Open Access Article
Anand Roy,
Anjali Singh,
S. Assa Aravindh,
Swaraj Servottam,
Umesh V. Waghmare and
C. N. R. Rao
*
New Chemistry Unit, Sheikh Saqr Laboratory, School of Advance Materials, Theoretical Science Unit, Jawaharlal Nehru Centre for Advanced Scientific Research, Jakkur, P. O., 560064, Bangalore, India. E-mail: cnrrao@jncasr.ac.in
First published on 31st January 2020
Cadmium phosphochlorides, Cd4P2Cl3 and Cd7P4Cl6, possess cadmium atoms differently bonded to chlorine and phosphorus ligands. A combined experimental and theoretical study has been carried out to examine the effect of manganese substitution in place of cadmium in these compounds. Experimentally it is found that manganese prefers the Cd7P4Cl6 phase over Cd4P2Cl3. First-principles calculations reveal, stabilization of Cd7P4Cl6 upon Mn-substitution with a significant reduction in the formation energy when Mn2+ is substituted at Cd-sites coordinated octahedrally by Cl-ligands. Substitution of Mn2+ at two different Cd-sites in these compounds not only alters their formation energy differently but also causes a notable change in the electronic structures. In contrast to n-type conductivity in pristine Cd7P4Cl6, Mn2+ substituted Cd7−yMnyP4Cl6 analogues exhibit p-type conductivity with a remarkable enhancement in the photochemical HER activity and stability of the system. Photochemical properties of pristine and substituted compounds are explained by studying the nature of charge carriers and their dynamics.
A striking difference between cadmium phosphochlorides and other semiconductor materials is the unique structural feature of the former wherein the unit cell consists of two distinct kinds of Cd-centers bonded differently to phosphorus and chlorine, giving rise to the possibility of selectivity in Mn-substitution between the two distinct Cd-sites. We chose the thermodynamically most stable Cd4P2Cl3 which is the closest analogue of CdS, to study the effect of manganese substitution. To our surprise, we observe the formation of the low-temperature Cd7P4Cl6 phase at higher temperatures in the presence of manganese. In order to understand this phenomenon and to explore the properties of the Mn-substituted phosphochlorides, we have carried out detailed first-principles calculations along with experimental studies. Theoretical calculations indicate that substitution of manganese in place of cadmium stabilizes the phosphochlorides, Mn-substitution being more favoured in Cd7P4Cl6 than in Cd4P2Cl3. Substitution of Mn2+ at the two different Cd-sites in Cd7P4Cl6 not only affects the formation energy differently but also changes the electronic band structures in a contrasting manner. The photochemical hydrogen evolution reaction (HER) activity of p-type Cd7−yMnyP4Cl6 is vastly superior to that of the pristine material wherein Cd5.8Mn1.2P4Cl6 shows an apparent quantum yield (AQY) of 47.6%. The influence of nature of charge carriers and their mechanism of action in pristine and Mn2+ substituted analogues have been studied.
). The presence of even small proportion of MnCl2 in the precursor mixture (XMn2+ = 0.04 with respect to Cd2+) affects the formation of Cd4P2Cl3 and gives rise to Cd7P4Cl6 (cubic, Pa
) (∼12%) along with Cd4P2Cl3 (∼88%) (Fig. 1a). Increase in the proportion of MnCl2 (XMn2+ = 0.08) results in an increase in the Cd7P4Cl6 in the mixture from 12 to 20%. Thus, above XMn2+ of 0.08 a greenish-yellow (major product) and an orange-yellow (minor product) compound, corresponding to pure Cd7P4Cl6 and mainly Cd4P2Cl3 are formed respectively (Fig. 1a and b). Furthermore, an increase in the MnCl2 content (XMn2+ ≥ 0.15) results in the exclusive formation of pure Cd7P4Cl6 with only a trace of Cd4P2Cl3 (Fig. 1b). Comparative PXRD patterns (using zero background sample holder) of pristine Cd7P4Cl6 and Mn-substituted Cd5.8Mn1.2P4Cl6 in the selective 2θ region is shown in Fig. 1c. A right shift (towards higher 2θ) in the diffraction peaks of Cd5.8Mn1.2P4Cl6 is observed with respect to the pristine compound, indicating a decrease in the lattice parameter of the system as a result of substitution. The Rietveld refinement of Cd5.8Mn1.2P4Cl6 indeed revealed a reduction in the lattice parameter (a = 11.926 Å) with respect to the pristine Cd7P4Cl6 (a = 11.940 Å)10 which can be attributed to the smaller size of Mn2+ ions (Fig. 1d).
