Transport properties of Co-based Heusler compounds Co2VAl and Co2VGa: spin-polarized DFT+U

A. H. Reshak*ab
aNew Technologies-Research Centre, University of West Bohemia, Univerzitni 8, 306 14 Pilsen, Czech Republic. E-mail: maalidph@yahoo.co.uk; Fax: +420 386 361255; Tel: +420 777729583
bCenter of Excellence Geopolymer and Green Technology, School of Material Engineering, University Malaysia Perlis, 01007 Kangar, Perlis, Malaysia

Received 20th April 2016 , Accepted 18th May 2016

First published on 20th May 2016


Abstract

Transport properties of two Co-based Heusler compounds Co2VAl and Co2VGa were calculated. The calculated spin-polarized electronic band structures for majority and minority spin reveal that the minority spin exhibits an indirect band gap of about 0.3 eV for Co2VAl and 0.2 eV for Co2VGa. It is clear that the minority spin density of states at EF, N(EF) vanishes for Co2VAl and Co2VGa which leads to unusual transport properties because only the majority density contributes to the states at EF. The calculated valence band’s electronic charge density distribution reveals that there exists a strong covalent bond between the atoms which is more favorable for the transport of carriers than an ionic one. The transport properties were calculated as a function of temperature at a fixed chemical potential and as a function of chemical potential at three constant temperatures. In Co2VAl a large value (1100 μV K−1) of Seebeck coefficient is obtained for spin-down electrons due to the existence of an almost flat conduction band along the L to Γ direction. Calculations show that Co2VAl exhibits a higher Seebeck coefficient than that obtained from Co2VGa and the increase in the Seebeck coefficient also leads to a maximum in the power factor. This makes the Co2VAl and Co2VGa compounds attractive candidates for materials used in spin voltage generators.


1. Introduction

Recently, Heusler compounds had an astonishing comeback. Because of their high spin polarization and their high Curie temperatures they have received exceptional attention for possible applications in the field of spin-electronics, magneto-resistive devices (spintronics) and thermoelectric devices.1–3 Recent review articles cover the basic physics of half-metallic ferromagnets (HMF)1 and aspects of their applications.2,3 The properties of Heusler compounds are strongly dependent on the atomic order. Band structure calculations show that already small amounts of disorder within the distribution of the atoms on the lattice sites cause distinct changes in their electronic structure, and thus also in their magnetic and transport properties.4–6 Therefore, a careful analysis of their crystal structure is essential to understand the structure-to-property relation of Heusler compounds. The transition from the ordered to the most prominent disordered Heusler structures (X2YZ) will be explained as follows;7–11 if the Y and the Z atoms are evenly distributed, the 4a and 4b positions become equivalent. This leads to a CsCl-like structure, also known as B2-type disorder. The symmetry is reduced and the resulting space group is Pm[3 with combining macron]m. The random distribution of X and Y or X and Z leads to BiF3-type disorder (space group Fm[3 with combining macron]m). Already small amounts of anti-site disorder cause changes in the electronic structure close to the Fermi energy, which lead to altered physical properties. The existence of martensitic transition in Heusler alloys makes them potential candidates for actuation devices and smart materials.12–16

The recent observation of the spin Seebeck effect allows a pure spin current to pass over a long distance17 and is directly applicable to the production of spin voltage generators which are crucial for driving spintronics devices.18–20 Recently, half-metallic compounds have attracted much interest because of their exceptional band structures at the Fermi level (EF). From the electronic band structure point of view, half-metals have metallic character for one spin channel and semiconducting for the other.21 There exists a high spin-polarization at the EF and many Heusler alloys belong to this family.21 Therefore, we think it would be worthwhile to investigate the electronic structure and hence the transport properties of two Co-based Heusler compounds (i.e. Co2VAl and Co2VGa). The calculations have been performed in order to understand the properties of the band gap in the minority spin channel and the peculiar transport properties of the Co2VAl and Co2VGa compounds.

