–2322, 910–921, 799–824 and 474–519 cm−1, respectively (Table S2).
According to the literature data, the νas(PO2) and δ(PH2) modes are observed in a similar wavenumber range, typically at 1154–1230 and 1132–1172 cm−1, respectively [20,30,31,32]. The same behavior is observed for the νs(PO2) and ω(PH2) modes that are usually observed at 1024–1062 and 1070–1126 cm−1, respectively [30,31,32]. Our Raman and IR spectra of [MHy][Mn(H2PO2)] show two separated groups of bands at 1154–1204 and 1048–1092 cm−1. The former bands can be assigned to νas(PO2) and δ(PH2) modes and the latter to νs(PO2) and ω(PH2) ones.
It is worth adding that there are two types of H2PO2 ions in the [MHy][Mn(H2PO2)] structure, one located at site of Cs and second at site of C1 symmetry [5]. Each fundamental mode of the former (later) anion should split into 4 (8) Davydov components (see Table S3). Indeed, Raman and IR spectra clearly show splitting of each mode into a few bands (Table S2). For instance, two τ(PH2) modes are observed due to two crystallographically distinct H2PO2 ions. Wavenumbers of these modes practically do not depend on polarization, pointing to negligible Davydov splitting. However, the νs(PH2), and νas(PH2) and δ(PO2) split into three components and the ρ(PH2) into four components, indicating significant Davydov splitting for these modes.
Lattice modes. Mn2+ ions are located in the structure at inversion center and as a result, T’(Mn2+) mode, are IR-active only (see Table S3). Our IR spectrum shows two strong IR bands at 229 and 280 cm−1 that have no Raman counterparts. These bands can be unambiguously assigned to T’(Mn2+) modes. It is worth noting here that this assignment is consistent with studies of zinc hypophosphites, which revealed T’(Zn2+) modes at 234–357 cm−1 [32].
Modes in the 137-162 cm−1 range are much stronger in the IR than Raman spectra while opposite behavior can be noticed for the modes located below 120 cm−1. This behavior is consistent with assignment of the former bands to the coupled modes involving translational motions of T’(Mn2+) and T’(H2PO2) ions and the latter ones to librational L(H2PO2) modes.