In order to estimate the manganese concentration in substituted compounds, we have carried out ICP-OES measurements. Samples were thoroughly washed in H2O–C2H5OH (1
:
3 volume ratio) mixture to remove the unreacted Mn-salt present on the surface. The substitution of manganese is more favoured in Cd7P4Cl6 compare to Cd4P2Cl3. Substitution of ≥12% manganese in Cd7−yMnyP4Cl6 in place of cadmium causes it to crystallize in pure form free from other compositions. We have obtained Cd7−yMnyP4Cl6 with y = 0.9–1.5, giving rise to the compositions, Cd6.1Mn0.9P4Cl6, Cd5.8Mn1.2P4Cl6, and Cd5.5Mn1.5P4Cl6 (Fig. 2a). The EDAX spectrum of Cd5.8Mn1.2P4Cl6 showed signals corresponding to Cd, Mn, P, and Cl throughout the sample, indicating uniform substitution of manganese (Fig. S1a†). An XPS scan of Cd5.8Mn1.2P4Cl6 revealed signals corresponding to Cd, Mn, P, and Cl (Fig. S1b†). High-resolution XPS-core level spectra of Cd and Mn (Fig. S1c and d†) show energy differences of 6.8 and 11.2 eV between 3d5/2 and 3d3/2 (in the case of Cd) and 2p3/2 and 2p1/2 (in the case of Mn) respectively, characteristic of Cd2+ and Mn2+ states. Phosphorus atoms in Cd7P4Cl6 form P–P dumbbell, wherein each P atom in dumbbell is bonded to 3 Cd atoms (P2Cd6 unit). The P–P bounded P atom possess a charge of (P2)4−. The crystal structure of Cd7P4Cl6 possess cationic three dimensional ∞3[Cd3P2]2+ and anionic [CdCl6]4− units and the crystallochemical formula can be written as [Cd3(P2)4−]2·[CdCl6] wherein Cd possess +2 oxidation state.10 In contrast, Cd4P2Cl3 has two different natures of phosphorus atoms wherein octahedrally coordinated (P2)4− dumbbell coexist with tetrahedrally coordinated P3− species (PCd4). The crystallochemical formula of Cd4P2Cl3 is Cd4(P3−)2(P2)4−Cl3.10 Fitting of core-level XPS signal of cadmium in pristine as well as substituted compounds reveal +2 oxidation state of cadmium and we did not observe any other Cd-species (Fig. S2†). Electron paramagnetic resonance (EPR) spectrum of Cd5.8Mn1.2P4Cl6 at room temperature showed a peak with ‘g’ value corresponding to 2.0059, confirming Mn2+ species in the compound. The absence of the hyperfine structure in the EPR spectrum is because of the high doping concentration of manganese in the host semiconductor (Fig. 2b).
![]() | ||
| Fig. 2 (a) Mole fractions of manganese in product (obtained by ICP-OES) as a function of the mole fractions of manganese used in the precursor. (b) Room temperature EPR spectrum of Cd5.8Mn1.2P4Cl6. | ||
Diffuse reflectance spectra of Mn-substituted Cd7−yMnyP4Cl6 compounds were collected in the powder form and converted into the absorbance mode employing Kubelka–Munk equation (Fig. S3a†). A small blue-shift (0.03 to 0.07 eV) occurs in the absorption spectra of Mn-substituted compounds compared to the pristine Cd7P4Cl6 (band gap ∼ 2.62 eV; λedge ∼ 472.30 nm). Tauc plots show linear behaviour between (αhν)1/n vs. hν for n = 1/2, indicating direct band gap in the compounds (Fig. S3b†).
In Cd6.75Mn0.25P4Cl6, (a) with Mn2+ substituted at the Cd1 site (MnCd1) (Fig. 3a), Mn–P, Mn–Cl1, Mn–Cl2, Mn–Cl3 and Mn–Cl4 bond lengths are 2.47 Å, 2.42 Å, 2.50 Å, 3.50 Å and 3.69 Å (Fig. 3a and b) and (b) with Mn2+ substituted at Cd2 site (MnCd2), Mn–Cl bond length is 2.58 Å (Fig. 3c and d). In Cd3.875Mn0.125P2Cl3, (a) MnCd1, Mn–P and Mn–Cl bond lengths are 2.42 Å and 2.44 Å (Fig. 4a and b) and (b) MnCd2, Mn–P1 (Mn–P2) and Mn–Cl1 (Mn–Cl2) bond lengths are 2.43 Å (2.45 Å) and 2.44 Å (2.47 Å) (Fig. 4c and d). Clearly, there is a significant reduction in metal–X bond length with Mn2+ substitution, as expected from their ionic radii.
