首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The Curie temperature (Tc) and magnetic entropy change (−ΔSm), and their relationship to the alloy composition of Tb–Co metallic glasses, were studied systematically in this paper. It was found that, in contrast to the situation in amorphous Gd–Co ribbons, the dependence of Tc on Tb content and the maximum −ΔSm vs. Tc -2/3 plots in Tb–Co binary amorphous alloys do not follow a linear relationship, both of which are supposed to be closely related to the non-linear compositional dependence of Tb–Co interaction due to the existence of orbital momentum in Tb.  相似文献   

2.
Solid-state reaction was used for Li7La3Zr2O12 material synthesis from Li2CO3, La2O3 and ZrO2 powders. Phase investigation of Li7La3Zr2O12 was carried out by x-ray diffraction (XRD), scanning electron microscopy (SEM) and energy-dispersive x-ray spectroscopy (EDS) methods. The thermodynamic characteristics were investigated by calorimetry measurements. The molar heat capacity (Cp,m), the standard enthalpy of formation from binary compounds (ΔoxHLLZO) and from elements (ΔfHLLZO), entropy (S0298), the Gibbs free energy of the Li7La3Zr2O12 formation (∆f G0298) and the Gibbs free energy of the LLZO reaction with metallic Li (∆rGLLZO/Li) were determined. The corresponding values are Cp,m = 518.135 + 0.599 × T − 8.339 × T−2, (temperature range is 298–800 K), ΔoxHLLZO = −186.4 kJ·mol−1, ΔfHLLZO = −9327.65 ± 7.9 kJ·mol−1, S0298 = 362.3 J·mol−1·K−1, ∆f G0298 = −9435.6 kJ·mol−1, and ∆rGLLZO/Li = 8.2 kJ·mol−1, respectively. Thermodynamic performance shows the possibility of Li7La3Zr2O12 usage in lithium-ion batteries.  相似文献   

3.
The microstructure, revealed by X-ray diffraction and transmission Mössbauer spectroscopy, magnetization versus temperature, external magnetizing field induction and mechanical hardness of the as-quenched Fe75Zr4Ti3Cu1B17 amorphous alloy with two refractory metals (Zr, Ti) have been measured. The X-ray diffraction is consistent with the Mössbauer spectra and is characteristic of a single-phase amorphous ferromagnet. The Curie point of the alloy is about 455 K, and the peak value of the isothermal magnetic entropy change, derived from the magnetization versus external magnetizing field induction curves, equals 1.7 J·kg−1·K−1. The refrigerant capacity of this alloy exhibits the linear dependence on the maximum magnetizing induction (Bm) and reaches a value of 110 J·kg−1 at Bm = 2 T. The average value of the instrumental hardness (HVIT) is about 14.5 GPa and is superior to other crystalline Fe-based metallic materials measured under the same conditions. HVIT does not change drastically, and the only statistically acceptable changes are visibly proving the single-phase character of the material.  相似文献   

4.
The effect of substitution of Fe by Cu on the crystal structure and magnetic properties of Fe72−xNi8Nb4CuxSi2B14 alloys (x = 0.6, 1.1, 1.6 at.%) in the form of ribbons was investigated. The chemical composition of the materials was established on the basis of the calculated minima of thermodynamic parameters: Gibbs free energy of amorphous phase formation ΔGamorph (minimum at 0.6 at.% of Cu) and Gibbs free energy of mixing ΔGmix (minimum at 1.6 at.% of Cu). The characteristic crystallization temperatures Tx1onset and Tx1 of the alpha-iron phase together with the activation energy Ea for the as-spun samples were determined by differential scanning calorimetry (DSC) with a heating rate of 10–100 °C/min. In order to determine the optimal soft magnetic properties, the wound cores were subjected to a controlled isothermal annealing process in the temperature range of 340–640 °C for 20 min. Coercivity Hc, saturation induction Bs and core power losses at B = 1 T and frequency f = 50 Hz P10/50 were determined for all samples. Moreover, for the samples with the lowest Hc and P10/50, the magnetic losses were determined in a wider frequency range 50 Hz–400 kHz. The real and imaginary parts of the magnetic permeability µ′, µ″ along with the cut-off frequency were determined for the samples annealed at 360, 460, and 560 °C. The best soft magnetic properties (i.e., the lowest value of Hc and P10/50) were observed for samples annealed at 460 °C, with Hc = 4.88–5.69 A/m, Bs = 1.18–1.24 T, P10/50 = 0.072–0.084 W/kg, µ′ = 8350–10,630 and cutoff frequency at 8–9.3 × 104 Hz. The structural study of as-spun and annealed ribbons was carried out using X-ray diffraction (XRD) and a transmission electron microscope (TEM).  相似文献   

5.
The aim of the paper was to study the structure, magnetic properties and critical behavior of the Fe60Co12Gd4Mo3B21 alloy. The X-ray diffractometry and the Mössbauer spectroscopy studies confirmed amorphous structure. The analysis of temperature evolution of the exponent n (ΔSM = C·(Bmax)n) and the Arrott plots showed the second order phase transition in investigated material. The analysis of critical behavior was carried out in order to reveal the critical exponents and precise TC value. The ascertained critical exponents were used to determine the theoretical value of the exponent n, which corresponded well with experimental results.  相似文献   

