Muhammad Awais Alia,
Maryam Noor Ul Aina,
Asim Mansha*a,
Sadia Asimb and
Ameer Fawad Zahoor*a
aDepartment of Chemistry, Government College University Faisalabad, Pakistan. E-mail: asimmansha@gcuf.edu.pk; fawad.zahoor@gcuf.edu.pk
bDepartment of Chemistry, Government College Women University Faisalabad, Pakistan
First published on 17th October 2024
First-principles density functional investigations of the structural, electronic, optical and thermodynamic properties of K3VO4, Na3VO4 and Zn3V2O8 were performed using generalized gradient approximation (GGA) via ultrasoft pseudopotential and density functional theory (DFT). Their electronic structure was analyzed with a focus on the nature of electronic states near band edges. The electronic band structure revealed that between 6% Fe and 6% Co, 6% Co significantly tuned the band gap with the emergence of new states at the gamma point. Notable variations were highlighted in the electronic properties of Na3V(1−x)FexO4, Na3V(1−x)CoxO4, K3V(1−x)FexO4, K3V(1−x)CoxO4, Zn3(1−x)V2(1−x)CoxO8 and Zn3(1−x)V2(1−x)FexO8 (where x = 0.06) due to the different natures of the unoccupied 3d states of Fe and Co. Density of states analysis as well as α (spin up) and β (spin down) magnetic moments showed that cobalt can reduce the band gap by positioning the valence band higher than O 2p orbitals and the conduction band lower than V 3d orbitals. Mulliken charge distribution revealed the presence of the 6s2 lone pair on Zn, greater population and short bond length in V–O bonds. Hence, the hardness and covalent character develops owing to the V–O bond. Elastic properties, including bulk modulus, shear modulus, Pugh ratio and Poisson ratio, were computed and showed Zn3V2O8 to be mechanically more stable than Na3VO4 and K3VO4. Optimal values of optical properties, such as absorption, reflectivity, dielectric function, refractive index and loss functions, demonstrated Zn3V2O8 as an efficient photocatalytic compound. The optimum trend within finite temperature ranges utilizing quasi-harmonic technique is illustrated by calculating thermodynamic parameters. Theoretical investigations presented here will open up a new line of exploration of the photocatalytic characteristics of orthovanadates.
The major role of a photocatalyst is to absorb incident light and create charge carriers, i.e., electrons and holes. With the absorption of light having wavelengths equal to the band gap of the semiconductor, electrons are excited from the valence band (VB) to the conduction band (CB), creating holes in the VB.7 These charge carriers can then interact with pollutants, undergo oxidation–reduction processes, and result in products being adsorbed on the semiconductor's surface. The hole in the valence band and electrons in the conduction band can recombine at the same time. Effective light absorption, high redox potential, and limiting the photoinduced charge recombination are helpful methods of chemical conversion that improve the photocatalytic process efficiency.8,9
Recent studies on photosensitive materials for solar energy conversion have mainly focused on two primary goals: modification of conventional photocatalysts and development of a novel photocatalyst or photocatalytic system.10 Moreover, due to their relative stability against photocorrosion (photo-oxidative disintegration), environmental friendliness, and ease of production, metal oxide semiconductors had attracted a lot of attention.11 To date, the most efficient semiconductor photocatalysts to be reported are Fe2O3,12 ZrO2,13 ZnO,14 SnO2,15 TiO2 (ref. 16) and bismuth halides.17
Vanadates have gained much attention due to their interesting physicochemical properties and exceptional performance.18 They are preferred over titanium and zinc oxide photocatalysts due to their narrower band gap,19 lower recombination rate,20,21 and higher quantum efficiency.22 Simultaneously, the complex structural chemistry of vanadates, which readily adapt to various shapes, including tetrahedral, distorted octahedral and dodecahedral, demonstrates their unique character. Due to the structural differences, they may be applied as photocatalysts or optoelectronic materials. The V-3d orbital usually offers a lower conduction band in the band gap structure, which is why vanadates are regarded as potent visible light-driven photocatalysts.23 Having isolated discrete units, the orthovanadates (VO4−3) class is the simplest of all vanadates. Numerous visible-light activated photocatalysts, such as YVO4,24 InVO4,25 and BiVO4,26 have been reported for hydrogen production and the photodegradation of dyes.