3.3. Temperature Dependence of Raman Modes

Raman spectra of the [MHy][Mn(H2PO2)] sample recorded as a function of temperature (80–360 K) are presented in Figure S1. Increasing of temperature leads to broadening and shift of bands. As a consequence, many bands are overlapped and hardly resolved above the room temperature. Since splitting as well as an emergence of additional or a disappearance of already present bands is not observed, the occurrence of a structural phase transition in [MHy][Mn(H2PO2)] can be excluded, in agreement with the literature [5]. This conclusion is further supported by the temperature evolution of wavenumbers and full width at half maximum (FWHM) values for a few selected Raman modes (Figure 4 and Figure 5), which do not show any splitting or departures from the typical anharmonic behavior in the studied temperature range.
Figure 4 presents the temperature-dependent evolution of bands assigned to vibrations of PH2 and PO2 groups building H2PO2 liners. As can be seen, Raman bands exhibit weak hardening upon cooling from 360 K to 80 K, that is, the observed shifts do not exceed 6 cm−1. The largest shift concerns bands located at 823 and 830 cm−1 (360 K) attributed to rocking motions of PH2 groups (ρPH2), which upon cooling move toward lowest wavenumbers by about 6 cm−1. Much weaker changes, i.e., upshifts by 1 or 2 cm−1, can be observed for bands assigned to bending motions of PO2 groups (δPO2) located at 519 and 473 cm−1 (at 360 K). Interesting observations can also be made for temperature dependence of FWHM of bands corresponding to vibrations of PH2 and PO2 groups (Figure S5a). In this case, modes narrow significantly upon cooling. For instance, the bandwidth of 519 cm−1 (RT) Raman band (δPO2) decreases from 16.4 cm−1 (360 K) to 9.3 cm−1 (80 K). Much weaker narrowing from 4.4 cm−1 (360 K) to 3.0 cm−1 (80 K) applies to band located at 822 cm−1 (RT), which has the main contribution of the rocking vibration of PH2 groups. Differences in thermal behavior for different vibrational modes of PH2 and PO2 groups can be explained by various surrounding and their different ability to form H-bonding. Please note that two inequivalent H2PO2 units coordinate neighboring Mn2+ ions in a different way, and thus the created P-O-Mn angles vary in a broad range (127, 130 and 157°) [5]. These linkers are also involved in the formation of hydrogen bonds (HBs) with the NH2 groups from MHy+ cations, which is directly responsible for the off-center displacement of organic ions inside the cavities. Decrease of temperature leads to slowing down of thermally-activated motions followed by decrease of the unit cell volume and increase of the HBs strength. Since the impact of HBs is not the same for the PH2 and PO2 groups, different thermal behavior for modes attributed to their vibrations can be observed.
Similar behavior also applies to bands assigned to vibrational motions of MHy+ cations (Figure 4b). In this case, the observed shifts do not exceed 5 cm−1. The largest changes are observed for bands at 1142 cm−1 and 881 cm−1 (360 K). Upon cooling, the first peak corresponding to mixed rocking and torsion vibrations of the NH2 groups exhibits downshift to 1137 cm−1, whereas the second one originating from symmetric stretching vibrations of the CNN moieties changes its position to 877 cm−1. Meanwhile, minimal hardening of 0.8 cm−1 upon cooling is noticed for the band at 1237 cm−1 (RT) attributed to mixed rocking vibrations of the CH3 groups and wagging motions of the NH2 groups. As in the case of vibrations of H2PO2 linkers, observed differences in thermal evolution of modes attributed to vibrational motions of MHy+ ions can also be explained by different influence of HBs on particular groups in the structure.
The thermal behavior of FHWM of the 878 cm−1 (RT), 1237 cm−1 and 1139 cm−1 bands corresponding to the νsCNN, ρCH3+ωNH2 and ρCH3+τNH2+ modes, respectively, is presented at Figure 5b. As can be seen, upon cooling all mentioned bands exhibit significant narrowing. The larger narrowing, by almost 9 cm−1, is observed for the band attributed to mixed rocking vibration of the CH3 groups and twisting vibrations of the NH2+ groups. Observed results prove that change of temperature leads to strong change in the HB strengths.

4. Conclusions

Phonon properties of manganese hypophosphite framework templated by methylhydrazinium cations were studied using IR and Raman spectroscopies. We have presented selection rules for the Pnma orthorhombic phase together with proposed assignment of the observed IR and Raman bands to the respective internal and lattice vibrations. Presented ascription was based on our DFT calculations made for the isolated MHy+ ion in harmonic and anharmonic approximations. Polarized Raman spectra provided additional information on magnitude of Davydov splitting observed in the studied organic-inorganic hybrid. The temperature-dependent Raman studies allowed us to obtain deeper insight into structural changes occurring in the [MHy][Mn(H2POO)3] compound. Differences in the temperature evolution of wavenumbers and FWHMs of bands attributed to vibrational motions of the MHy+ as well as H2PO2 ions can be explained by the different influence of HB on individual groups that built this structure.

Supplementary Materials

The following are available online, Figure S1: The temperature-dependent Raman spectra, Table S1: Selected bond lengths and bond angles of MHy+ cation. Table S2: Raman and IR wavenumbers of [MHy][Mn(H2POO)3] at RT, Table S3: The correlation diagram and irreducible representations for the Pnma orthorhombic phase of the [MHy][Mn(H2POO)3] crystal.