6.
Ferroelectric property that induces electrocaloric effect was investigated in Ba(GexTi1−x)O3 ceramics, known as BTGx. X-ray diffraction analysis shows pure perovskite phases in tetragonal symmetry compatible with the P4mm (No. 99) space group. Dielectric permittivity exhibits first-order ferroelectric-paraelectric phase transition, confirmed by specific heat measurements, similar to that observed in BaTiO3 (BTO) crystal. Curie temperature varies weakly as a function of Ge-content. Using the direct and indirect method, we confirmed that the adiabatic temperature change ΔT reached its higher value of 0.9 K under 8 kV/cm for the composition BTG6, corresponding to an electrocaloric responsivity ΔT/ΔE of 1.13 × 10−6 K.m/V. Such electrocaloric responsivity significantly exceeds those obtained so far in other barium titanate-based lead-free electrocaloric ceramic materials. Energy storage investigations show promising results: stored energy density of ~17 mJ/cm3 and an energy efficiency of ~88% in the composition BTG5. These results classify the studied materials as candidates for cooling devices and energy storage applications.  相似文献   

7.
The arc-melting method was adopted to prepare the compound La0.5Pr0.5(Fe1−xCox)11.4Si1.6 (x = 0, 0.02, 0.04, 0.06, 0.08), and the magnetocaloric effect of the compound was investigated. As indicated by the powder X-ray diffraction (XRD) results, after receiving 7-day high temperature annealing at 1373 K, all the compounds formed a single-phase cubic NaZn13 crystal structure. As indicated by the magnetic measurement, the most significant magnetic entropy change |∆SM(T)| of the sample decreased from 28.92 J/kg·K to 4.22 J/kg·K with the increase of the Co content under the 0–1.5 T magnetic field, while the Curie temperature TC increased from 185 K to the room temperature 296 K, which indicated that this series of alloys are the room temperature magnetic refrigerant material with practical value. By using the ferromagnetic Curie temperature theory and analyzing the effect of Co doping on the exchange integral of these alloys, the mechanism that the Curie temperature of La0.5Pr0.5(Fe1−xCox)11.4Si1.6 and La0.8Ce0.2(Fe1−xCox)11.4Si1.6 increased with the increase in the Co content was reasonably explained. Accordingly, this paper can provide a theoretical reference for subsequent studies.  相似文献   

8.
The effects of the substitution of Fe by Co or Ni on both the structure and the magnetic properties of FeB amorphous alloy were investigated using first-principle molecular dynamics. The pair distribution function, Voronoi polyhedra, and density of states of Fe80−xTMxB20 (x = 0, 10, 20, 30, and 40 at.%, TM(Transition Metal): Co, Ni) amorphous alloys were calculated. The results show that with the increase in Co content, the saturation magnetization of Fe80−xCoxB20 (x = 0, 10, 20, 30, and 40 at.%) amorphous alloys initially increases and then decreases upon reaching the maximum at x = 10 at.%, while for Fe80−xNixB20 (x = 0, 10, 20, 30, and 40 at.%), the saturation magnetization decreases monotonously with the increase in Ni content. Accordingly, for the two kinds of amorphous alloys, the obtained simulation results on the variation trends of the saturation magnetization with the change in alloy composition are in good agreement with the experimental observation. Furthermore, the relative maximum magnetic moment was recorded for Fe70Co10B20 amorphous alloy, due to the induced increased magnetic moments of the Fe atoms surrounding the Co atom in the case of low Co dopant, as well as the increase in the exchange splitting energy caused by the enhancement of local atomic symmetry.  相似文献   

9.
In this work, the stability of Sr2(FeMo)O6−δ-type perovskites was tailored by the substitution of Mo with Ti. Redox stable Sr2Fe1.4TixMo0.6−xO6−δ (x = 0.1, 0.2 and 0.3) perovskites were successfully obtained and evaluated as potential electrode materials for SOFCs. The crystal structure as a function of temperature, microstructure, redox stability, and thermal expansion properties in reducing and oxidizing atmospheres, oxygen content change, and transport properties in air and reducing conditions, as well as chemical stability and compatibility towards typical electrolytes have been systematically studied. All Sr2Fe1.4TixMo0.6−xO6−δ compounds exhibit a regular crystal structure with Pm-3m space group, showing excellent stability in oxidizing and reducing conditions. The increase of Ti-doping content in materials increases the thermal expansion coefficient (TEC), oxygen content change, and electrical conductivity in air, while it decreases the conductivity in reducing condition. All three materials are stable and compatible with studied electrolytes. Interestingly, redox stable Sr2Fe1.4Ti0.1Mo0.5O6−δ, possessing 1 μm grain size, low TEC (15.3 × 10−6 K−1), large oxygen content change of 0.72 mol·mol−1 between 30 and 900 °C, satisfactory conductivity of 4.1–7.3 S·cm−1 in 5% H2 at 600–800 °C, and good transport coefficients D and k, could be considered as a potential anode material for SOFCs, and are thus of great interest for further studies.  相似文献   

10.
Li3FeN2 material was synthesized by the two-step solid-state method from Li3N (adiabatic camera) and FeN2 (tube furnace) powders. Phase investigation of Li3N, FeN2, and Li3FeN2 was carried out. The discharge capacity of Li3FeN2 is 343 mAh g−1, which is about 44.7% of the theoretic capacity. The ternary nitride Li3FeN2 molar heat capacity is calculated using the formula Cp,m = 77.831 + 0.130 × T − 6289 × T−2, (T is absolute temperature, temperature range is 298–900 K, pressure is constant). The thermodynamic characteristics of Li3FeN2 have the following values: entropy S0298 = 116.2 J mol−1 K−1, molar enthalpy of dissolution ΔdHLFN = −206.537 ± 2.8 kJ mol−1, the standard enthalpy of formation ΔfH0 = −291.331 ± 5.7 kJ mol−1, entropy S0298 = 113.2 J mol−1 K−1 (Neumann–Kopp rule) and 116.2 J mol−1 K−1 (W. Herz rule), the standard Gibbs free energy of formation ΔfG0298 = −276.7 kJ mol−1.  相似文献   