This research work focuses on the systematic theoretical comparison of alkali-metal orthovanadate (K3VO4, Na3VO4) with transition metal orthovanadate (Zn3V2O8). Having exceptional chemical strength, low band gap, economical, and an ecofriendly nature, one of the most promising photocatalysts for pollutant degradation and water splitting process is Zn3V2O8. Sajid et al. reported that Zn3(VO4)2 and its hydrates are effective photocatalysts and solar energy conversion materials.27 Slobodin et al. proposed the color features and luminescence behavior of AVO3 (A = K, Rb, and Cs), and reported on the synthesis of M3V2O8 (M = Mg, Zn) by solid-state reaction method.28
Cobalt (Co) has been reported as an efficient dopant with a suitable valence band structure to photo-oxidize water to H2, while having a band gap covering the entire range of visible light. Examples include the Co-doped Fe2O3–TiO2,29 Co2+-doped-TiO2,30 LaCoxFe1−xO3 (x = 0.01, 0.05, 0.1),31 and BiFe1−xCoxO3 (x = 0, 0.025, 0.05, 0.075 & 0.10).32 In the same way, iron (Fe) has also been found as an effective dopant in various water splitting photocatalysts, such as Fe3+-doped TiO2,33 Fe-doped TiO2 and SrTiO3 nanotubes,34 Fe-doped BiVO4,35 and many others. Subsequent to the reported methodical photocatalytic efficiency of Fe and Co dopants, this research has also explored the enhancement of the photocatalytic efficiency with 6% Fe and 6% Co doped on Na3VO4, K3VO4 and Zn3V2O8.
A review of the literature revealed that neither theoretical nor experimental studies of the electronic, elastic, optical, and thermodynamic parameters of these compounds have been reported yet. Herein, we report the first attempt to computationally design these water spitting photocatalysts with theoretical insights into the evolution of (OH−˙) and O2−˙ radicals. The aforementioned motivation inspired us to carry out all these calculations using the full potential linear augmented plane wave (FPLAW) approach to ascertain the previous efforts on this subject, and offer reference data to experimentalists. Therefore, our investigation of all these parameters and band structure analysis with doping effects exhibited a reasonable explanation for the photocatalytic water splitting and degradation activity, which can assist in designing potentially effective vanadium-based photocatalyst materials.
Fig. 1 Mixed polyhedral and ball-and-stick model of the optimized crystal structures of (a) Zn3V2O8 (orthorhombic), (b) Na3VO4 (tetragonal) and (c) K3VO4 (tetragonal). |
There are two independent lattice parameters, such as ‘a’ and ‘c’, for the tetragonal phase structure. The unit cell parameters are summarized in Table 1. There are two inequivalent coordinated sites in the Na3VO4 crystal system shown in Fig. 1b. In the first site, it forms a distorted tetrahedron (NaO4) that gets bonded with four oxygen atoms, thereby sharing corners with the VO4 tetrahedron. Secondly, Na coordinates with four oxygen atoms and forms NaO4, which shares sites with the VO4 and NaO4 tetrahedra.42 Similar to Na3VO4, the tetragonal crystal structure K3VO4 crystallizes with the I2m space group, in which V ions are tetrahedrally coordinated. Meanwhile, K1+ ions have two types of chemically distinct environments, which are depicted in Fig. 1. In the first environment, K1+ is bonded to four oxygen atoms, forming a distorted trigonal pyramid that shares edges with the VO4 tetrahedra. At the second site, the K1+ ions combine with four oxygen atoms to form the KO4 tetrahedra.19
Crystals | Lattice constants | Angles | Cell volume (Å)3 | Lattice energy (eV) | ||||
---|---|---|---|---|---|---|---|---|
a (Å) | b (Å) | c (Å) | α | β | γ | |||
Zn3V2O8 | 6.15 | 11.70 | 8.34 | 90 | 90 | 90 | 600.87 | −50340.43 |
Zn3(1−0.06)V2(1−0.06)Fe0.06O8 | 6.25 | 11.87 | 8.13 | 90 | 90 | 90 | 603.62 | −52422.60 |
Zn3(1−0.06)V2(1−0.06)Co0.06O8 | 6.20 | 11.76 | 8.36 | 90 | 90 | 90 | 609.85 | −52831.13 |
Na3VO4 | 5.02 | 5.02 | 9.79 | 90 | 90 | 90 | 247.01 | −15288.88 |
Na3V(1−0.06)Fe0.06O4 | 5.06 | 5.06 | 9.78 | 90 | 90 | 90 | 250.42 | −15072.91 |
Na3V(1−0.06)Co0.06O4 | 5.09 | 5.09 | 9.80 | 90 | 90 | 90 | 254.42 | −15134.94 |
K3VO4 | 5.72 | 5.72 | 10.05 | 90 | 90 | 90 | 329.38 | −12140.37 |
K3V(1−0.06)Fe0.06O4 | 5.8 | 5.8 | 10 | 90 | 90 | 90 | 338.68 | −11985.15 |
K3V(1−0.06)Co0.06O4 | 5.95 | 5.95 | 9.31 | 90 | 90 | 90 | 329.59 | −11994.68 |
When foreign atoms are doped into the host lattice, the lattice distorts as a result of changes in the ionic radii and electronic states. However, the lattice distortions in the Co-doped and Fe-doped Zn3V2O8, K3VO4, and Na3VO4 are not apparent after placing the dopant atom, which could be explained by the low impurity concentration and similar radii.