We estimate the formation energy (Ef) of Mn2+ substitution in Cd7P4Cl6 and Cd4P2Cl3 to find out the favourable site for Mn substitution,
| Ef = EMn–Cd1/Cd2 − EB − EMn + ECd |
| Compound | Ef (eV) | ΔE(Emag − Enon-mag) (eV) | Magnetic moment (μB) |
|---|---|---|---|
| Cd6.75Mn0.25P4Cl6(MnCd1) | −6.59 | −2.20 | 4.43 |
| Cd6.75Mn0.25P4Cl6(MnCd2) | −7.10 | −2.20 | 4.61 |
| Cd3.875Mn0.125P2Cl3(MnCd1) | −6.77 | −2.67 | 4.47 |
| Cd3.875Mn0.125P2Cl3(MnCd2) | −6.67 | −2.24 | 4.40 |
| MnCd7P4Cl6 | −3.12 | −2.28 | 4.50 |
| MnCd4P2Cl3 | −5.04 | −1.58 | 3.58 |
We now examine the electronic structures of Mn2+ substituted Cd7P4Cl6 and Cd4P2Cl3 compounds and compare them with those of the pristine forms to understand the effects of substitution. Cd6.75Mn0.25P4Cl6 (MnCd1) is a direct (at Γ-point) band gap semiconductor with a gap of 1.4 eV (Fig. 5b). The band gap of Cd7P4Cl6(MnCd1) reduces by 0.4 eV upon Mn2+ substitution at Cd1 site. The states at valence band maximum (VBM) and conduction band minimum (CBM) are spin split, with VBM is constituted of spin-up states whereas CBM is constituted of spin-down states. The VBM is mainly contributed by spin-up states of P 3p orbitals whereas CBM is constituted of spin-down states of Mn-3d orbitals (Fig. 6a). Noting that Cd6.75Mn0.25P4Cl6(MnCd2) is the lowest energy configuration, we find from its electronic structure that it is a direct (at Γ-point) gap semiconductor with a band gap of 1.8 eV, without any notable change in the gap with respect to pristine Cd7P4Cl6 (Fig. 5c). In contrast to Cd6.75Mn0.25P4Cl6(MnCd1), we do not find spin split states at VBM and CBM. The VBM is mainly contributed by spin-up and spin-down states of P 3p and Cl 3p orbitals whereas CBM is constituted of spin-down states of Mn 3d orbitals (Fig. 6b). Cd3.875Mn0.125P2Cl3(MnCd1) and Cd3.875Mn0.125P2Cl3(MnCd2) are direct (at Γ-point) gap semiconductor with a band gap of 1.3 and 1.0 eV respectively (Fig. 5e and f). Mn2+ substitution in Cd4P2Cl3 at Cd2 site affects its band gap notably. From the projected density of states, VBM of Cd3.875Mn0.125P2Cl3(MnCd1) is mainly contributed by spin-up states of P 3p and Mn 3d orbitals, and CBM is constituted of spin-down states of Mn 3d orbitals (Fig. 6c). States at the VBM of Cd3.75Mn0.25P2Cl3(MnCd2) are mainly contributed by spin-up states of P 3p and Mn 3d orbitals (Fig. 6d). CBM is constituted of spin-down states of Mn 3d and 5s Cd orbitals (Fig. 6d). We find spin split states at VBM and CBM of those configurations which have Mn2+ coordinated with a greater number of P atoms, e.g., Mn2+ is octahedrally coordinated with 4 Cl and 2 P in Cd6.75Mn0.25P4Cl6(MnCd1) whereas, in Cd6.75Mn0.25P4Cl6(MnCd2), Mn is octahedrally coordinated with 6 Cl atoms (no P atoms). Similarly, Mn2+ is tetrahedrally coordinated with 3 Cl, and 1 P in Cd3.875Mn0.125P2Cl3(MnCd1) and Mn2+ is tetrahedrally coordinated with 2 P and 2 Cl in Cd3.875Mn.125P2Cl3(MnCd2). Splitting of spin-up and spin-down bands arises due to broken inversion and crystal field symmetry in the system. Cd6.75Mn0.25P4Cl6(MnCd1) exhibits spin-split states at VBM and CBM as Mn2+ is bonded to 4 Cl and 2 P atoms, forming a distorted octahedra with broken inversion symmetry (Fig. 3a). In contrast, we do not find spin split states in Cd6.75Mn0.25P4Cl6(MnCd2) since Mn2+ is bonded to 6 Cl atoms forming symmetric octahedra maintaining inversion symmetry (Fig. 3d). Similarly, Mn2+ is at the center of proper and distorted tetrahedra in Cd3.875Mn0.125P2Cl3(MnCd1) and Cd3.875Mn0.125P2Cl3(MnCd2) respectively (Fig. 4b and d). Both these compounds exhibit spin-split states at VBM and CBM; the magnitude of splitting is weak in Cd3.875Mn0.125P2Cl3(MnCd1) as the P atom is along the principal axis and 3-fold symmetry is maintained by 3 Cl atoms bonded to Mn. In Cd3.875Mn0.125P2Cl3(MnCd2), Mn2+ is bonded to 2 P and 2 Cl atoms, leading to broken 3-fold symmetry and stronger spin splitting of energy bands. Hence, Cd3.875Mn0.125P2Cl3(MnCd2) and Cd6.75Mn0.25P4Cl6(MnCd1) show spin-spit states at VBM and CBM. From the spin-resolved projected electronic density of states (DOS) (Fig. 6) of Mn2+ substituted Cd7P4Cl6 and Cd4P2Cl3 compounds, it is evident that there is a strong coupling between the Mn 3d orbitals and P 3p and Cl 3p orbitals of the host semiconductor.