2. Details of calculations

The Co-based Heusler compounds Co2VAl and Co2VGa crystallize in cubic symmetry with the Fm[3 with combining macron]m space group.22,23 The experimental lattice constants are a = 5.770 Å (ref. 22 and 23) for Co2VAl and a = 5.786 Å (ref. 22 and 23) for Co2VGa. Using the generalized gradient approximation (PBE-GGA)24 we have optimized the lattice constants. These are a = 5.751 Å for Co2VAl and a = 5.763 Å for Co2VGa, which show good agreement with the experimental data.22,23 The full potential linear augmented plane wave plus the local orbitals (FPLAPW + lo) method as implemented in the WIEN2k code25 within PBE-GGA was used to perform comprehensive calculations for the Co2VAl and Co2VGa compounds. It is well-known that for oxides and other highly correlated compounds, local density approximation (LDA) and the generalized gradient approximation (GGA) are known to fail to give the correct ground state. In these systems, the electrons are highly localized. The Coulomb repulsion between the electrons in open shells should be taken into account. As there is no exchange correlation function that can include this in an orbital independent way, a simpler approach is to add the Hubbard-like on-site repulsion to the Kohn–Sham Hamiltonian. This is known as an LDA+U or GGA+U calculation. There are different ways in which this can be implemented. In the present work, we used the method of Anisimov et al.26 and Liechtenstein et al.27 where the Coulomb (U) and exchange (J) parameters are used. From the obtained relaxed geometry the ground state properties were determined using FPLAPW + lo25,28,29 within GGA+U (U-Hubbard Hamiltonian). We applied U on the 3d orbital of Co atoms and V atoms, and we have tested several U values until we reached the energy band gap values that agree well with the previous results.30 The U values used here are 0.15 Ryd and 0.187 Ryd (ref. 30) for Co and V, respectively. Based on the calculated electronic band structures we have calculated the transport properties of Co2VAl and Co2VGa utilizing the semi-classical Boltzmann theory as incorporated within the BoltzTraP code.31 Simulations of the transport properties are a transition from first- to second-principles methods. The first-principles method used here is the all-electron full potential linear augmented plane wave method whereas the second-principles method is the BoltzTraP code,31 which solves the semi-classical Bloch–Boltzmann transport equations.31 Transport properties were obtained from the ground state within the limits of the Boltzmann theory32–34 and the constant relaxation time approximation as implemented in the BoltzTraP code.31 In short, BoltzTraP performs a Fourier expansion of the quantum chemical band energies. This allows us to obtain the electronic group velocity v and inverse mass tensor, as the first and second derivatives of the bands with respect to k. Applying v to the semi-classical Boltzmann equations, the transport tensors can be evaluated.

The potential for the construction of the basis functions inside the sphere of the muffin-tin was spherically symmetric, whereas it was constant outside the sphere. Self-consistency is obtained using 800 image file: c6ra10226c-t1.tif points in the irreducible Brillouin zone (IBZ). The self-consistent calculations are converged since the total energy of the system is stable within 0.00001 Ryd. The spin-polarized electronic band structure and the transport properties were performed within 20[thin space (1/6-em)]000 image file: c6ra10226c-t2.tif points in the IBZ as the accurate calculations of transport properties of metals require dense sampling of the Brillouin zone. It is well known that first-principles calculations are a powerful and useful tool to predict the crystal structure and its properties related to the electron configuration of a material before its synthesis.35–37

3. Results and discussion

3.1. Salient features of the spin-polarized electronic band structure

The calculated spin-polarized electronic band structures for the spin-up (↑) majority spin and spin-down (↓) minority spin are shown in Fig. 1b, d, g and i. It has been found that the minority spin exhibits an indirect band gap of about 0.3 eV for Co2VAl and 0.2 eV for Co2VGa, as the valence band maximum (VBM) is located at the Γ point of the BZ and the conduction band minimum (CBM) at the X point of the BZ. The obtained values using GGA+U are much better than the previous calculated energy band gap of 0.238 eV (0.189 eV) of Co2VAl (Co2VGa) using LDA+U.38 Meanwhile, the majority spin exhibits a metallic structure with a density of states at EF, N(EF) of about 15.30 (state per eV per unit cell) for Co2VAl and 13.22 (state per eV per unit cell) for Co2VGa. The calculated density of states at EF enables us to calculate the bare electronic specific heat coefficient (γ) which is about 2.65 and 2.29 mJ per mol cell K2 for Co2VAl and Co2VGa. It is clear that the minority spin density of states at EF, N(EF) vanishes for Co2VAl and Co2VGa which should lead to unusual transport properties because only the majority density contributes to the states at EF. Similar behavior was observed in the Co2MnAl and Co2MnSn compounds.39 Therefore, the bands which cross EF are responsible for the transport properties of the compound and those bands which are not crossing EF will contribute negligibly to the transport properties.21 The calculated Fermi surface of the majority spin for Co2VAl and Co2VGa are shown in Fig. 1a and f. We noticed that in the Fermi surface there are white regions which represent the hole concentration while the colored regions correspond to the presence of electrons.40 The colors of the Fermi surface also give an idea about the speed of the electrons. The red color represents the highest speed, yellow, green and blue have intermediate speeds whereas the violet color shows the lowest speed. The colors of the Fermi surface confirm that replacing Al by Ga causes reduction/increase in the speed of the electrons at the Fermi surface. Usually the transport properties are related to the electrons in the system, and these electrons are defined through the Fermi surface, which determines the electrical conductivity.40
image file: c6ra10226c-f1.tif
Fig. 1 (a) Calculated Fermi surface of the Co2VAl compound. The white regions represent the hole concentration while the colored regions correspond to the presence of electrons. The red color represents the highest speed, yellow, green and blue have intermediate speeds whereas the violet color shows the lowest speed; (b) calculated spin-up electronic band structure of the Co2VAl compound; (c) calculated carrier concentration as a function of chemical potential (μEF = ±0.2 eV) at three constant temperatures (300, 600 and 900 K) for the spin-up configuration of the Co2VAl compound; (d) calculated spin-down electronic band structure of the Co2VAl compound; (e) calculated carrier concentration as a function of chemical potential (μEF = ±0.2 eV) at three constant temperatures (300, 600 and 900 K) for the spin-down configuration of the Co2VAl compound; (f) calculated Fermi surface of the Co2VGa compound. The white regions represent the hole concentration while the colored regions correspond to the presence of electrons. The red color represents the highest speed, yellow, green and blue have intermediate speeds whereas the violet color shows the lowest speed; (g) calculated spin-up electronic band structure of the Co2VGa compound; (h) calculated carrier concentration as a function of chemical potential (μEF = ±0.2 eV) at three constant temperatures (300, 600 and 900 K) for the spin-up configuration of the Co2VGa compound; (i) calculated spin-down electronic band structure of the Co2VGa compound; (j) calculated carrier concentration as a function of chemical potential (μEF = ±0.2 eV) at three constant temperatures (300, 600 and 900 K) for the spin-down configuration of the Co2VGa compound. We would like to mention that the colors in the electronic band structures mean nothing, they are just to illustrate the bands in a clear way.