Fig. 2 Brillouin zone showing k points of (a) orthorhombic Zn3V2O8 as well as (b) tetragonal Na3VO4 and K3VO4. |
Fig. 3 Band structure with the indicated dashed line at 0 eV as the Fermi level of (a) Zn3V2O8, (b) Zn3(1−0.06)V2(1−0.06)Fe0.06O8 and (c) Zn3(1−0.06)V2(1−0.06)Co0.06O8. |
Fig. 4 Band structure of (a) Na3VO4, (b) Na3V(1−0.06)Fe0.06O4, and (c) Na3V(1−0.06)Co0.06O4, in which the Fermi level is referenced as 0 eV as indicated by the horizontal dashed line. |
Fig. 5 Band structure of (a) K3VO4, (b) K3V(1−0.06)Fe0.06O4 and (c) K3V(1−0.06)Co0.06O4, in which the Fermi level is referenced as 0 eV as indicated by the horizontal dashed line. |
The band gap varies inversely with the atomic mass. This agrees with the sequence of band gap values for the selected compounds, such as Na3VO4 > K3VO4 > Zn3V2O8.51 However, among all compounds, the dramatic decrease in the band gap of Zn3V2O8 is due to the transition metal involving the d-orbital in bonding. Specifically, Zn (3d-orbital) is involved in the excitation of electrons from VB to CB.19
The upper valence band of Zn3V2O8 ranging from 0–3.5 eV is mainly dominated by the O-2p state, which is the major hallmark of oxide-based semiconductor materials, as shown in Fig. 6. The V-3d and Zn-3d orbitals are also involved in the upper valence band. One of the most striking attributes is the involvement of Zn-3d in the entire VB. However, p–d hybridization results in a strong covalent interaction between the Zn-3d and O-2p states.50 The calculated DOS and PDOS for the Fe-doped and Co-doped systems with 6% concentration are compared in Fig. 7a and b. The Fermi level of each doped system lies within the dopant states, which means that both dopants inherently have empty states that can accept the active electrons. The dopant states become eminently closer in the case of the Co-dopant, and the energy gap is noted to be significantly narrower for Zn3(1−0.06)V2(1−0.06)Co0.06O8. Moreover, expansion of the VBM can be attributed to the increasing contribution of Co-3d states in the conduction band, as evident from the PDOS plots shown in Fig. 6b. Similarly, Fig. 7a and b show that the contribution and distribution of O-2p states in the valence bands are not significantly altered.
Fig. 6 Calculated total density of states and partial density of states of pure Zn3V2O8 showing the O-2p states and V-3d states below and above the Fermi level (Eg). |
On the other hand, the VBM of the tetragonal phase system is composed of O 2p states. There is a contribution of V 3d, Na 3s, and K 4s in the CBM of these isostructural crystal systems, as shown in Fig. 8 and 10. Fig. 9a and 11a reveal that the Fe-dopant prominently induces its 3d before the V 3d state in Na3V(1−0.06)Fe0.06O4 and K3V(1−0.06)Fe0.06O4, thus placing the CBM near the Fermi level (Eg). Meanwhile, in the case of the Co-doped alkali metal orthovanadates, such as Na3V(1−0.06)Co0.06O4 and K3V(1−0.06)Co0.06O4, the CBM has been computed to shift closer to Eg, resulting in the vital reduction of the band gap due to the low-lying Co 3d state (Fig. 9a and b).