Our estimates of magnetic moments of Mn-doped Cd7P4Cl6 and Cd4P2Cl3 are tabulated in Table 1. We find that 3d orbitals of Mn are occupied as (dxy↑)1 (dyz↑)1 (dzx↑)1 (dz2↑)1 (dx2−y2↑)1, resulting in an electronic configuration with S = 5.0 μB. However, hybridizing interaction of 3d states of Mn with 3p states of Cl and P atoms results in partial cancellation of the magnetic moment μ = 5 − δ, resulting in effective local magnetic moments 4.61 μB in Cd6.75Mn0.25P4Cl6(MnCd2). Cd6.75Mn0.25P4Cl6(MnCd2) has a higher magnetic moment than of that of Mn-substituted Cd7P4Cl6 and Cd4P2Cl3 compounds.
Halogen vacancies are inherently present in cadmium phosphochlorides compounds.11 The exact composition of Cd7−yMnyP4Cl6 compounds by EDAX and XPS measurements reveal some amount of Cl and P vacancies in these compounds. We simulated Cl and P vacancies in energetically most favourable Mn2+ substituted (at Cd2 site) Cd7P4Cl6, considering 5 inequivalent configurations of vacancy pairs. In the configuration I, we removed a pair of P atoms from the crystal, each of which is bonded with three Cd1 atoms. Two Cl atoms are removed from octahedral coordination of Mn atom in configuration II, whereas one atom is removed from 6 Cl atoms bonded with Mn, and another Cl is extracted from octahedrally coordinated Cd2 atom in configuration III. Configurations IV and V have mixed pairs of Cl and P vacancies, in both cases, Cl atom is taken from octahedrally coordinated Mn atom, and P is removed from P1–P2 pair, such that the distance of Cl vacancy from P1 and P2 atoms is 5.33 Å and 3.85 Å respectively. We simulate pure Cl and P vacancy pairs in pristine Cd7P4Cl6, in the configuration I, we remove a P pair from crystal and remove 2 Cl atoms from octahedrally coordinated Cd2 atom in configuration II. From the formation energies of vacancies (Table 2), we conclude that it is relatively easy to introduce a Cl vacancy than a P vacancy in pristine as well as Mn substituted Cd7P4Cl6 compound. Secondly, it is easier to introduce P vacancy in Mn-substituted Cd7P4Cl6 compound than in the pristine one. Cl vacancies are easier to be introduced in Mn substituted Cd7P4Cl6 compound than in the pristine one when Cl atom is removed each from octahedrally coordinated Mn and Cd2 atoms rather than introducing both the vacancies at octahedrally coordinated Mn or Cd2 atom. We present the electronic structure of the configurations III and V as configuration III has the lowest defect formation energy, and configuration V has lower Ef than of I and IV configurations and understands the combined effect of Cl and P vacancies. In the electronic structure, we find that Cl and P vacancy pairs make Cd6.75Mn0.25P4Cl6(MnCd2) an indirect (Γ–R) band gap semiconductor with gaps of 1.68 eV and 1.14 eV for III and V configurations (Fig. 7a and c). Associated with Cl vacancy, we identify a new defect band on the bottom of the CBM (Fig. 7a), which is also seen in the density of states (Fig. 7b). The peak has contributions mainly from minority spin states of Cl-3p and Cd-5s orbitals. The introduction of P and Cl vacancies leads to defect bands just above the VBM and below the CBM (Fig. 7c), as evident in the density of states (Fig. 7d). Spin-up and spin-down states at VBM and CBM here are mainly contributed by P 3p, Cd 5s, and Cd 5p orbitals.