11.
The attractive/repulsive relationship between superconductivity and magnetic ordering has fascinated the condensed matter physics community for a century. In the early days, magnetic impurities doped into a superconductor were found to quickly suppress superconductivity. Later, a variety of systems, such as cuprates, heavy fermions, and Fe pnictides, showed superconductivity in a narrow region near the border to antiferromagnetism (AFM) as a function of pressure or doping. However, the coexistence of superconductivity and ferromagnetic (FM) or AFM ordering is found in a few compounds [RRh4B4 (R = Nd, Sm, Tm, Er), R′Mo6X8 (R′ = Tb, Dy, Er, Ho, and X = S, Se), UMGe (M = Ge, Rh, Co), CeCoIn5, EuFe2(As1−xPx)2, etc.], providing evidence for their compatibility. Here, we present a third situation, where superconductivity coexists with FM and near the border of AFM in Fe1−xPdxTe. The doping of Pd for Fe gradually suppresses the first-order AFM ordering at temperature TN/S, and turns into short-range AFM correlation with a characteristic peak in magnetic susceptibility at TN. Superconductivity sets in when TN reaches zero. However, there is a gigantic ferromagnetic dome imposed in the superconducting-AFM (short-range) cross-over regime. Such a system is ideal for studying the interplay between superconductivity and two types of magnetic (FM and AFM) interactions.Since the first discovery of superconductivity (SC) a century ago, the effects of magnetic impurities and the possibility of magnetic ordering in superconductors has been a central topic of condensed matter physics. Due to strong spin scattering (1, 2), it has generally been believed that the conduction electrons cannot be both magnetically ordered and superconducting. Even though it is thought that Cooper pairs in cuprates, heavy fermions, and Fe-based compounds are mediated by spin fluctuations (35), SC generally occurs after suppressing the magnetic ordering either through chemical doping or the application of hydrostatic pressure (610). However, there is growing evidence for the coexistence of superconductivity with either ferromagnetic (FM) (1120) or antiferromagnetic (AFM) ordering (2124). With the decrease of temperature (T), some of these systems show magnetic ordering before the superconducting transition (Tc) (1417, 20), some are ordered in a reversed sequence (1113, 18, 19, 22), some have the two orderings occur concomitantly (22, 25), and some show reentrant superconductivity (partially) overlapping with a magnetically ordered phase (1113, 26). Despite extensive investigations of interaction between SC and magnetic moments, there is so far no unified theory for the coexistence of SC and magnetism. With the lack of theoretic guidance, the existing experimental findings lead to two schools of thought: one is that both orders result from the same conduction electrons as evidenced by their synchronized magnetic and superconducting orders (22), and the other is that there are two separate sets of electrons responsible for magnetic ordering and superconductivity, respectively (19, 21, 25). What remains incomprehensible is the case where superconductivity and magnetic ordering coexist but are in competition with each other, as seen in Fe-based systems (23, 24, 27).The discovery of superconductivity in Fe-based compounds has sparked enormous interest in the scientific community. Although Fe is the most well known ferromagnet, all parent compounds of Fe-based superconductors exhibit AFM ordering. Though superconductivity is induced after suppressing the AFM ordering, it can coexist with either remaining AFM ordering (23, 24, 27) or new FM ordering (18, 19), and this provides an ideal platform for studying the interplay between superconductivity and magnetism. Among the Fe-based superconductors, the chalcogenide FeTe1−xSex is unique in several aspects: (i) it is the only compound composed of slabs of Fe(Te/Se)4 stacked together without an interlayer spacer; (ii) it becomes superconducting via isovalent doping of Se for Te, with the highest Tc occurring at the 50% doping level; and (iii) the parent compound Fe1+yTe shows nonmetallic electrical conduction and forms (π, π)-type AFM ordering with a large magnetic moment (28, 29), in contrast to the (π, 0)-type ordering with a small magnetic moment seen in Fe pnictides. The unusual AFM order in Fe1+yTe cannot be explained by a simple Fermi-surface nesting picture, thus leading to arguments for a correlated local-moment scenario (27). Because of these differences, the application of hydrostatic pressure or partial doping on the Fe site, such as with Co or Ni, does not generate the generic phase diagram as seen for other Fe-based superconductors (3033).To gain insight into the relationship between magnetism and superconductivity, we choose to substitute Pd for Fe in FeTe. This decision is motivated by the following: (i) Pd is known to exhibit FM instability (34, 35) even though it is paramagnetic in bulk, and (ii) PdTe is a superconductor with Tc ∼4.5 K (3642). The partial replacement of Fe by Pd will allow us to understand the roles of Fe, Pd, and Te in both magnetism and superconductivity. Based on electrical transport, and magnetic and thermodynamic property measurements, we show that the ground state of Fe1−xPdxTe varies from AFM to FM ordering, to a superconducting state; this is one of rare cases where superconductivity flirts with both AFM and FM.Single crystals of Fe1−xPdxTe were grown via high-temperature melting, with the procedure described in SI Text. The crystal structure and the phase purity were measured by both powder and single-crystal X-ray diffraction. Fig. 1 shows powder X-ray diffraction patterns (Fig. 1A) and lattice parameters obtained from single-crystal X-ray refinement (Fig. 1B) for different doping levels; it confirms that the undoped FeTe forms a tetragonal structure, belonging to the P4/nmm space group. At room temperature, the lattice parameters are a = 3.8202 Å and c = 6.2686 Å. Similar to previous observations (43), our single-crystal refinement result indicates that there are ∼9% extra Fe atoms (T2 site of Fig. 1B) incorporated at interstitial sites of the Te layers. As soon as Pd is introduced into the system, the interstitial sites of Fe1−xPdxTe are occupied by Pd atoms (i.e., T2 = Pd when x > 0); though it remains in the same tetragonal space group, its lattice parameters a and c both increase with increasing Pd doping concentration for xxs ∼0.6 (Fig. 1B). This finding indicates that Pd doping creates negative pressure, which is surprising because Pd2+ (∼0.80 Å) has a nearly identical ionic radius as that of Fe2+ (∼0.77 Å). The crystal structure of Fe1−xPdxTe remains tetragonal up to xs (∼0.6). Above xs, Fe1−xPdxTe crystallizes in a hexagonal structure (space group P63/mmc), and its lattice parameters also increase with increasing x (Fig. 1B); this results in an increase of the unit cell volume with increasing x in the entire doping range (Fig. 1B). In the hexagonal structure, the system no longer incorporates any interstitial sites (Tables S1S4). Though the hexagonal structure at x > xs can be regarded as a deformed FeTe structure, the local environment of Fe/Pd is transformed from a tetrahedron (x < xs) to an octahedron (x > xs) (44).Open in a separate windowFig. 