Fig. 8 Calculated total density of states (TDOS) and partial density of states of pure Na3VO4, showing O-2p and V-3d states below and above the Fermi level (Eg). |
Fig. 10 Calculated total density of states (TDOS) and partial density of states of pure K3VO4, showing the O-2p states and V-3d states below and above the Fermi level (Eg). |
In the total electron density, the major α (spin up) and β (spin down) components of the spin-polarized electrons are appreciably distinguished by the positive and negative spin densities of the s, p and d orbitals.56 The α and β spin states for the undoped and doped Zn3V2O8 Na3VO4 and K3VO4 are shown in Fig. 12–14, respectively. In Zn3V2O8, the valence band is attributed mainly to the α-spin states of O 2p with a minor contribution from the 3d α-spin electrons of Zn and V. Particularly, the higher absolute value of the α-dominated spin density of the V-3d electrons in the conduction band compared to that of the β-dominated electrons indicates the presence of more α electrons than β electrons, as manifested in Fig. 12a. However, the α-spin magnetic moment predominantly originates in the conduction band from the spin-polarized components of the V-3d and Co-3d orbitals in Zn2.82V1.88Co0.12O4, and the V-3d and Fe-3d orbitals in Zn2.82V1.88Fe0.12O4 with an energetic-preferential spin splitting near the Fermi level, as shown in Fig. 12b and c.
Fig. 13 Spin-up (↑) and spin-down (↓) partial density of states of (a) Na3VO4, (b)- Na3V(1−0.06)Co0.06O4, and (c) Na3V(1−0.06)Fe0.06O4 with indicated dashed lines as the Fermi level (Eg). |
Fig. 14 Spin-up (↑) and spin-down (↓) partial density of states of (a) K3VO4, (b) K3V(1−0.06)Co0.06O4, and (c) K3V(1−0.06)Fe0.06O4 with indicated dashed lines as the Fermi level (Eg). |
For Na3VO4 and K3VO4, the entire crystal lattice comprises the α-dominated spin states of O-2p and V-3d near the Fermi level (Fig. 13 and 14). The valence band maximum (VBM) and conduction band minimum (CBM) are mainly from the α-spin states of the O-2p and V-3d orbitals. Meanwhile, the 3d α-spin states of the Fe and Co dopants significantly shifted the CBM close to the Fermi level. However, the β-dominated electrons occur less frequently in VBM and CBM (Fig. 13b, c and 14b, c). It is evident from the spin polarization that Co and Fe are magnetic metals because both up-spin and down-spin DOS cross the Fermi level with a clear polarization between the DOS of the two spin channels.57 In Zn3V2O8, Fe and Co established the p-type extrinsic semiconductor with the shifting of the Fermi level towards VB.58 Meanwhile, in the case of Na3VO4 and K3VO4, DOS more prominently crosses the Fermi level, indicating Fe-doped and Co-doped p-type extrinsic semiconductors.
Compounds | Population R–O | Bond length R–O/Å | Bond length V–O/Å | Population V–O |
---|---|---|---|---|
Zn3V2O8 | 0.33, 0.20, 0.17 | 2.04, 2.23, 2.18 | 1.71, 1.74, 1.81 | 0.76, 0.67, 0.64 |
Na3VO4 | 0.10, 0.08 | 2.32, 2.36 | 1.73 | 0.74 |
K3VO4 | 0.04, −0.05 | 2.60, 2.8 | 1.74 | 0.76 |
The R and V atoms of the photocatalytic systems have higher positive charge, while oxygen has a negative charge. The data sets in Table 3 demonstrate that charge transfer is carried out from R and V to O atoms. Moreover, the higher positive charge transfer values demonstrated the greater overlap between the atoms of the bonding atoms.44
Compounds | Species | Mulliken atomic population | Mulliken charge/e | Hirshfeld charge/e | |||
---|---|---|---|---|---|---|---|
s | p | d | Total | ||||
Zn3V2O8 | Zn | 0.34, 0.31 | 0.65, 0.67 | 9.97, 9.98 | 10.97, 10.96 | 1.03, 1.04 | 0.46, 0.45 |
V | 2.25 | 6.43 | 3.35 | 12.03 | 0.97 | 0.46 | |
O | 1.86, 1.85 | 4.73, 4.79, 4.84 | 0.00 | 6.59, 6.70, 6.64 | −0.59, −0.70, −0.64 | −0.28, −0.29 | |
Na3VO4 | Na | 2.14 | 6.09, 6.10 | 0.00 | 8.24 | 0.76 | 0.34, 0.32 |
V | 2.25 | 6.82 | 3.42 | 12.48 | 0.52 | 0.49 | |
O | 1.87 | 4.83 | 0.00 | 6.70 | −0.70 | −0.37 | |
K3VO4 | K | 2.18, 2.11 | 5.81, 5.96 | 0.00, 0.00 | 7.99, 8.07 | 1.01, 0.93 | 0.36, 0.45 |
V | 2.27 | 6.87 | 3.44 | 12.59 | 0.41 | 0.45 | |
O | 1.90 | 4.92 | 0.00 | 6.82 | −0.82 | −0.40 |
The characteristic of a bond as either covalent or ionic can be evaluated by the overlap population of the orbitals. A complete ionic bond has a value of zero in the overlap bond population. Conversely, bonds having values greater than zero have a higher covalent character.59 So, the stronger covalency is implied by the greater bond population, whereas the opposite is true for the ionic character. The V–O bond populations for these three systems (0.76, 0.74, 0.64, 0.67) are much higher than the R–O bonds (0.04, −0.05, 0.10, 0.08, 0.33, 0.20, 0.17). As compared to the R–O bond, the covalent character is stronger in the V–O bond. Na3VO4 and Zn3V2O8 exhibit more covalent character, as listed in Table 2. Meanwhile, the bonding of K3VO4 is actually a mixture of covalent–ionic properties. Having greater bond population and shorter bond lengths, the V–O bonds also have a considerable impact on the hardness, instead of the R–O bonds, in the R3VO4 and R3V2O8 crystal systems.