| Compound | Vacancy configuration | Ef (eV/vacancy) |
|---|---|---|
| Cd7P4Cl6 | I (2 P) | 7.86 |
| Cd7P4Cl6 | II (2 Cl) | 4.65 |
| Cd6.75Mn0.25P4Cl6(MnCd2) | I (2 P) | 7.73 |
| Cd6.75Mn0.25P4Cl6(MnCd2) | II (2 Cl) | 4.84 |
| Cd6.75Mn0.25P4Cl6(MnCd2) | III (Cl–Cl) | 4.48 |
| Cd6.75Mn0.25P4Cl6(MnCd2) | IV (Cl–P1) | 7.16 |
| Cd6.75Mn0.25P4Cl6(MnCd2) | V (Cl–P2) | 7.00 |
In order to understand the enhancement in the photochemical HER activity, we have studied photogenerated charge carrier dynamics of these compounds using photoelectrochemical measurements. The open-circuit potential of the photoelectrodes demonstrates the process of charge carrier generation and recombination in materials.12,13 Pristine Cd7P4Cl6 exhibits a cathodic shift in the potential upon light irradiation, indicating n-type conductivity, similar to the earlier report (Fig. 9a).14 In contrast to the cathodic shift of the pristine compound, the manganese substituted analogue (Cd5.8Mn1.2P4Cl6) exhibits anodic shift, indicating p-type conductivity (Fig. 9b). In order to verify the nature of conduction in these compounds, Hall measurements were carried out at room temperature on the rectangular pallet of samples. Unlike the negative Hall coefficient of pristine Cd7P4Cl6 (n-type conductivity) (Fig. 9c), Cd5.8Mn1.2P4Cl6 exhibited positive Hall coefficient due to p-type conductivity (Fig. 9d). The variation of Hall voltage with the applied positive and negative magnetic field is given in ESI (Fig. S6†).
![]() | ||
| Fig. 9 Open circuit potential (OCP) of (a) pristine Cd7P4Cl6, and (b) Cd5.8Mn1.2P4Cl6. Variation of the Hall voltage with applied magnetic field in (c) Cd7P4Cl6 and (d) Cd5.8Mn1.2P4Cl6. | ||
The origin of the change in conductivity upon Mn-substitution can be understood by the electronic structure of compounds (Fig. 6). The electronic states at the CBM are dominated by 3d states of Mn. Their narrow bandwidth indicates their localized nature and hence, a large effective mass and lower mobility of the electron carriers. In contrast, the states in the VBM are dominated by p-orbital of Cl and P strongly hybridized with 3d orbitals of Mn, producing notable dispersion of bands giving a low effective mass and high mobility of hole carriers. Hence, holes contribute more significantly to conduction than electrons upon Mn-substitution in these compounds causing a change in n to p-type behaviour. We note that the overall p-type conduction in Mn-substituted compounds is weaker than the n-type conduction of the pristine ones.
Since the photocatalytic activity of semiconductor materials is prone to the nature of charge carriers, we have examined the effect of the change in the type of majority carriers on the photocatalytic properties.14 It is known that when an extrinsic semiconductor comes in contact with an electrolyte bending of bands occurs at the surface to equilibrate the Fermi level.15 An n-type semiconductor with the Fermi level near CBM undergoes upward band bending in contact with the electrolyte, resulting in electron migration towards the bulk of the semiconductor and hole migration towards the interface.14 This causes a decrease in the hot electron concentration at the interface and negatively impacts the HER (Fig. 10). In contrast, a p-type semiconductor with the Fermi level close to the VBM undergoes a downward band bending in the contact of the electrolyte. Such bending of the band helps in the separation of electron and holes at the surface wherein the holes are thrown to the bulk of semiconductor and electron are pushed to the space charge region, resulting in the enrichment of photogenerated electron population in the space charge region (semiconductor–H2O interface) (Fig. 10).16 A greater population of hot electrons at the interface (space charge region) positively impacts the reduction of H2O into H2. These observations lead support to our finding that p-type Cd5.8Mn1.2P4Cl6 exhibits superior photochemical HER performance than their pristine n-type analogues.
Footnote |
| † Electronic supplementary information (ESI) available: EDAX and XPS spectra, UV/vis. absorption spectra, Tauc plots, PXRD of compounds before and after photochemical reactions, Hall measurement, apparent quantum yield calculations. See DOI: 10.1039/c9ra10711h |
| This journal is © The Royal Society of Chemistry 2020 |