1.Fe1−xPdxTe: (A) Powder X-ray diffraction patterns and (B) its unit cell parameters at room temperature obtained from single-crystal X-ray refinement.To correlate the structural information with physical properties, we show, in Fig. 2, the doping dependence of structural, electrical, and magnetic properties of Fe1−xPdxTe, with specifics presented later. The undoped (x = 0) Fe1.09Te undergoes both an AFM ordering and a structural transition from paramagnetic (PM) tetragonal (high temperatures) to an AFM-ordered monoclinic (low temperatures) phase at TN/S ∼70 K, with the first-order characteristic. Upon Pd doping, TN/S is suppressed (solid circles), reaching zero at x = xN/S ∼0.15 (dash line). In this doping region, the electrical resistivity (ρ) changes from nonmetallic (NM) character (dρ/dT < 0) above TN/S to metallic (M) behavior (dρ/dT > 0) below TN/S. When the characteristic feature for the first-order transition is absent at xN/S, the magnetic susceptibility of Fe1−xPdxTe exhibits a peak at TN. The value of TN decreases with increasing x (Fig. 2, open circles), which eventually reaches zero at xc ∼0.88. The magnetic susceptibility peak is a signature for short-range (SR) AFM correlation developed below TN, with no obvious fingerprint in the electrical resistivity. Above xc, superconductivity appears (Fig. 2, solid diamonds). The transition temperature Tc increases with increasing x. Remarkably, the system shows ferromagnetism (Fig. 2, solid squares) in the doping range of 0.78 ≤ x ≤ 0.98 with the maximum TFM (∼190 K) at xc, where AFM and SC vanish. In addition to the variable magnetic phases, there is doping-induced structural phase transition, being tetragonal at x < xs ∼0.6 (Fig. 2, dark blue) and hexagonal at x > xs (light blue). Such a structural transition is responsible for a change of electronic structure, manifested by NM behavior at x < xs and a metallic character at x > xs.Open in a separate windowFig. 2.Phase diagram of Fe1−xPdxTe. AFM, antiferromagnetic; AFM-M, antiferromagnetically ordered metallic state; FM, ferromagnetically ordered state; M, metallic; NM, nonmetal; PM, paramagnetic; PM-M, paramagnetic metal; SC, superconductivity; SR, short-range; SR-AFM, short-range AFM correlation.We now present the detailed experimental results in support of the phase diagram. Fig. 3A shows the temperature dependence of ρ of Fe1−xPdxTe for 0 ≤ x ≤ 0.5 (the experimental technique is described in SI Text). Similar to previous observations (45), the electrical resistivity of undoped (x = 0) Fe1.09Te increases with decreasing temperature at high temperatures, but sharply decreases below TN/S ∼70 K; this is caused by the first-order coupled structural and AFM phase transition (45, 46). Below TN/S, the structure of Fe1.09Te becomes monoclinic. Upon Pd doping, TN/S is suppressed, and the first-order transition is no longer observed at x = 0.2. The smooth and monotonic temperature dependence of ρ indicates the absence of both structural and AFM transitions at 0.1 < x < 0.2. Though it sustains a nonmetallic temperature dependence (dρ/dT < 0) down to 2 K, the magnitude of the electrical resistivity deceases with increasing x in the entire doping range above TN/S (Fig. 3 A and C). For 0.5 < x ≤ 1.0, the electrical resistivity of Fe1−xPdxTe continues to decrease with increasing x, as shown in Fig. 3B. More interesting is that ρ shows metallic behavior (dρ/dT > 0) in the entire temperature range measured for 0.5 < x ≤ 1.0. We believe that the cross-over from nonmetallic to metallic character is associated with the structural change at xs. Furthermore, there is a sharp drop of resistivity at low temperatures for 0.88 < x ≤ 1.0, as shown in Fig. 3D. This drop is due to the emergence of superconductivity because the magnetic susceptibility shows diamagnetism as well in the temperature range (Fig. 3H). Note that the superconducting transition temperature Tc increases with increasing x.Open in a separate windowFig. 3.(A and B) Temperature dependence of the electrical resistivity (ρ) of Pd1−xFexTe at the indicated compositions. (C and D) Temperature dependence of ρ near the first-order AFM/structural transition in the low-doping region, and the superconducting transition in the high-doping region plotted as ρ/ρ(5 K) vs. T, respectively. (E–G) Temperature dependence of the magnetic susceptibility (χ) of Fe1−xPdxTe measured at a magnetic field of 1 T. G Inset shows χ vs. x = 0.99 and 1.0; H is the χ(T) measured at 20 oersted (Oe) for 0.97 ≤ x ≤ 1.0.With Pd doping, the magnetic properties of Fe1−xPdxTe also change. Fig. 3 E–H shows the temperature dependence of the magnetic susceptibility, χ, at indicated values x. For 0 ≤ x < 0.2, χ initially increases with decreasing temperature, and then drops steeply at TS/N (Fig. 3E); this confirms the first-order nature of the structural/AFM transition at 0 ≤ x < 0.2. Such a feature is absent for xxN/S ∼0.15. Instead, there is a characteristic peak in magnetic susceptibility at TN (Fig. 3F). Because there is no anomaly in electrical resistivity (Fig. 3 A–C) or specific heat, the susceptibility peak cannot be due to a true phase transition, but rather marks the onset of SR AFM correlation as seen in other materials (47). With increasing x, TN moves to lower temperatures, and the magnitude of χ decreases as well (Fig. 3F). Though the characteristic peak is expected to completely vanish around x = xc ∼0.88, the magnetic susceptibility reveals a dramatic increase below another characteristic temperature TFM at x ≥ 0.78, as shown in Fig. 3G, which indicates the onset of ferromagnetic ordering at TFM. With increasing x, TFM initially increases then decreases after reaching the maximum (∼190 K) at x = xc ∼0.88. Interestingly, the low-temperature susceptibility peak is still present for samples with x = 0.78, 0.80, 0.83, and 0.88, which suggests the coexistence of FM ordering and AFM interactions in this doping region. Above xc, χ(T) continues to show FM behavior at high temperatures, but becomes negative at low temperatures (below Tc) under low magnetic fields, as depicted in Fig. 3H. The negative χ below Tc indicates the emergence of superconductivity, because their electrical resistivities also drop sharply (Fig. 3D). Whereas TFM decreases, Tc increases with increasing x (above xc). As shown Fig. 3G Inset, χ(T) exhibits paramagnetic behavior above Tc for x = 0.99 and 1.0, indicating the complete suppression of FM. However, Tc continues to increase with x (Fig. 3H).Clearly, there is a coexistence of AFM and FM ordering at 0.78 ≤ x ≤ 0.88, and of FM ordering and SC at 0.88 ≤ x ≤ 0.98 at low temperatures for Fe1−xPdxTe, which is further supported by the following results. First, the high-temperature magnetic susceptibility data allows us to extract both Curie–Weiss temperature CW) and effective magnetic moment (μeff) by fitting our experimental data to the Curie–Weiss law , where NA is the Avogadro’s constant and kB is Boltzmann’s constant. The x dependences of θCW and μeff are shown in Fig. 4 A and B, respectively. Note that θCW < 0 for x ≤ 0.8, indicating that the dominant magnetic interaction is AFM in this doping region. The fact that | θCW | >> TN/S implies 2D AFM ordering for x < 0.2, whereas a small θCW value at 0.2 ≤ x ≤ 0.8 is consistent with our interpretation that there is short-range AFM correlation. And θCW > 0 for 0.8 ≤ x < 0.98, which confirms the ferromagnetic interaction in this doping range. Although θCW is comparable to the corresponding transition temperature TFM, the effective magnetic moment is small, as shown in Fig. 4B, which implies that the FM ordering results mainly from Fe, even though Pd has to be the vehicle of coupling between Fe atoms (see discussion below). According to the single-crystal X-ray refinement results (Tables S1S3), the actual Fe concentration