The crystal structures and electronic band structures have led to the conclusion that vanadium atoms develop covalent and ionic interactions with O to produce VO43+, which confirmed the lack of a lone pair electron in the V atoms. Therefore, the interaction forces between R2+ and V5+ are directly related to the V–O bonds. The distortion in the Zn–O bond lengths in Zn3V2O8 is due to the 4s2 or 6s2 lone pair of the transition metal (Zn+2). The strong photocatalytic activity of visible light is greatly influenced by this attribute,44 which seems favorable for Zn3V2O8.
• Born stability standard for the tetragonal system (Na3VO4 and K3VO4):
C11 > |C12|; 2C132 < C33 (C11 + C12), C44 > 0; C66 > 0 |
• Criteria for the mechanical stability of the orthorhombic system (Zn3V2O8) is:
(C11 > 0; C11C12 > C122 |
C11C22C33 + 2C12C13C23 − C11C232 –C22C132 − C33C122 > 0 |
C44 > 0; C55 > 0; C66 > 0 |
Table 4 shows that the values of all elastic parameters are positive, and meet all aforementioned requirements. So, the crystal systems are mechanically stable.
Compounds | C11 | C33 | C44 | C66 | C12 | C13 | C16 | C22 | C23 | C55 |
---|---|---|---|---|---|---|---|---|---|---|
Zn3V2O8 | 173.98 | 164.66 | 53.86 | 19.44 | 52.25 | 87.42 | — | 182.03 | 97.04 | 61.23 |
Na3VO4 | 71.70 | 52.97 | 26.51 | 20.75 | 31.70 | 32.84 | 0.00 | — | — | — |
K3VO4 | 50.40 | 32.07 | 18.80 | 11.45 | 19.19 | 29.58 | 0.00 | — | — | — |
To assess the aggregate average elastic characteristics of the crystal structure, additional approximations named as isostress (Reuss state) and isostrain (Voigt state) were used. Hill demonstrated that the Voigt and Reuss approximations accurately exhibit the upper and lower limits of the elastic constants. The Voigt (Bv) and Reuss (BR) bulk modulus is given as:64
For the tetragonal system, the bulk modulus can be calculated as:65
The Voigt–Reuss shear modulus is then derived as:
The VRH averages for the shear modulus (G) and bulk modulus (B) are given as:
While the Young's modulus (E) and Poisson ratio (v) are computed as:
The bulk and shear modulus for the orthorhombic system can be calculated by the following equations:66
BR = χ(C11(C22 + C33 − 2C33) + C22(C33 − 2C13) − 2C33C12 + C12(2C23 − C12) + C13(2C12 − C13) + C23(2C13 − C23))−1 |
The way to calculate the Young's modulus and Poisson ratio is the same as stated for the tetragonal systems. The Voigt and Reuss equations define the upper and lower limits of polycrystalline elastic constants, respectively, and the arithmetic mean of these two limits is used to estimate the realistic bulk and shear modulus. The bulk modulus (B) measures the resistance to volume change, along with the bond length caused by applied pressure. The significant strength of the solid bonds is an indication of the higher bulk modulus.67 The condition of the bulk modulus stated as C12 < B < C11 seems to satisfy all these crystal systems.
Panda & Chandran68 developed equations to investigate the Young's modulus and Poisson ratio. The resistance to reversible deformation under shear stress was investigated by the shear modulus (G), which is known as the stiffness of a material. A greater shear modulus suggests a stronger degree of bonding between atoms.69 As a result, the Zn3V2O8 system is stiffer and exhibits stronger bonding than the other two systems.