Our results reveal that the Co2Val and Co2VGa compounds exhibit half-metallic (HM) nature. The electronic band structure in the area around the CBM and the VBM show that both CBM and VBM have parabolic bands in the vicinity of the Fermi level (Fig. 1b, d, g and i). This implies that Co2Val and Co2VGa show highest k-dispersion bands around EF and thus lowest effect masses and hence the highest mobility carriers.

To obtain further insight into the type of states associated with each orbital, the projected density of states (DOS) were calculated. The angular momentum character of various structures in the Co2VAl and Co2VGa compounds for the spin up (↑) and spin down (↓) states can be obtained by calculating the angular momentum projected density of states (PDOS), as shown in Fig. 2a–d. It is demonstrated that replacing Al by Ga introduces high peaks in the density of states below EF and that the charge density near the DOS peaks is substantially attracted toward Ga atoms due to their high electro-negativity.41 It is clear that the Co-s/p/d, V-s/p/d and Ga-s/p/d (Al-s/p) orbitals are distributed in the valence and conduction bands along the whole energy scale and this shows that there exists a strong hybridization between the states, for instance in the Co2VAl compound the Al-p state is hybridized with the Co-s/p and V-s/p states, Al-s with V-s and Co-s, and Co-d with V-d. Meanwhile, for the Co2VGa compound the Ga-p state hybridized with the Co-s/p and V-s states, Co-s/p with the Ga-s and V-p states, and Co-d with the V-d state. The hybridization degree favors enhancing the covalent bonding.42 To support this statement we have calculated the valence band electronic charge density distribution of the Co2VAl and Co2VGa compounds.


image file: c6ra10226c-f2.tif
Fig. 2 (a and b) The calculated angular momentum projected density of states (PDOS) of the Co2VAl compound for the spin up (↑) and spin down (↓) states; (c and d) the calculated angular momentum projected density of states (PDOS) of the Co2VGa compound for the spin up (↑) and spin down (↓) states.

3.2. Valence electronic charge density

To visualize the charge transfer and the chemical bonding character, the valence band electronic charge density distributions are calculated and analyzed in detail. Fig. 3a–i illustrates the calculated total valence charge density distribution in two crystallographic planes along the (1 0 0) and (1 0 1) directions for spin-up and spin-down cases. According to the Pauling scale, the electro-negativities of Co, V, Al and Ga atoms are 1.88, 1.63, 1.61 and 1.81, respectively. Thus the electro-negativity differences between the atoms are small, therefore the small electro-negativity differences indicate that there exist covalent bonds.43,44 Covalent bonding is more favorable for the transport of carriers than ionic bonding.45 Also, due to the electro-negativity differences between Co, V, Al and Ga atoms, some valence electrons are transferred towards Co and Ga atoms as it is clear that the Co and Ga atoms are surrounded by uniform blue spheres which indicate the maximum charge accumulation according to the thermoscale (Fig. 3i). We would like to highlight that the crystallographic plane along the (1 0 0) direction shows only Al(Ga) and V, whereas the crystallographic plane along the (1 0 1) direction shows all atoms. Therefore, the calculated valence band’s electronic charge density distribution helps us to gain deep insight into the nature of the chemical bonding, anisotropy in bonding and to explore the transport of the carriers.
image file: c6ra10226c-f3.tif
Fig. 3 The calculated valence electronic charge density in two crystallographic planes for spin-up/down configurations: (a and b) the calculated valence electronic charge density for the Co2VAl compound in the (1 0 0) crystallographic plane for spin-up and spin-down; (c and d) the calculated valence electronic charge density for the Co2VAl compound in the (1 0 1) crystallographic plane for spin-up and spin-down; (e and f) the calculated valence electronic charge density for the Co2VGa compound in the (1 0 0) crystallographic plane for spin-up and spin-down; (g and h) the calculated valence electronic charge density for the Co2VGa compound in the (1 0 1) crystallographic plane for spin-up and spin-down; (i) the thermoscale which shows the value of charge accumulation around the atoms, where the red color exhibits zero charge accumulation (0.0000) while the blue color indicates the maximum charge accumulation (+1.0000), the colors between red and blue represent the degree of charge accumulation between 0.0 and 1.0.