is very close to the nominal value for 0.8 ≤ x ≤ 0.98. We thus estimate the effective magnetic moment of Fe using the nominal Fe concentrations; this gives μeff = 1.60μB/Fe (x = 0.80), 2.42μB/Fe (x = 0.83), 3.79μB/Fe (x = 0.88), 4.21μB/Fe (x = 0.92), 3.53μB/Fe (x = 0.95), and 2.52μB/Fe (x = 0.98). Note that the highest magnetic moment for x = 0.92 is close to the Fe moment in Fe1.09Te ( 4.65μB/Fe), as plotted in Fig. 4B. In the latter case, the actual ordered magnetic moment is ∼3.31μB/Fe (28), smaller than that obtained from the Curie–Weiss constant; this is explained as due to the itinerant nature of the electrons in FeTe (28). For Fe1−xPdxTe, we may estimate the ordered magnetic moment from the magnetization. Fig. 4C shows the field dependence (H) of magnetization (M) at T = 3 K for x between 0.80 and 0.98. Note that M(H) reveals a well-defined hysteresis loop, confirming the nature of FM ordering for 0.8 ≤ x ≤ 0.98. Though M(H) is not saturated up to 6 T, we may estimate the lower bound of the ordered magnetic moment using M(H = 6 T) data shown in Fig. 4B. At T = 3 K, μ ∼0.33μB/Fe (x = 0.80), 0.48μB/Fe (x = 0.83), 1.49μB/Fe (x = 0.88), 1.85μB/Fe (x = 0.92), 1.67μB/Fe (x = 0.95), and 0.22μB/Fe (x = 0.98). In addition to the fact that M (H = 6 T) < Msat, the small ordered magnetic moment could result from (i) the itinerant nature of the electrons and (ii) the reduced concentration of Fe in the highly Pd doped region. Nevertheless, the specific heat reveals anomalies in the compounds with high TFM (Fig. 4D), indicating a true phase transition at TFM.Open in a separate windowFig. 4.Fe1-xPdxTe (A) Doping dependence of Curie–Weiss temperature. (B) Doping dependence of the effective magnetic moment and magnetization at 6 T. (C) Magnetization vs. magnetic field at 3 K for x = 0.8, 0.88, 0.92, 0.95, and 0.98. (D) Specific heat vs. temperature for x = 0.92 and 0.88.From the results presented here, it is most likely that the FM dome is intrinsic, due to the ordering of the Fe magnetic moments. Within the dome, our single-crystal X-ray refinement indicates that the structure remains the same as that of pure PdTe with actual Fe concentration close to the nominal value, and there are no interstitial sites of either Pd or Fe, which strongly suggests that the FM ordering is not due to the formation of Fe clusters. In early studies, FM was detected in alloys of Fe in Pd (i.e., Pd1−xFex) at compositions down to 0.08% Fe (48), because magnetic moments on the Fe sites polarize the surrounding Pd matrix (4851). It seems that Fe doping in PdTe compound gives rise to a similar effect, i.e., Pd matrix mediates the magnetic interactions between Fe atoms.The central question is, how can both FM ordering and superconductivity coexist in the region of 0.88 ≤ x ≤ 0.98? Having excluded the possibility of structural phase separation, one may consider the scenario of electronic phase separation with either macroscopic coexistence of singlet SC and FM in different regions or microscopic coexistence of triplet SC and FM (11, 21, 52). If Fe1−xPdxTe were a triplet superconductor, a small amount of impurity or disorder would completely suppress superconductivity, as seen in Sr2RuO4 (53). Though we observe anomalous specific heat below Tc of PdTe (42), the insensitivity to Fe doping does not support triplet SC. For the former case, further experimental investigations are necessary, such as combined scanning electron microscopy and tunneling microscopy to directly probe both SC and FM regions. Theoretically, almost all Pd d orbitals in PdTe are occupied due to the covalency of the Pd–Te bonding, according to recent first-principle calculations (44), and this leads to a strong suppression of the local magnetic moment (44). As shown in Fig. 3G (Inset), the magnetic susceptibility of PdTe (x = 1) indeed exhibits paramagnetic behavior above Tc. Upon substitution of Pd by Fe, Tc decreases nearly linearly with increasing Fe concentration (1 − x; Fig. 2). As Fe concentration reaches ∼3% (1 − x = 0.03), both SC and FM ordering coexist. The fact that FM ordering does not immediately kill superconductivity is likely due to the unique electronic structure of PdTe. First-principles calculations indicate that the states in the proximity of the Fermi level consist mainly of Te p electrons, which are weakly hybridized with Pd eg orbitals (44); this is in dramatic contrast to the electronic structure of FeTe, in which the states near the Fermi level derived from Fe with direct Fe–Fe interactions, and the Te p states lie well below the Fermi level and hybridized weakly with the Fe d states (54). For PdTe, the partial doping of Fe may lead to an even weaker hybridization between Te p and Pd eg orbitals, due to the shift of Fermi level. Our experimental observation is in support of this scenario: with increasing Fe concentration, superconductivity is gradually suppressed to zero, at which the FM ordering reaches the maximum (Figs. 2 and and4).4). After Tc reaches zero at xc, the disappearance of the superconducting energy gap allows the rearrangement of electronic structure, which is in favor of antiferromagnetic interaction; the latter apparently competes with the existing FM interaction, thus leading to the complete suppression of FM at x < 0.78.In summary, the substitution of Fe by Pd results in an extremely rich phase diagram for Fe1−xPdxTe. Powder and single-crystal X-ray diffraction measurements both indicate that there is doping-induced structural transition from a tetragonal phase at x < xs ∼0.6 to a hexagonal phase at x > xs. Correspondingly, the electrical resistivity changes from nonmetallic character at x < xs to metallic behavior at x > xs. Magnetically, the system undergoes a first-order AFM transition at TN/S for xxN/S ∼0.15, with the structure transition occurring at the same temperature from tetragonal to monoclinic. When TN/S approaches zero at xN/S, there is no longer a temperature-induced structural transition, whereas short-range AFM correlation exists for xN/S < x ≤ 0.88. Superconductivity sets in at xxc ∼0.88. In addition, a gigantic FM dome with 0.78 ≤ x ≤ 0.98 is centered at the critical concentration xc, where both AFM and SC vanish. The coexistence of FM ordering with either SR AFM or SC results most likely from the unique electronic structure of Fe1−xPdxTe. With the low Pd concentration, both electrical conduction and magnetism of Fe1−xPdxTe are determined by Fe/Pd d electrons, because they are near the Fermi level. Due to the ferromagnetic instability and the extended orbitals of 4d electrons of Pd, both AFM interactions and electrical resistivity in Fe1−xPdxTe decrease with increasing x. Before the complete suppression of SR AFM, FM ordering occurs due to the Fe magnetic moment polarization in the Pd matrix. However, the hexagonal structure at x > xs with Pd/Fe in an octahedral environment leads to more localized d electrons from Pd/Fe but itinerant p electrons from Te. It seems that the weak hybridization between pd electrons allows for the magnetic ordering of Fe/Pd moments and superconducting ordering of p conduction electrons. Nevertheless, these two orderings compete with each other. Hence, the physics of Fe1−xPdxTe in the hexagonal phase is in contrast to known Fe-based superconductors, in which physical properties are mainly determined by Fe d states. Fe1−xPdxTe provides another and rare example for studying the interplay between SC, FM, and AFM ordering, where two separate sets of electrons are responsible for FM ordering and superconductivity, respectively.  相似文献   