Author Contributions

Conceptualization, A.C.-L., M.P. and M.M.; data curation, M.M.; formal analysis, J.H.; funding acquisition, M.M.; investigation, A.C.-L. and M.P.; methodology, M.P.; project administration, M.M.; resources, M.M.; software, E.K.; supervision, M.M.; validation, A.C.-L., M.P. and M.M.; visualization, A.C.-L.; writing—original draft, A.C.-L., M.P., E.K. and M.M.; writing—review and editing, M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by National Science Center (Narodowe Centrum Nauki) in Poland, grant number 2018/31/B/ST5/00455.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, F.; Cao, Y.; Chen, C.; Chen, Q.; Wu, X.; Li, X.; Qin, T.; Huang, W. Materials Toward the Upscaling of Perovskite Solar Cells: Progress, Challenges, and Strategies. Adv. Funct. Mater. 2018, 28, 1803753. [Google Scholar] [CrossRef]
  2. 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]
  3. 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]
  4. Mączka, M.; Stefańska, D.; Zaręba, J.K.; Nyk, M.; Sieradzki, A. Temperature-dependent luminescence and second-harmonic generation of perovskite-type manganese and cadmium dicyanamide frameworks templated by tetrapropylammonium cations. J. Alloys Compd. 2020, 821, 153464. [Google Scholar] [CrossRef]
  5. 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]
  6. Jain, P.; Ramachandran, V.; Clark, R.J.; Zhou, H.D.; Toby, B.H.; Dalal, N.S.; Kroto, H.W.; Cheetham, A.K. Multiferroic behavior associated with an order-disorder hydrogen bonding transition in metal-organic frameworks (MOFs) with the perovskite ABX3 architecture. J. Am. Chem. Soc. 2009, 131, 13625–13627. [Google Scholar] [CrossRef]
  7. Tian, Y.; Stroppa, A.; Chai, Y.; Yan, L.; Wang, S.; Barone, P.; Picozzi, S.; Sun, Y. Cross coupling between electric and magnetic orders in a multiferroic metal-organic framework. Sci. Rep. 2015, 4, 6062. [Google Scholar] [CrossRef] [Green Version]
  8. Mączka, M.; Gągor, A.; Ptak, M.; Paraguassu, W.; Da Silva, T.A.; Sieradzki, A.; Pikul, A. Phase transitions and coexistence of magnetic and electric orders in the methylhydrazinium metal formate frameworks. Chem. Mater. 2017, 29, 2264–2275. [Google Scholar] [CrossRef]
  9. Bermúdez-Garcia, 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] [Green Version]
  10. 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+). Dalton Trans. 2019, 48, 13006–13016. [Google Scholar] [CrossRef]
  11. Geng, F.-J.; Zhou, L.; Shi, P.-P.; Wang, X.-L.; Zheng, X.; Zhang, Y.; Fu, D.-W.; Ye, Q. Perovskite-type organic–inorganic hybrid NLO switches tuned by guest cations. J. Mater. Chem. C 2017, 5, 1529–1536. [Google Scholar] [CrossRef]
  12. Mączka, M.; Ptak, M.; Gągor, A.; Stefańska, D.; Zaręba, 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]
  13. 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]
  14. Mączka, M.; Collings, I.E.; Furtado Leite, F.; Paraguassu, W. Raman and single-crystal X-ray diffraction evidence of pressure-induced phase transitions in a perovskite-like framework of [(C3H7)4N][Mn(N(CN)2)3]. Dalton Trans. 2019, 48, 9072–9078. [Google Scholar] [CrossRef]
  15. Mączka, M.; Ptak, M.; Gągor, A.; Sieradzki, A.; Peksa, P.; Usevicius, G.; Simenas, M.; Leite, F.F.; Paraguassu, W. Temperature- and pressure-dependent studies of a highly flexible and compressible perovskite-like cadmium dicyanamide framework templated with protonated tetrapropylamine. J. Mater. Chem. C 2019, 7, 2408–2420. [Google Scholar] [CrossRef]
