Journal of Materials Science & Technology  2019 , 35 (8): 1655-1661 https://doi.org/10.1016/j.jmst.2019.03.030

Orginal Article

Microstructure and soft-magnetic properties of FeCoPCCu nanocrystalline alloys

Long Houa1, Xingdu Fana1, Qianqian Wanga, Weiming Yangb, Baolong Shenab*

a Jiangsu Key Laboratory for Advanced Metallic Materials, School of Materials Science and Engineering, Southeast University, Nanjing 211189, China
b Institute of Massive Amorphous Metal Science, China University of Mining and Technology, Xuzhou 221116, China

Corresponding authors:   *Corresponding author at: Jiangsu Key Laboratory for Advanced Metallic Mate-rials, School of Materials Science and Engineering, Southeast University, Nanjing 211189, China. E-mail address: blshen@seu.edu.cn (B. Shen).

Received: 2018-12-7

Revised:  2019-01-17

Accepted:  2019-01-20

Online:  2019-08-05

Copyright:  2019 Editorial board of Journal of Materials Science & Technology Copyright reserved, Editorial board of Journal of Materials Science & Technology

About authors:

1Authors contributed equally to this work.

More

Abstract

Fe83.2-xCoxP10C6Cu0.8 (x = 0, 4, 6, 8 and 10) alloys with a high amorphous-forming ability and good soft-magnetic properties were successfully synthesized. Saturation magnetic flux density (Bs) is effectively enhanced from 1.53 T to 1.61 T for as-quenched alloy by minor Co addition, which is consistent well with the result of the linear relationship between average magnetic moment and magnetic valence. For Co-contained alloys, the value of corecivity (Hc) is mainly determined by magneto-crystalline anisotropy, while effective permeability (μe) is dominated by grain size and average saturation polarization. After proper heat treatment, the Fe79.2Co4P10C6Cu0.8 nanocrystalline alloy exhibited excellent soft-magnetic properties including a high Bs of 1.8 T, a low Hc of 6.6 A/m and a high μe of 15,510, which is closely related to the high volume fraction of α-(Fe, Co) grains and refined uniform nanocrystalline microstructure.

Keywords: Anocrystalline alloy ; Microstructure ; Co addition ; Soft-magnetic properties ; Ferromagnetic exchange-coupling

0

PDF (3487KB) Metadata Metrics Related articles

Cite this article Export EndNote Ris Bibtex

Long Hou, Xingdu Fan, Qianqian Wang, Weiming Yang, Baolong Shen. Microstructure and soft-magnetic properties of FeCoPCCu nanocrystalline alloys[J]. Journal of Materials Science & Technology, 2019, 35(8): 1655-1661 https://doi.org/10.1016/j.jmst.2019.03.030

1. Introduction

Fe-based nanocrystalline alloys with amorphous/crystalline composite structures have attracted great attention due to their excellent soft-magnetic properties such as high saturation magnetic flux density (Bs), low coercivity (Hc) and high effective permeability (μe) [[1], [2], [3]]. Among them, FeSiBNbCu (FINEMET) alloy has been widely used in electronic devices due to its high μe and low core loss [4,5]. However, the most commonly used FINEMET alloy exhibits a rather low Bs of only 1.24 T, which limits its industrial applications, as Bs is an important property in applications, the higher the Bs is, the smaller and more efficient the electronic devices can be made. Therefore, with the rapid development of modern electronics industry, it is important to develop nanocrystalline alloys with a higher Bs. For ferromagnetic soft magnetic alloys, the Bs is basically determined by ferromagnetic elements contents such as Fe, Co, and Ni, but non-ferromagnetic elements are also important for obtaining good magnetic softness, particularly, in some high entropy soft magnetic alloys [6]. The optimization of non-ferromagnetic elements is required to achieve an optimum combination of properties, especially to get the balance between mechanical and magnetic properties. For Fe-based amorphous/nanocrystalline alloys, the addition of non-ferromagnetic metalloid elements such as C [7,8], Si [9] and P [10,11] has been experimentally proved to favor amorphous formation, while metal elements such as Nb [12,13], Mn [14], Hf [15], and Mo [16,17] can inhibit the quick growth of nanocrystals due to their large atomic size and slow diffusion. Generally, a high Fe content is essential to achieve a high Bs, but an increase in the Fe content implies a decrease in non-ferromagnetic elements contents inside the alloy, leading to a series of problems such as the decrease of amorphous-forming ability (AFA), the decreasing thermal stability of residual amorphous phase, and the abnormal growth of crystalline grains, etc. Accordingly, how to achieve high Bs, high AFA and excellent soft-magnetic properties with high Fe content has always been a balance question in the study of nanocrystalline alloys.