12.
Developing new soft magnetic amorphous alloys with a low cost and high saturation magnetization (Bs) in a simple alloy system has attracted substantial attention for industrialization and commercialization. Herein, the glass-forming ability (GFA), thermodynamic properties, soft magnetic properties, and atomic structures of Fe80+xSi5−xB15 (x = 0–4) amorphous soft magnetic alloys were investigated by ab initio molecular dynamics (AIMD) simulations and experiments. The pair distribution function (PDF), Voronoi polyhedron (VP), coordination number (CN), and chemical short- range order (CSRO) were analyzed based on the AIMD simulations for elucidating the correlations between the atomic structures with the glass-forming ability and magnetic properties. For the studied compositions, the Fe82Si3B15 amorphous alloy was found to exhibit the strongest solute–solute avoidance effect, the longest Fe-Fe bond, a relatively high partial CN for the Fe-Fe pair, and the most pronounced tendency to form more stable clusters. The simulation results indicated that Fe82Si3B15 was the optimum composition balancing the saturation magnetization and the GFA. This prediction was confirmed by experimental observations. The presented work provides a reference for synthesizing new Fe-Si-B magnetic amorphous alloys.  相似文献   

13.
Incorporating with inhomogeneous phases with high electroluminescence (EL) intensity to prepare smart meta-superconductors (SMSCs) is an effective method for increasing the superconducting transition temperature (Tc) and has been confirmed in both MgB2 and Bi(Pb)SrCaCuO systems. However, the increase of ΔTc (ΔTc = TcTcpure) has been quite small because of the low optimal concentrations of inhomogeneous phases. In this work, three kinds of MgB2 raw materials, namely, aMgB2, bMgB2, and cMgB2, were prepared with particle sizes decreasing in order. Inhomogeneous phases, Y2O3:Eu3+ and Y2O3:Eu3+/Ag, were also prepared and doped into MgB2 to study the influence of doping concentration on the ΔTc of MgB2 with different particle sizes. Results show that reducing the MgB2 particle size increases the optimal doping concentration of inhomogeneous phases, thereby increasing ΔTc. The optimal doping concentrations for aMgB2, bMgB2, and cMgB2 are 0.5%, 0.8%, and 1.2%, respectively. The corresponding ΔTc values are 0.4, 0.9, and 1.2 K, respectively. This work open a new approach to reinforcing increase of ΔTc in MgB2 SMSCs.  相似文献   