3.3. Transport properties

3.3.1. Charge carrier concentration and electrical conductivity. We have investigated the influence of temperature on the carrier concentration of the Heusler compounds Co2VAl and Co2VGa for the majority and minority spin at a certain value of the chemical potential (μ = EF). It has been found that the carrier concentration of the majority spin increases exponentially with increasing temperature as illustrated in Fig. 4a and b and the negative sign indicates that the materials exhibit n-type conduction, whereas for the minority spin the carrier concentration (Fig. 4c and d) shows a constant value until 400 K (450 K) for Co2VAl (Co2VGa), then increases dramatically with increasing temperature and the materials exhibit p-type conduction. The observed trends of the carrier concentration for the majority and minority spin confirm the half-metallic (HM) nature of Co2VAl and Co2VGa. To support this statement we have investigated the carrier concentration for (↑) and (↓) at three constant temperatures (300, 600 and 900 K) and μEF = ±0.2 eV in the vicinity of the Fermi level as shown in Fig. 1c, e, h and j. It has been noticed that the difference between chemical potential and Fermi energy (μEF) is positive for valence bands and negative for conduction bands. It is clear from the electronic band structure (Fig. 1b, d, g and i) that Co2VAl and Co2VGa have parabolic bands in the vicinity of the Fermi level, therefore the carriers exhibit low effective mass and hence high mobility. It has been found that the investigated materials exhibit a maximum carrier concentration in the vicinity of the Fermi level. Usually the temperature difference between two materials is responsible for producing a thermoelectric voltage. To gain high thermoelectric efficiency, it is necessary that the material possesses high electrical conductivity, a large Seebeck coefficient and low thermal conductivity.46 Therefore, to achieve the highest electrical conductivity, high mobility carriers are required. To achieve this, a material with small effective masses is needed. Therefore, we have calculated the effective mass of electrons (m*e) for spin-up and spin-down. Usually we estimated the value of effective mass of electrons from the conduction band minimum curvature. The diagonal elements of the effective mass tensor, m*e, for the electrons in the conduction band are calculated following this expression;
 
image file: c6ra10226c-t3.tif(1)

image file: c6ra10226c-f4.tif
Fig. 4 (a and b) The temperature induced carrier concentration per unit cell (e per uc) for spin-up electrons versus temperature for the Co2VAl and Co2VGa compounds; (c and d) the temperature induced carrier concentration per unit cell (e per uc) for spin-down electrons versus temperature for the Co2VAl and Co2VGa compounds.

The effective mass of electrons is assessed by fitting the electronic band structure to a parabolic function eqn (1). The calculated electron effective mass ratio (m*e/me) around the X point of the BZ is about 0.00846 (0.01075) for the spin-down configuration of Co2VAl (Co2VGa), while it is 0.01497 (0.02681) for the spin-up configuration of Co2VAl (Co2VGa). Therefore, the spin-up electrons possess higher mobility than those of spin-down. The effective mass ratio of the heavy holes (m*hh/me) around the Γ point of the BZ is about 0.03036 (0.02873) for Co2VAl (Co2VGa). This implies that heavy holes possess lower mobility than the electrons.

To ascertain that the investigated Heusler compounds Co2VAl and Co2VGa are expected to give maximum efficiency at EF, we have investigated the electrical conductivity for spin-up/down as a function of temperature at a certain value of chemical potential (μ = EF) as shown in Fig. 5a–d, which represents the temperature variation of σ/τ. Fig. 5a and b illustrate the electrical conductivity of the two compounds for spin-up. It has been noticed that at low temperature (<100 K) the electrical conductivity of Co2VAl reduced with increasing temperature, then afterwards a prompt increase occurred with rising temperature. Meanwhile, Co2VGa shows a sharp rise up to 100 K, then a significant change in the trend of σ/τ occurs to reach the saturated value at around 600 K that is attributed to the different bands which form the Fermi surface as shown in Fig. 1a and f. Meanwhile, for the spin-down configuration (Fig. 5c and d) the compounds have semiconductor like behavior; the electrical conductivity exhibits almost a constant value of about 1.11 × 1015 (3.4 × 1015) (Ω ms)−1 up to 400 (450) K for Co2VAl (Co2VGa). Above these temperatures a rapid increase occurs to reach the maximum value of about 4.14 × 1017 (1.30 × 1018) (Ω ms)−1 for Co2VAl (Co2VGa) at 900 K. At room temperature the value of σ/τ for spin-down electrons is very low in comparison to the spin-up electrons. Furthermore, we have calculated the total σ/τ. The total σ/τ is computed by using the two current model.47,48 It is clear that σ/τ (↑) for Co2VAl (Fig. 5a) shows a slight deviation from the linear temperature dependence, whereas total σ/τ (Fig. 5e) varies almost linearly with temperature. Therefore, at a certain value of chemical potential the carrier concentration and electrical conductivity are temperature-dependent.


image file: c6ra10226c-f5.tif
Fig. 5 (a and b) The electrical conductivity for spin-up electrons versus temperature for the Co2VAl and Co2VGa compounds; (c and d) the electrical conductivity for spin-down electrons versus temperature for the Co2VAl and Co2VGa compounds; (e and f) the total electrical conductivity versus temperature for the Co2VAl and Co2VGa compounds; (g and h) the electrical conductivity of Co2VAl and Co2VGa for spin-up electrons as a function of chemical potential (μEF = ±0.15 eV) at three constant temperatures (300, 600 and 900 K); (i and j) the electrical conductivity of Co2VAl and Co2VGa for spin-down electrons as a function of chemical potential (μEF = ±0.15 eV) at three constant temperatures (300, 600 and 900 K).