Recently, by crystallizing hetero-amorphous phase, nanocrystalline FeSiBPCu alloy (NANOMET) has been developed with a Bs of >1.8 T, a low Hc of < 10 A/m, and especially a low core loss of 0.32 W/kg at 1.5 T and 50 Hz [[18], [19], [20]]. Moreover, this alloy also possesses low material cost as it does not contain noble metal elements. However, the relatively low AFA, by which the ribbon thickness is limited to 20 μm or less, constrains its application [18,19,21]. More recently, the FePCCu nanocrystalline alloys were successfully developed with good soft-magnetic properties including a relatively high Bs of 1.64-1.65 T, a low Hc of 3.3-3.9 A/m, and a high μe of 21000, respectively, as well as a higher AFA with 25 μm in critical thickness [22,23]. However, the soft-magnetic properties, especially the Bs of these alloys, are still not comparable to those of the NANOMET alloy.

It has been reported that partial substitution of Co for Fe in Fe-based amorphous/nanocrystalline alloys is effective for AFA improvement and Bs tuning [21,[24], [25], [26], [27]]. However, in some FeSiB-type alloys such as the FeSiBNbCu alloy, the AFA is significantly enhanced from ribbon to bulk glassy alloy with a diameter of 1.5 mm by 10 at.% Co addition [28], whereas the Bs does not increase, but exhibits a slight decrease instead. Similar results are also found in the FeSiBPCu alloys [29,30].

Therefore, with the aim of synthesizing novel alloys with high AFA without losing the Bs, the Fe83.2-xCoxP10C6Cu0.8 (x = 0, 4, 6, 8 and 10) alloys have been developed. The amorphous formation, thermal stability and microstructure evolution of the FePCCu alloys with different amounts of Co substitution were investigated. The mechanism of Co addition on soft-magnetic properties especially the Bs was also discussed.

2. Experimental

Alloy ingots with nominal compositions of Fe83.2-xCoxP10C6Cu0.8 (x = 0, 4, 6, 8 and 10) were prepared by induction melting the mixtures of pure Fe (99.99 mass%), Cu (99.995 mass%), Co (99.99 mass%), pre-alloyed Fe-P ingots (consisting of 75 mass% Fe and 25 mass% P) and Fe-C ingots (consisting of 96 mass% Fe and 4 mass% C) in an induction melting furnace under the protection of an argon atmosphere. Part of master alloy ingot was re-melted in a special quartz tube and then sprayed rapidly into the surface of a high-speed rotating single roller (roller speed 40 m/s, cavity pressure difference 0.015-0.02 MPa) to form as-quenched (AQ) ribbons. The width and thickness of ribbon samples are 1 mm and 22-28 μm, respectively.

Microstructure of ribbons was characterized by X-ray diffraction (XRD, Bruker D8 Discover diffractometer) with Cu Kα1 radiation and transmission electron microscopy (TEM, JEM2000ex). Thermal parameters were measured by using a differential scanning calorimeter (DSC, NETZSCH 404 F3) under a flow of high purity argon with a heating rate of 0.67 K/s. AQ ribbons were cut into 6 mm and then isothermally annealed in the vacuum chamber to develop nanocrystalline alloys. The Bs and Hc were measured using a vibrating sample magnetometer (VSM, Lake Shore 7410) under an applied field of 800 kA/m, and a B-H loop tracer (RIKEN BHS-40) under a field of 1 kA/m, respectively; μe was measured by an impedance analyzer (Agilent 4294 A) under a field of 1 A/m. For all magnetic measurements, three specimens were measured for each alloy to ensure credible results.

3. Results

Fig. 1 shows the XRD patterns of free surface of the Fe83.2-xCoxP10C6Cu0.8 (x = 0, 4, 6, 8 and 10) AQ ribbons. The Co-free alloy shows a small crystallization peak at 2θ = 44.5° which corresponds to the (110)-reflection of α-Fe crystalline phase, indicating a poor AFA. However, for the alloys with Co addition, only a typical broad diffraction feature without any detectable sharp peaks is observed, indicating the formation of amorphous structures. As a result, the addition of Co element can effectively improve the AFA of FePCCu alloys.

Fig. 1.   XRD patterns of AQ Fe83.2-xCoxP10C6Cu0.8 (x = 0, 4, 6, 8 and 10) alloy ribbons.