Pugh's ratio defines the nature of bonding, which is either ionic or covalent, on the basis of the ductile and brittle character of the fabricated material. G/B < 6 indicates the ductility (ionic bonding) of the material, and the reverse is for the brittle (covalent bonding) nature.70 Therefore, K3VO4 and Zn3V2O8 have higher ductile nature.
The directionality of the covalent bonds is quantified by the Poisson ratio. In contrast to ionic materials, which have a Poisson ratio ≥0.25, covalent materials have a smaller value (v = 0.1). The lower and higher boundaries of the central force in solids are 0.25 and 0.5, respectively.71 Our results indicate that all compounds have ionic contribution and central interatomic forces.
Another crucial factor used to assess the constant nature of the structural qualities in all directions is the anisotropy of materials. A suitable parameter is required to describe the degree of anisotropy, as all single crystals are actually anisotropic. Anisotropy factors for the tetragonal structure are as follows:44
Using elastic constants, calculations can also be done for the orthorhombic system:
The calculated values of all these parameters are listed in Table 5. As the index for these materials is not zero, all are considered to be anisotropic in nature. Moreover, the elastic anisotropy is associated with the inter-atomic bond strength.
Compounds | Bv | BR | BH | GV | GR | GH | Y | ν | G/B | A1 | A2 | A3 | ξ |
---|---|---|---|---|---|---|---|---|---|---|---|---|---|
Zn3V2O8 | 110.45 | 109.5 | 109.98 | 45.83 | 37.3 | 41.57 | 110.69 | 0.33 | 0.378 | 1.31 | 1.60 | 0.31 | 0.44 |
Na3VO4 | 43.46 | 42.57 | 43.02 | 21.35 | 19.83 | 20.59 | 53.26 | 0.29 | 2.86 | 1.79 | — | 0.14 | 0.57 |
K3VO4 | 32.18 | 31.27 | 31.72 | 13.44 | 7.59 | 10.52 | 28.40 | 0.35 | 0.33 | 3.24 | — | 0.73 | 0.52 |
To define the relative positions of the cation and anion sub-lattices under the maintained volume-strain distortions (for which the positions are variable according to symmetry), Kleemann established a significant factor known as the internal strain parameter (ξ). It describes the coloration of the bond stretching and bending.44
The minimum bond bending is expressed when ξ = 0. In the case of ξ = 1, there will be minimum bond stretching. Table 5 illustrates that bond-bending is prominent in these photocatalytic materials.
(1) |
The most prominent estimated absorption apexes are 252269 cm−1, 177245 cm−1, and 152260 cm−1 for Zn3V2O8, Na3VO4 and K3VO4, respectively. The absorption values fluctuate with increasing photon energy. Meanwhile, optimum values were obtained at about 15.8 eV, 7.39 eV, and 26.9 eV for Zn3V2O8, Na3VO4 and K3VO4, respectively. The absorption cut-off band edge lies at 2.94 eV. This result indicates that the most likely electron transition to occur is VB to CB, as it has been studied for the BiVO4 photocatalyst.72 All these materials have almost zero absorption in the area of low-frequency photons. The photocatalytic degradation of pollutants is directly related to absorption.73 Photo-induced charge carriers (electron and hole pairs) are produced when light interacts and activates the photocatalytic material, causing the migration of photo-sensitive electrons from VB to CB. When these electron and hole pairs interact with water, it results in highly reactive hydroxyl and superoxide radicals. Therefore, these radicals interact with pollutants, and partially or completely degrade them.14 A lower band gap has a synergetic effect on enhancing the photocatalytic process by allowing the photocatalyst to absorb more photons, and become more sensitive to absorbing incident light.74 Hence, from Fig. 15a, Zn3V2O8 exhibits higher absorption than the other two photocatalyst materials.
(2) |
The high energy region and low energy region are two zones for describing the macroscopic and microscopic parameters of the materials.6 The critical onset values for the high energy region are 25 eV, 21 eV, and 27 eV for Zn3V2O8, Na3VO4 and K3VO4, respectively. As in the visible region, photocatalysts cannot attenuate incident light, which results in less plasmonic excitation. However, the composition and electronic structure can be designed via a low energy function region that is almost less than 1 eV, as shown in Fig. 15c.
ε(ω) = εRe(ω) + iεIm(ω) | (3) |
The initial threshold of the dielectric function's imaginary part is basically referred to as the first absorption peak. Such peaks correlate with the electronic transition between the occupied and unoccupied states near the Fermi level. The electron transition takes place between the V-3d and O-2p states. In addition to some other peaks, the critical onset points for the orthorhombic and tetragonal phases are 5.26 eV and 6.20 eV, respectively. By comparing the function of these three materials, the highest dielectric function of Zn3V2O8 confirms its excellent dielectric properties (Fig. 15d).