14.
Lead zirconate titanate (PZT)-based ceramics are used in numerous advanced applications, including sensors, displays, actuators, resonators, chips; however, the poor mechanical characteristics of these materials severely limits their utility in composite materials. To address this issue, we herein fabricate transgranular type PZT ceramic nanocomposites by a novel method. Thermodynamically metastable single perovskite-type Pb0.99(Zr0.52+xTi0.48)0.98Nb0.02O3+1.96x powders are prepared from a citrate precursor before both monoclinic and tetragonal ZrO2 nanoparticles ranging from 20 to 80 nm are precipitated in situ at a sintering temperature of 1260 °C. The effects of ZrO2 content on the microstructure, dielectric, and piezoelectric properties are investigated and the mechanism, by which ZrO2 toughened PZT is analyzed in detail. The ZrO2 nanoparticles underwent a tetragonal to monoclinic phase transition upon cooling. The fracture mode changed from intergranular to transgranular with increasing ZrO2 content. The incorporation of ZrO2 nanoparticles improved the mechanical and piezoelectric properties. The optimized piezoelectric properties (εT33/ε0 = 1398, tan δ = 0.024 d33 = 354 pC N−1, kp = 0.66 Qm = 78) are obtained when x = 0.02. Tc initially increased and subsequently decreased with increasing ZrO2 content. The highest Tc = (387 °C) and lowest εT33/ε0 was obtained at x = 0.01.  相似文献   