In addition to that, we have calculated σ/τ of Co2VAl and Co2VGa for spin-up/down electrons as a function of chemical potential (μEF = ±0.15 eV) at three constant temperatures (300, 600 and 900 K) as shown in Fig. 5g–j. It is clear that above EF a significant increase in the electrical conductivity occurs to reach a maximum value of about 1.0 × 1021 (0.92 × 1021) (Ω ms)−1 at μEF = +0.15 eV for spin-up electrons of Co2VAl (Co2VGa). Meanwhile, for spin-down electrons it is about 8.5 × 1020 (7.5 × 1020) (Ω ms)−1 for Co2VAl (Co2VGa) at μEF = +0.15 eV. It is clear that the temperature has an insignificant influence on the electrical conductivity in the chemical potential (μEF) range between −0.15 eV and +0.15 eV.

3.3.2. Electronic thermal conductivity. The thermal conductivity (k) consists of the electronic contribution ke (electrons and holes transporting heat) and the phonon contribution kl (phonons traveling through the lattice). The BoltzTraP code calculates only the electronic part ke.31 Using the BoltzTraP code we have calculated the electronic thermal conductivity of Co2VAl and Co2VGa for (↑) and (↓) as a function of temperature at the Fermi level (a certain value of chemical potential (μ = EF)) as shown in Fig. 6a–d which represents the variation in ke/τ at different temperatures. It is clear that for the spin-up configuration of Co2VAl and Co2VGa, ke/τ increases linearly with increasing temperature. This tendency is not valid for the spin-down configuration (Fig. 6c and d) since ke/τ exhibits a zero value up to 400 K for Co2VAl and 450 K for Co2VGa, then a rapid increase occurs with rising temperature. The results show that the electronic thermal conductivity is temperature-dependent.
image file: c6ra10226c-f6.tif
Fig. 6 (a and b) The electronic thermal conductivity for spin-up electrons versus temperature for the Co2VAl and Co2VGa compounds; (c and d) the electronic thermal conductivity for spin-down electrons versus temperature for the Co2VAl and Co2VGa compounds; (e and f) the electronic thermal conductivity of Co2VAl and Co2VGa for spin-up electrons as a function of chemical potential (μEF = ±0.15 eV) at three constant temperatures (300, 600 and 900 K); (g and h) the electronic thermal conductivity of Co2VAl and Co2VGa for spin-down electrons as a function of chemical potential (μEF = ±0.15 eV) at three constant temperatures (300, 600 and 900 K).

In addition to that, we have calculated ke/τ of Co2VAl and Co2VGa for spin-up/down electrons as a function of chemical potential (μEF = ±0.15 eV) at three constant temperatures (300, 600 and 900 K) as shown in Fig. 6e–h. It is clear that a significant increase in ke/τ occurs with increasing temperature and the temperature 900 K induced the highest ke/τ values, while 300 K induced the lowest ke/τ values, confirming that 300 K is the optimal temperature which gives the lowest ke/τ values in the chemical potential range μEF = +0.15 eV.

3.3.3. Seebeck coefficient. The Seebeck coefficient (S) is an important quantity which is related to the electronic band structure of the materials. We have calculated the Seebeck coefficients of Co2VAl and Co2VGa as a function of chemical potential (μEF = ±0.15 eV) at three constant temperatures (300, 600 and 900 K). In Fig. 7a–d we illustrate S along the chemical potential between 0.15 and −0.15 eV. One can see in the vicinity of EF that the Seebeck coefficient for the spin-up configuration exhibits several pronounced structures at EF, below and above EF. The differences in the structures’ height and location between the two compounds is attributed to the Fermi surface’s configuration. It has been found that for the spin-up case the Seebeck coefficient of Co2VGa at around μEF = +0.07 eV exhibits a higher value than that of Co2VAl. Fig. 7c and d illustrate the Seebeck coefficients of both compounds for the spin-down case. One can see in the vicinity of EF the Seebeck coefficients exhibit two pronounced structures just below and above EF, with the highest values of Seebeck coefficient at 300 K (1100.0 and −1000.0 μV K−1 for Co2VAl, and 1000.0 and −800.0 μV K−1 for Co2VGa). It has been noticed that in the Co2VAl compound a large value (1100 μV K−1) of Seebeck coefficient is obtained for spin-down electrons due to the existence of an almost flat conduction band along the L to Γ direction (see Fig. 1d). It has been found that for both compounds the obtained value of S for spin-down electrons is greater than that of spin-up electrons and the total S of Co2VAl is larger than that of Co2VGa. This is attributed to the fact that there are more valence electrons in Co2VGa than in Co2VAl. For an increase in the valence electrons the absolute value of the Seebeck coefficient decreases. This is explained by the increase in the electron concentration in the bands. By increasing the number of valence electrons, additional electrons are added to the d-band at the Fermi energy.49 This leads to an increase in the carrier concentration. The increase in the carrier concentration leads to a decrease in the Seebeck coefficient.46 The interrelationship between carrier concentration and Seebeck coefficient can be seen from relatively simple models of electron transport. For simple metals or degenerate semiconductors with parabolic bands and energy independent scattering, the Seebeck coefficient S is given by;
 
image file: c6ra10226c-t4.tif(2)
where n is the carrier concentration and m* is the effective mass of the carrier. It can be clearly seen that S depends on the carrier concentration and on the effective mass m*. The latter depends on the shape of the bands.