N(ω) = n(ω) + ik(ω) | (4) |
The variable n(ω) determines the extent of the electromagnetic wave slowdown in comparison to the vacuum speed. The real part of the complex function of such photocatalyst materials shows an inverse pattern. As depicted in Fig. 15g, n(ω) has a larger refractive index in the early stage of photon energy than k(ω), which is virtually zero. Therefore, the real and imaginary parts of the complex N(ω) function are slightly greater for Zn3V2O8 and K3VO4, as compared to Na3VO4.
Fig. 16 Calculated thermodynamic parameters, (a) Gibb's free energy, (b) enthalpy, (c) entropy, (d) heat capacity and (e) Debye temperature of Na3VO4, K3VO4 and Zn3V2O8. |
The Gibbs free energy (G) effectively defines the stability of a compound.81 Fig. 16a shows the variation in the Gibbs free energy of these systems with temperature. The decreasing trend of the Gibbs free energy for the systems is as follows: K3VO4 > Na3VO4 > Zn3V2O8. The Gibbs free energy corresponds to the amount of energy of the system that is required for work to be done. Thus, a system having a lower Gibbs energy (G) is thought to be thermodynamically favorable.82 The rising trend in the enthalpy and entropy curves with increasing temperature can also be observed in Fig. 16b and c, respectively. The behavior of these two factors is the opposite of that of the Gibbs free energy.83 Because it possesses the highest entropy, enthalpy, and lowest Gibbs energy, the Zn3V2O8 photocatalyst is thermodynamically more stable.
Solid-state physics also involves the investigation of a crystal's specific heat capacity at constant volume.84 The phonon thermal softening at high temperature is highlighted by the increasing trend of the Cv curves. When the temperature is low, Cv is proportional to T3. However, at elevated temperature, the heat capacity complies with the Dulong–Petit law, and reaches the asymptotic limit, i.e., 3nR. Here, n is the atom's number in the primitive cell, while R is the general gas constant.41 The highest heat capacity of Zn3V2O8 is 277.29 cal K−1 at 1000 K (Fig. 16d).
For qualitative insight into the atomic thermal vibrations, the Debye temperature (θD) was studied, as these two parameters are parallel (Fig. 16e). According to the Debye theory, the Debye temperature is compatible with the phonon frequency.85 Additionally, it represents the stability of the structure and the strength of the bonds, which directly corresponds to several physical characteristics, like the melting temperature and specific heat. The Debye temperature is derived from elastic constants, and observed heat measurements are equivalent at low temperatures. Computing θD via elastic constants is one of the standard methods of calculation. Thus, the Debye temperature (θD) can be estimated from the average sound velocity (νm), as described in the following equation:86
The transverse and longitudinal sound velocities (νt and νl) of a solid material can be determined from the Navier equation using the values of bulk and shear modulus.
The values of all these parameters are summarized in Table 6. To the best of our knowledge, there is no comparable experimental or theoretical data. Therefore, the stated findings are considered as the first-ever prediction. Thus, these results still need an experimental confirmation.
Crystal systems | V (Å3) | ρ (kg m−3) | νl (m s−1) | νt (m s−1) | νm (ms−1) | θD |
---|---|---|---|---|---|---|
Zv3V2O8 | 600.83 | 46500 | 6446 | 15548 | 7295.71 | 1520.52 |
Na3VO4 | 329.38 | 24600 | 2803 | 5348 | 3229.71 | 449.90 |
K3VO4 | 247.01 | 26100 | 2004 | 3891 | 2244.78 | 344.72 |
EVB = EVB + Eg |
Compounds | χ | Eg (eV) | CB (eV) | VB (eV) |
---|---|---|---|---|
Zn3V2O8 | 5.80 | 2.69 | −0.63 | 2.06 |
Zn3(1−0.06)V2(1−0.06)Fe0.06O8 | 5.90 | 1.72 | −0.36 | 1.36 |
Zn3(1−0.06)V2(1−0.06)Co0.06O8 | 5.93 | 1.68 | −0.31 | 1.37 |
Na3VO4 | 4.77 | 3.80 | −1.63 | 2.17 |
Na3V(1−0.06)Fe0.06O4 | 4.78 | 2.99 | −1.23 | 1.76 |
Na3V(1−0.06)Co0.06O4 | 4.78 | 2.59 | −1.02 | 1.57 |
K3VO4 | 4.50 | 3.65 | −1.83 | 1.82 |
K3V(1−0.06)Fe0.06O4 | 4.50 | 2.76 | −1.38 | 1.38 |
K3V(1−0.06)Co0.06O4 | 4.53 | 2.58 | −1.26 | 1.32 |
Fig. 17 Schematic with band edge positions of pure and doped Na3VO4, K3VO4, and Zn3V2O8 in the photocatalytic water splitting process. |