  16. Wu, Y.; Shaker, S.; Brivio, F.; Murugavel, R.; Bristowe, P.D.; Cheetham, A.K. [Am]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] [Green Version]
  17. Gao, H.-Q.; Wei, W.-J.; Tan, Y.-H.; Tan, Y.-Z. Phase Transition and negative Thermal Expansion in Guanidinium Magnesium-Hypophosphite Hybrid Perovskite. Chem. Mater. 2020, 32, 6886–6891. [Google Scholar] [CrossRef]
  18. Wu, Y.; Halat, D.M.; Wei, F.; Binford, T.; Seymour, I.D.; Gaultois, M.W.; Shaker, S.; Wang, J.; Grey, C.P.; Cheetham, A.K. Mixed X-site formate–hypophosphite hybrid perovskites. Chem. A Eur. J. 2018, 24, 11309–11313. [Google Scholar] [CrossRef]
  19. Wu, Y.; Binford, T.; Hill, J.A.; Shaker, S.; Wang, J.; Cheetham, A.K. Hypophosphite hybrid perovskites: A platform for unconventional tilts and shifts. Chem. Commun. 2018, 54, 3751–3754. [Google Scholar] [CrossRef]
  20. Ptak, M.; Mączka, M. Phonon properties and mechanism of order-disorder phase transition in formamidinium manganese hypophosphite single crystal. Spectrochim. Acta A 2020, 230, 118010. [Google Scholar] [CrossRef]
  21. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Montgomery, J.A., Jr.; Vreven, T.; Kudin, K.N.; Burant, J.C.; et al. Gaussian 03, Revision A.1; Gaussian, Inc.: Pittsburgh, PA, USA, 2003. [Google Scholar]
  22. Becke, D. Density-functional thermochemistry. IV. A new dynamical correlation functional and implications for exact-exchange mixing. J. Chem. Phys. 1996, 104, 1040–1046. [Google Scholar] [CrossRef]
  23. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [Green Version]
  24. Parr, R.G.; Yang, W. Density-Functional Theory of Atoms and Molecules; Oxford University Press: New York, NY, USA, 1989. [Google Scholar]
  25. McLean, A.D.; Chandler, G.S. Contracted Gaussian basis sets for molecular calculations. I. Second row atoms, Z = 11–18. J. Chem. Phys. 1980, 72, 5639–5648. [Google Scholar] [CrossRef]
  26. Krishnan, R.; Binkley, J.S.; Seeger, R.; Pople, J.A. Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650–654. [Google Scholar] [CrossRef]
  27. Zhurko, G.A.; Zhurko, D.A. Chemcraft Graphical Program of Visualization of Computed Results. Available online: http://chemcraftprog.com (accessed on 6 July 2020).
  28. Rostkowska, H.; Lapinski, L.; Nowak, M.J. Analysis of the normal modes of molecules with D3h symmetry Infrared spectra of monomeric s-triazine and cyanuric acid. Vib. Spectrosc. 2009, 49, 43–51. [Google Scholar] [CrossRef]
  29. Tabacik, V.; Abenoza, M. Quasi general harmonic force field and mode anharmonicity of the hypophosphite ion normal vibrations. J. Mol. Struct. 1975, 27, 369–381. [Google Scholar] [CrossRef]
  30. Tsuboi, M. Vibrational Spectra of Phosphite and hypophosphite anions, and the characteristic frequencies of PO3 and PO2 groups. J. Am. Chem. Soc. 1957, 79, 1351–1354. [Google Scholar] [CrossRef]
  31. Liu, R.; Moody, P.R.; Vanburen, A.S.; Clark, J.A.; Crauser, J.A.; Tate, D.R. On assignemnt of fundametal vibrational modes of hypophosphite anion and its deuterated analogue. Vib. Spectrosc. 1996, 10, 325–329. [Google Scholar] [CrossRef]
  32. Tanner, P.A.; Yu-Long, L.; Mak, T.C.W. Synthesis, crystal structures and vibrational spectra of zinc hypophosphites. Polyhedron 1997, 16, 495–505. [Google Scholar] [CrossRef]
Sample Availability: Samples of the compound ([CH3NH2NH2][Mn(H2PO2)3] is available from the authors.
Figure 1. The crystal structure packing of [MHy][Mn(H2POO)3] taken from Ref. [5] (a) and the structural model of MHy+ cation (with the atomic numbering used in the Density Functional Theory (DFT) calculations) and H2POO anion (b).