image file: c6ra10226c-f7.tif
Fig. 7 (a and b) The Seebeck coefficient of Co2VAl and Co2VGa for spin-up electrons as a function of chemical potential (μEF = ±0.15 eV) at three constant temperatures (300, 600 and 900 K); (c and d) the Seebeck coefficient of Co2VAl and Co2VGa for spin-down electrons as a function of chemical potential (μEF = ±0.15 eV) at three constant temperatures (300, 600 and 900 K).

From Fig. 7a–d it is clear that the obtained values of Seebeck coefficients are negative/positive for the entire range of the chemical potential suggesting the presence of n/p-type charge carriers. For spin-up electrons the temperature has a significant influence on S along the chemical potential range μEF = ±0.15 eV, whereas for the spin-down electrons the temperature shows a significant influence in the chemical potential range between +0.05 eV and −0.05 eV only.

From the above, we should emphasize that both materials posses n-/p-type conductions in the vicinity of EF. The sign of S indicates the type of dominant charge carriers: S with a positive sign represents the p-type materials, whereas n-type materials have a negative S.21,31,50,51

3.3.4. Power factor. The power factor is defined as P = S2σ/τ, where S2 is the Seebeck coefficient’s square, σ is the electrical conductivity and τ is the relaxation time. Following this formula one can see that P is directly proportional to S2 and σ/τ. Therefore, in order to gain a high power factor one needs to maintain the values of S2 and σ/τ. It is well known that the figure of merit is a very important quantity for calculating the transport properties of materials. The dimensionless figure of merit is written as ZT = S2σT/ke + kl (ref. 40 and 52), which shows that the power factor comes in the numerator of the figure of merit, thus the power factor is an important quantity and plays a principle role in evaluating the transport properties of the materials.

In order to ascertain the influence of varying the chemical potential on the power factor, we have calculated the power factor for spin-up and spin-down at three fixed temperatures (300, 600 and 900 K) as a function of chemical potential (μEF) between −0.15 and +0.15 eV as shown in Fig. 8a–d. It has been found that the investigated material exhibits the highest power factor just above and below EF. It is interesting to see that increasing the temperature causes significant increases in the power factor. The increase in the Seebeck coefficient also leads to a maximum of the power factor at μEF = ±0.05 eV for Co2VAl (Fig. 8c) and at −0.03 eV and 0.05 eV for Co2VGa (Fig. 8d). It is slightly shifted compared to the maximum of the Seebeck coefficient due to the decreased conductivity.


image file: c6ra10226c-f8.tif
Fig. 8 (a and b) The power factor of Co2VAl and Co2VGa for spin-up electrons as a function of chemical potential (μEF = ±0.15 eV) at three constant temperatures (300, 600 and 900 K); (c and d) the power factor of Co2VAl and Co2VGa for spin-down electrons as a function of chemical potential (μEF = ±0.15 eV) at three constant temperatures (300, 600 and 900 K).

We would like to mention here that in our previous works53–56 we have calculated the transport properties using the FPLAPW method within the BoltzTraP code on several systems whose transport properties are known experimentally, and in those previous calculations we found very good agreement with the experimental data. Thus, we believe that our calculations reported in this paper would produce very accurate and reliable results.

4. Conclusions

We have calculated the spin-polarized electronic band structure of Co2VAl and Co2VGa and hence the transport properties based on the semi-classical Boltzmann theory as incorporated within the BoltzTraP code. The calculated spin-polarized electronic band structure reveals that these compounds possess parabolic bands in the vicinity of the Fermi level, therefore the carriers exhibit low effective mass and hence high mobility. It has been found that the transport coefficients increase/reduce with increasing temperature, therefore the spin-up/down transport coefficients are temperature-dependent. The minority spin density of states at EF, N(EF) vanishes for Co2VAl and Co2VGa which should lead to unusual transport properties because only the majority density contributes to the states at EF. The calculated valence band’s electronic charge density distribution reveals that there exists a strong covalent bond between the atoms which is more favorable for the transport of carriers than an ionic one. Calculations show that Co2VAl exhibits a higher Seebeck coefficient than that obtained from Co2VGa. The increase in the Seebeck coefficient also leads to a maximum power factor. This makes them attractive candidates for materials used in spin voltage generators.