The following equation is used to calculate the electronegativity of a material:87
According to the equation, the generation of superoxide acid (−HO2) and superoxide radical (O2−˙) is an essential step for the solar degradation of pollutants. However, photogenerated holes can oxidize the OH− ion and fabricate the reactive hydroxyl radical (OH−), as shown in the chemical equations below:92
O2 + H+ + e− → HO2 |
O2 + e− → O2−˙ |
OH− + h+ → OH−˙ |
As discussed earlier, the conduction band potential for all of these systems is more negative as compared to the redox potential of O2/O−2˙ (−0.33 eV), indicating the reduction of O2 to the O2−˙ radical anion. Although the conduction band edge potential of the photocatalyst systems is more negative than that of O2/H2O2 (0.685 eV potential), this suggests that H2O2 is formed when electrons react with adsorbed oxygen molecules. This H2O2 combines with electrons to produce hydroxyl radicals. Furthermore, the valence band potential is more positive than that of HO−/HO˙ potential (1.99 eV), indicating that photogenerated holes in these systems oxidize hydroxyl ions to the reactive hydroxyl radicals required in the photocatalytic degradation of pollutants.93
All of the three host systems, either pure, Fe-doped or Co-doped photocatalysts, have a higher positive valence band potential than that of O2/H2O (1.23 eV), which is consistent with all of the aforementioned observations, and indicates a strong oxidation potential for the evolution of O2. Highly reactive radicals (OH−˙ and O2−˙) will oxidize both organic and inorganic substances, resulting in water and other non-toxic molecules.94 Interestingly, the CB positions of Na3VO4, Na3V0.94Fe0.06O4, Na3V0.94Co0.06O4, K3VO4, K3V0.94Fe0.06O4, K3V0.94Co0.06O4, Zn3V2O8, Zn2.82V1.88Fe0.12O4, and Zn2.82V1.88Co0.12O4 are negative as compared to the H2/H+ (0 eV) redox potential. A more positive potential provides the strong oxidizing potential for oxygen evolution in these photocatalysts, despite the fact that they are efficient at producing hydrogen gas. Higher positive VB and more negative CB potentials are among the best benchmarks for photocatalytic processes.95 The peculiar higher photocatalytic activity of Zn3V2O8 compared to other compounds can be ascribed to the location of the higher V-3d and Fe-3d or Co-3d states, compared with those of Na3VO4 and K3VO4. For efficient photocatalysis, electron–hole pairs should be energetically and geometrically separated.
Solubility is a key parameter in determining the photocatalytic efficiency. At present, most of the photocatalytic materials usually comprised insoluble semiconductors, while soluble semiconductor materials have restricted validity in photocatalysis.96 Soluble inorganics can dissolve in water due to the attraction of positive and negative charges, resulting in the disintegration of crystalline structures and the loss of energy band structures for light absorption.97 Meanwhile, the addition of an insoluble solute to its saturated solution will remain as crystals, and these undissolved photocatalytic materials will retain the geometrical shapes required for light absorption and photogenerated carrier transfer.96 Higher lattice energies are typically associated with less solubility because they demand more energy to break the ionic bonds within the lattice.98 Lattice energy minimization techniques have been used in attempt to enhance the effectiveness of water-soluble photocatalysts.99 The lattice energy of Zn3V2O8 (−50340.43 eV) is highest among all three studied orthovanadates, indicating its lowest solubility as compared to K3VO4 and Na3VO4 (Table 1). Moreover, the smaller size and greater charge on the Zn2+ ion than Na+ and K+ ions enables it to develop the strongest forces of attraction within the lattice, which are difficult to overcome, hence making it significantly less soluble in water. Furthermore, neither distortion of the crystal lattice nor significant reduction in the lattice energy was observed after doping. The solubility trend for all un-doped and doped orthovanadates is as follows: Zn3V2O8 > Zn3(1−0.06)V2(1−0.06)Co0.06O8 > Zn3(1−0.06)V2(1−0.06)Fe0.06O8 > Na3VO4 > Na3V(1−0.06)Co0.06O4 > Na3V(1−0.06)Fe0.06O4 > K3VO4 > K3V(1−0.06)Co0.06O4 > K3V(1−0.06)Fe0.06O4.
This journal is © The Royal Society of Chemistry 2024 |