Figure 1. The crystal structure packing of [MHy][Mn(H2POO)3] taken from Ref. [5] (a) and the structural model of MHy+ cation (with the atomic numbering used in the Density Functional Theory (DFT) calculations) and H2POO anion (b).
Molecules 25 05215 g001
Figure 2. The polycrystalline IR and Raman spectra of the [MHy][Mn(H2POO)3] perovskite measured at room temperature.
Figure 2. The polycrystalline IR and Raman spectra of the [MHy][Mn(H2POO)3] perovskite measured at room temperature.
Molecules 25 05215 g002
Figure 3. Polarized Raman spectra of the [MHy][Mn(H2POO)3] single crystal in the (a) 2200–3360 cm−1 and (b) 50–1720 cm−1 range.
Figure 3. Polarized Raman spectra of the [MHy][Mn(H2POO)3] single crystal in the (a) 2200–3360 cm−1 and (b) 50–1720 cm−1 range.
Molecules 25 05215 g003
Figure 4. The temperature evolution of Raman wavenumbers of bands corresponding to vibrations of groups that built up (a) H2PO2 and (b) MHy+ ions.
Figure 4. The temperature evolution of Raman wavenumbers of bands corresponding to vibrations of groups that built up (a) H2PO2 and (b) MHy+ ions.
Molecules 25 05215 g004
Figure 5. The temperature evolution of Raman full widths at half maximum (FWHMs) of bands corresponding to vibrations of groups that built up (a) H2PO2 and (b) MHy+ ions.
Figure 5. The temperature evolution of Raman full widths at half maximum (FWHMs) of bands corresponding to vibrations of groups that built up (a) H2PO2 and (b) MHy+ ions.
Molecules 25 05215 g005
Table 1. Calculated (harmonic and anharmonic) wavenumbers for the MHy+ cation.
Table 1. Calculated (harmonic and anharmonic) wavenumbers for the MHy+ cation.
No.νharmνharm 1νanharmPED
1356434213388νasNH2-99
2347733383306νsNH2-100
3346133233292νasNH2+-99
4342532883289νsNH2+-98
5318030533049νasCH3-100
6316630393029νasCH3-100
7307529523014νsCH3-100
8170016661679δNH2-80 + δNH2+-20
9167016371607δNH2+-80 + δNH2-20
10149914691472δasCH3-91
11149914691458δasCH3-62 + ρNH2-22 + τNH2+-14
12149014601449τNH2+-38 + δasCH3-35 + ρNH2-26
13146114321439δsCH3-97
14144314151401ωNH2+-86 + δasCH3-15
15133613091297τNH2+-41 + ρCH3-36 + ρNH2-22
16121911941176ρCH3-50 + ωNH2-25 + νNH2+-NH2-13 + δNH2+-13
17111710951091ρCH3-42 + ρNH2-25 + τNH2+-32
18106510431031ωNH2-37 + ρCH3-30 + νNH2+-NH2-20 + νNH2+-CH3-12
19953934907νNH2+-CH3-48 + ωNH2-27 + νNH2+-NH2-22
20841824821ρNH2+-75 + ρCH3-22
21836819811νNH2+-NH2-46 + νNH2+-CH3-37 + ωNH2-12
22405397404δCNN-87
23281275252τNH2-73 + τCH3-27
24225220245τCH3-73 + τNH2-27
1 scaling factor = 0.98 (2499–0 cm−1) + 0.96 (3500–2500 cm−1); abbreviations: in-plane vibrations: ν—stretching; δbending; ρ—rocking; out-of-plane vibrations: ω—wagging; τ—twisting
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Ciupa-Litwa, A.; Ptak, M.; Kucharska, E.; Hanuza, J.; Mączka, M. Vibrational Properties and DFT Calculations of Perovskite-Type Methylhydrazinium Manganese Hypophosphite. Molecules 2020, 25, 5215. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25215215

AMA Style

Ciupa-Litwa A, Ptak M, Kucharska E, Hanuza J, Mączka M. Vibrational Properties and DFT Calculations of Perovskite-Type Methylhydrazinium Manganese Hypophosphite. Molecules. 2020; 25(21):5215. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25215215

Chicago/Turabian Style

Ciupa-Litwa, Aneta, Maciej Ptak, Edyta Kucharska, Jerzy Hanuza, and Mirosław Mączka. 2020. "Vibrational Properties and DFT Calculations of Perovskite-Type Methylhydrazinium Manganese Hypophosphite" Molecules 25, no. 21: 5215. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules25215215

Article Metrics

Back to TopTop