Acknowledgements

The result was developed within the CENTEM project, reg. no. CZ.1.05/2.1.00/03.0088, co-funded by the ERDF as part of the Ministry of Education, Youth and Sports OP RDI programme and, in the follow-up sustainability stage, supported through CENTEM PLUS (LO1402) by financial means from the Ministry of Education, Youth and Sports under the National Sustainability Programme I. Computational resources were provided by MetaCentrum (LM2010005) and CERIT-SC (CZ.1.05/3.2.00/08.0144) infrastructures.

References

  1. M. I. Katsnelson, V. Y. Irkhin, L. Chioncel and R. A. de Groot, Rev. Mod. Phys., 2008, 80, 315 CrossRef CAS.
  2. C. Felser, G. H. Fecher and B. Balke, Angew. Chem., Int. Ed., 2007, 46, 668 CrossRef CAS PubMed.
  3. B. Balke, S. Wurmehl, G. H. Fecher, C. Felser and J. Kübler, Sci. Technol. Adv. Mater., 2008, 9, 014102 CrossRef.
  4. Y. Miura, K. Nagao and M. Shirai, Phys. Rev. B: Condens. Matter Mater. Phys., 2004, 69, 144413 CrossRef.
  5. H. C. Kandpal, V. Ksenofontov, M. Wojcik, R. Seshadri and C. Felser, J. Phys. D: Appl. Phys., 2007, 40, 1587 CrossRef CAS.
  6. S. Picozzi, A. Continenza and A. J. Freeman, Phys. Rev. B: Condens. Matter Mater. Phys., 2004, 69, 094423 CrossRef.
  7. P. J. Webster, Contemp. Phys., 1969, 10, 559 CrossRef CAS.
  8. G. E. Bacon and J. S. Plant, J. Phys. F: Met. Phys., 1971, 1, 524 CAS.
  9. P. J. Webster and K. R. A. Ziebeck, Landolt-Börnstein new series group III, 1988, vol. 19c, p. 75 Search PubMed.
  10. K. R. A. Ziebeck and K. U. Neumann, Landolt-Börnstein new series group III, 2001, vol. 32c, p. 64 Search PubMed.
  11. T. Graf, F. Casper, J. Winterlik, B. Balke, G. H. Fecher and C. Felser, Z. Anorg. Allg. Chem., 2009, 635, 976 CrossRef CAS.
  12. A. Sozinov, A. A. Likhachev, N. Lanska and K. Ullako, Appl. Phys. Lett., 2002, 80, 1746 CrossRef CAS.
  13. A. Planes and J. Ortin, J. Appl. Phys., 1992, 71, 950 CrossRef.
  14. R. Kainuma, H. Nakano, K. Oikawa, K. Ishida and T. Nishizawa, Mater. Res. Soc. Symp. Proc., 1992, 246, 403 CrossRef CAS.
  15. R. Kainuma, N. Ono and K. Ishida, Mater. Res. Soc. Symp. Proc., 1995, 360, 467 CrossRef CAS.
  16. R. Kainuma, H. Nakano and K. Ishida, Metall. Mater. Trans. A, 1996, 27, 4153 CrossRef.
  17. K. Uchida, S. Takahashi, K. Harii, J. Ieda, W. Koshibae, S. Ando, K. Maekawa and E. Saitoh, Nature, 2008, 455, 778 CrossRef CAS PubMed.
  18. S. A. Wolf, D. D. Awschalom, R. A. Buhrman, J. M. Daughton, S. von Molnar, M. L. Roukes, A. Y. Chtchelkanova and D. M. Treger, Science, 2001, 294, 1488 CrossRef CAS PubMed.
  19. I. Zutic, J. Fabian and S. D. Sarma, Rev. Mod. Phys., 2004, 76, 323 CrossRef CAS.
  20. C. Chappert, A. Fert and F. N. Van Dau, Nat. Mater., 2007, 6, 813 CrossRef CAS PubMed.
  21. S. Sharma and S. K. Pandey, J. Phys.: Condens. Matter, 2014, 26, 215501 CrossRef CAS PubMed.
  22. P. J. Webster and K. R. A. Ziebeck, J. Phys. Chem. Solids, 1973, 34, 1647 CrossRef CAS.
  23. P. J. Webster and K. R. A. Ziebeck, Alloys and compounds of d elements with Main Group Elements, Part 2, ed. H. R. J. Wijn and Landolt-Bornstein, Springer Berlin, New Series, Group III, 1998, 2001, vol. 19/c, pp. 75–184 Search PubMed.
  24. J. P. Perdew, S. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996, 77, 3865 CrossRef CAS PubMed.
  25. P. Blaha, K. Schwarz, G. K. H. Madsen, D. Kvasnicka and J. Luitz, WIEN2k, An augmented plane wave plus local orbitals program for calculating crystal properties, Vienna University of Technology, Austria, 2001 Search PubMed.
  26. V. I. Anisimov, I. V. Solvyev, M. A. Korotin, M. T. Czyzyk and C. A. Sawatzky, Phys. Rev. B: Condens. Matter Mater. Phys., 1993, 48, 16929 CrossRef CAS.