15.
On the attribution and additivity of binding energies   总被引:14,自引:5,他引:14       下载免费PDF全文
It can be useful to describe the Gibbs free energy changes for the binding to a protein of a molecule, A—B, and of its component parts, A and B, in terms of the “intrinsic binding energies” of A and B, ΔGAi and ΔGBi, and a “connection Gibbs energy,” ΔGs that is derived largely from changes in translational and rotational entropy. This empirical approach avoids the difficult or insoluble problem of interpreting observed ΔH and TΔS values for aqueous solutions. The ΔGi and ΔGs terms can be large for binding to enzymes and other proteins.  相似文献   

16.
Bulk ceria-zirconia solid solutions (Ce1−xZrxO2−δ, CZO) are highly suited for application as oxygen storage materials in automotive three-way catalytic converters (TWC) due to the high levels of achievable oxygen non-stoichiometry δ. In thin film CZO, the oxygen storage properties are expected to be further enhanced. The present study addresses this aspect. CZO thin films with 0 ≤ x ≤ 1 were investigated. A unique nano-thermogravimetric method for thin films that is based on the resonant nanobalance approach for high-temperature characterization of oxygen non-stoichiometry in CZO was implemented. The high-temperature electrical conductivity and the non-stoichiometry δ of CZO were measured under oxygen partial pressures pO2 in the range of 10−24–0.2 bar. Markedly enhanced reducibility and electronic conductivity of CeO2-ZrO2 as compared to CeO2−δ and ZrO2 were observed. A comparison of temperature- and pO2-dependences of the non-stoichiometry of thin films with literature data for bulk Ce1−xZrxO2−δ shows enhanced reducibility in the former. The maximum conductivity was found for Ce0.8Zr0.2O2−δ, whereas Ce0.5Zr0.5O2-δ showed the highest non-stoichiometry, yielding δ = 0.16 at 900 °C and pO2 of 10−14 bar. The defect interactions in Ce1−xZrxO2−δ are analyzed in the framework of defect models for ceria and zirconia.  相似文献   

17.
This paper refers to the structural and magnetic properties of [(Fe80Nb6B14)0.88Dy0.12]1−xZrx (x = 0; 0.01; 0.02; 0.05; 0.1; 0.2; 0.3; 0.5) alloys obtained by the vacuum mold suction casting method. The analysis of the phase contribution indicated a change in the compositions of the alloys. For x < 0.05, occurrence of the dominant Dy2Fe14B phase was observed, while a further increase in the Zr content led to the increasing contribution of the Fe–Zr compounds and, simultaneously, separation of crystalline Dy. The dilution of (Fe80Nb6B14)0.88Dy0.12 in Zr strongly influenced the magnetization processes of the examined alloys. Generally, with the increasing x parameter, we observed a decrease in coercivity; however, the unexpected increase in magnetic saturation and remanence for x = 0.2 and x = 0.3 was shown and discussed.  相似文献   

18.
PbTi1−xFexO3−δ (x = 0, 0.3, 0.5, and 0.7) ceramics were prepared using the classical solid-state reaction method. The investigated system presented properties that were derived from composition, microstructure, and oxygen deficiency. The phase investigations indicated that all of the samples were well crystallized, and the formation of a cubic structure with small traces of impurities was promoted, in addition to a tetragonal structure, as Fe3+ concentration increased. The scanning electron microscopy (SEM) images for PbTi1−xFexO3−δ ceramics revealed microstructures that were inhomogeneous with an intergranular porosity. The dielectric permittivity increased systematically with Fe3+ concentration, increasing up to x = 0.7. A complex impedance analysis revealed the presence of multiple semicircles in the spectra, demonstrating a local electrical inhomogeneity due the different microstructures and amounts of oxygen vacancies distributed within the sample. The increase of the substitution with Fe3+ ions onto Ti4+ sites led to the improvement of the magnetic properties due to the gradual increase in the interactions between Fe3+ ions, which were mediated by the presence of oxygen vacancies. The PbTi1−xFexO3−δ became a multifunctional system with reasonable dielectric, piezoelectric, and magnetic characteristics, making it suitable for application in magnetoelectric devices.  相似文献   

19.
The new generation of high-frequency and high-efficiency motors has high demands on the soft magnetic properties, mechanical properties and corrosion resistance of its core materials. Bulk amorphous and nanocrystalline alloys not only meet its performance requirements but also conform to the current technical concept of integrated forming. At present, spark plasma sintering (SPS) is expected to break through the cooling-capacity limitation of traditional casting technology with high possibility to fabricate bulk metallic glasses (BMGs). In this study, Fe84Si7B5C2Cr2 soft magnetic amorphous powders with high sphericity were prepared by a new atomization technology, and its characteristic temperature was measured by DSC to determine the SPS temperature. The SEM, XRD, VSM and universal testing machine were used to analyze the compacts at different sintering temperatures. The results show that the powders cannot be consolidated by cold pressing (50 and 500 MPa) or SPS temperature below 753 K (glass transition temperature Tg = 767 K), and the tap density is only 4.46 g·cm−3. When SPS temperature reached above 773 K, however, the compact could be prepared smoothly, and the density, saturation magnetization, coercivity and compressive strength of the compacts increased with the elevated sintering temperature. In addition, due to superheating, crystallization occurred even when the sintering temperature was lower than 829 K (with the first crystallization onset temperature being Tx1 = 829 K). The compact was almost completely crystallized at 813 K, resulting in a sharp increase in the coercivity of the compact from 55.55 A·m−1 at 793 K to 443.17 A·m−1. It is noted that the nanocrystals kept growing in size as the temperature increased to 833 K, which increased the coercivity remarkably but showed an enhanced saturation magnetization.  相似文献   

20.
First-order isostructural magnetoelastic transition with large magnetization difference and controllable thermal hysteresis are highly desirable in the development of high-performance magnetocaloric materials used for energy-efficient and environmental-friendly magnetic refrigeration. Here, we demonstrate large magnetocaloric effect covering the temperature range from 325 K to 245 K in Laves phase Hf1−xTaxFe2 (x = 0.13, 0.14, 0.15, 0.16) alloys undergoing the magnetoelastic transition from antiferromagnetic (AFM) state to ferromagnetic (FM) state on decreasing the temperature. It is shown that with the increase of Ta content, the nature of AFM to FM transition is gradually changed from second-order to first-order. Based on the direct measurements, large reversible adiabatic temperature change (ΔTad) values of 2.7 K and 3.4 K have been achieved under a low magnetic field change of 1.5 T in the Hf0.85Ta0.15Fe2 and Hf0.84Ta0.16Fe2 alloys with the first-order magnetoelastic transition, respectively. Such remarkable magnetocaloric response is attributed to the rather low thermal hysteresis upon the transition as these two alloys are close to intermediate composition point of second-order transition converting to first-order transition.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号