  27. A. I. Liechtenstein, V. I. Anisimov and J. Zaanen, Phys. Rev. B: Condens. Matter Mater. Phys., 1995, 52, R5467 CrossRef CAS.
  28. O. K. Andersen, Phys. Rev. B: Condens. Matter Mater. Phys., 1975, 12, 3060 CrossRef CAS.
  29. J. P. Perdew and Y. Wang, Phys. Rev. B: Condens. Matter Mater. Phys., 1992, 45, 13244 CrossRef.
  30. D. P. Rai, Sandeep, A. Shankar, M. P. Ghimire and R. K. Thapa, Phys. Scr., 2012, 86, 045702 CrossRef.
  31. G. K. H. Madsen and D. J. Singh, Comput. Phys. Commun., 2006, 175, 67–71 CrossRef CAS.
  32. P. B. Allen, in Quantum Theory of Real Materials, ed. J. R. Chelikowsky and S. G. Louie, Kluwer, Boston, 1996, pp. 219–250 Search PubMed.
  33. J. M. Ziman, Electrons and Phonons, Clarendon, Oxford, 2001 Search PubMed.
  34. C. M. Hurd, The Hall Effect in Metals and Alloys, Plenum, New York, 1972 Search PubMed.
  35. K. J. Plucinski, I. V. Kityk, J. Kasperczyk and B. Sahraoui, Semicond. Sci. Technol., 2001, 16, 467 CrossRef CAS.
  36. H. S. Saini, M. Singh, A. H. Reshak and M. K. Kashyap, J. Magn. Magn. Mater., 2013, 331, 1–6 CrossRef CAS.
  37. Y. Saeed, S. Nazir, A. Shaukat and A. H. Reshak, J. Magn. Magn. Mater., 2010, 322, 3214–3222 CrossRef CAS.
  38. C. Felser and G. H. Fecher, Spintronics From Materials to Devices, Springer Dordrecht Heidelberg, New York London, eBook, 2013,  DOI:10.1007/978-90-481-3832-6, ISBN 978-90-481-3831-9, ISBN 978-90-481-3832-6.
  39. J. Kübler, A. R. Williams and C. B. Sommers, Phys. Rev. B: Condens. Matter Mater. Phys., 1983, 28, 1745 CrossRef.
  40. A. H. Reshak and S. Azam, J. Magn. Magn. Mater., 2013, 342, 80–86 CrossRef CAS.
  41. J.-H. Lee, G. Wu and J. C. Grossman, Phys. Rev. Lett., 2010, 104, 016602 CrossRef PubMed.
  42. A. H. Reshak, Z. A. Alahmed, J. Bila, V. V. Atuchin, B. G. Bazarov, O. D. Chimitova, M. S. Molokeev, I. P. Prosvirin and A. P. Yelisseyev, J. Phys. Chem. C, 2016, 120, 10559–10568 CAS.
  43. Schlusseltechnologien Key Technologies, 41st IFF Springschool, 2010, pp. A1–18 Search PubMed.
  44. S. Khan and A. H. Reshak, Polyhedron, 2015, 85, 962–970 CrossRef CAS.
  45. F. Wu, H. Z. Song, J. F. Jia and X. Hu, Prog. Nat. Sci., 2013, 23(4), 408–412 CrossRef.
  46. G. J. Snyder and E. S. Toberer, Nat. Mater., 2008, 7, 105–114 CrossRef CAS PubMed.
  47. H. J. Xiang and D. J. Singh, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 76, 195111 CrossRef.
  48. A. S. Botana, P. M. Botta, C. D. L. Calle, A. Pineiro, V. Pardo, D. Baldomir and J. A. Alonso, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 83, 184420 CrossRef.
  49. S. Ouardi, B. Balke, A. Gloskovskii, G. H. Fecher, C. Felser, G. Schönhense, T. Ishikawa, T. Uemura, M. Yamamoto, H. Sukegawa, W. Wang, K. Inomata, Y. Yamashita, H. Yoshikawa, S. Ueda and K. Kobayashi, J. Phys. D: Appl. Phys., 2009, 42, 084010 CrossRef.
  50. B. Xu, X. Li, G. Yu, J. Zhang, S. Ma, Y. Wang and L. Yi, J. Alloys Compd., 2013, 565, 22–28 CrossRef CAS.
  51. T. J. Scheidemantel, C. Ambrosch-Draxl, T. Thonhauser, J. V. Badding and J. O. Sofo, Phys. Rev. B: Condens. Matter Mater. Phys., 2003, 68, 125210 CrossRef.
  52. C. Calvin Hu, Modern Semiconductor Devices for Integrated Circuits, Part I: Electrons and holes in a semiconductor, November 11, 2011 Search PubMed.
  53. A. H. Reshak and S. Auluck, Comput. Mater. Sci., 2015, 96, 90–95 CrossRef CAS.
  54. A. H. Reshak, J. Phys. Chem. Solids, 2015, 78, 46–52 CrossRef CAS.
  55. A. H. Reshak, Renewable Energy, 2015, 76, 36–44 CrossRef CAS.
  56. A. H. Reshak, RSC Adv., 2014, 4, 63137 RSC; A. H. Reshak, RSC Adv., 2015, 5, 47569 RSC.

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.