The thermal performance of the AQ ribbons was investigated by DSC, as shown in Fig. 2. Two exothermic peaks for each alloy can be easily observed, suggesting the crystallization proceeds in two stages. It was reported that the first exothermic peak corresponds to the crystallization of α-Fe phase while the second one corresponds to that of Fe3P and Fe3C compounds [22]. The onset temperature of the first crystallization (Tx1) has almost no change for the alloy with 4 at.% Co addition compared with that for Co-free alloy, while it increases slightly with further increase in Co to 10 at.%. Meanwhile, the second crystallization temperature (Tx2) increases slightly for alloy with 4 at.% Co, then decreases with further increase of Co to 10 at%. Accordingly, the temperature interval (ΔTx = Tx2 - Tx1) is first slightly enlarged and then decreased as listed in Table 1. The large ΔTx with minor Co substitution provides a wider crystallization window for the formation of α-(Fe, Co) phase without the second precipitation phases, which favors the achievement of good magnetic softness.

Fig. 2.   DSC curves of AQ Fe83.2-xCoxP10C6Cu0.8 (x = 0, 4, 6, 8 and 10) alloy ribbons with a heating rate of 0.67 °C/s.

Table 1   Thermal parameters and magnetic properties of Fe83.2-xCoxP10C6Cu0.8 nanocrystalline alloys annealed at 470 °C for 2 min.

AlloysThermal parametersMagnetic properties
Tx1
(oC)
Tx2
(oC)
ΔTx
(K)
Bs
(T)
Hc
(A/m)
μe
(1 kHz)
Fe83.2P10C6Cu0.83734851121.718.69910
Fe79.2Co4P10C6Cu0.83734891161.806.615,510
Fe77.2Co6P10C6Cu0.83764871111.7715.34800
Fe75.2Co8P10C6Cu0.83764851091.7519.14340
Fe73.2Co10P10C6Cu0.83804831031.7444.95030
Fe83.25P10C6Cu0.75 [22]3844911071.653.321100
Fe83.25P9C7Cu0.75 [23]3854901051.643.921000

New window

Fig. 3 shows the annealing temperature (Ta) dependence of Hc for Fe83.2-xCoxP10C6Cu0.8 alloys. The inset is the variation curve of Hc with annealing time at 390 °C for Fe79.2Co4P10C6Cu0.8 alloy, and this temperature is over Tx1 to ensure the crystallization. Because the Fe79.2Co4P10C6Cu0.8 alloy annealed for 2 min showed the lowest Hc, the optimum annealing time was determined to be 2 min for subsequently processed samples. As shown in Fig. 3, except for the Co-free alloy, all alloys exhibit a similar variation tendency, i.e. with increasing Ta from 350 to 470 °C, the values of Hc slightly decrease and then rise, followed by a distinct increase at 490 °C. The decrement of Hc of the alloy annealed below Tx1 is mainly caused by internal stress relief and the uniformity of amorphous structure enhancement, but between Tx1 and Tx2 is attributed to the precipitation of α-(Fe, Co) phase. When Ta is near or higher than Tx2, the quick growth of α-(Fe, Co) grains and the precipitation of Fe3P compound will greatly degrade the soft-magnetic properties. As for the alloy with no Co addition, the Hc is larger than that of the alloy with 4 at% Co, which is owed to the pre-existing α-Fe crystalline grains according to the XRD result. Meanwhile, there exists a slight increase in Hc before it rapidly increases due to the second inhomogeneous crystallization behavior. The Ta dependence of μe of Fe83.2-xCoxP10C6Cu0.8 alloys was also investigated. As shown in Fig. 4, with the increase in Ta, μe firstly increases gradually and reaches its maximum at 470 °C and then obviously drops. Thus, the optimal annealing temperature is 470 °C, at which all alloys exhibit the maximum μe.

Fig. 3.   Annealing temperatures (Ta) dependence of coercivity (Hc) for Fe83.2-xCoxP10C6Cu0.8 (x = 0, 4, 6, 8 and 10) alloy ribbons, and the inset is the dependence of Hc on annealing time (t) for Fe79.2Co4P10C6Cu0.8 alloy annealed at 390 °C.

Fig. 4.   Dependence of effective permeability (μe) on Ta for Fe83.2-xCoxP10C6Cu0.8 alloy ribbons.

The magnetic performance and corresponding microstructure of FeCoPCCu alloys during the annealing process were analyzed. Here, we only investigated the typical Fe83.2-xCoxP10C6Cu0.8 (x = 0, 4 and 10) alloys in detail. The Ta dependence of Bs of the Fe83.2-xCoxP10C6Cu0.8 alloys is shown in Fig. 5(a). The value of Bs clearly increases from 1.53 T to 1.61 T for the AQ alloys with different Co addition, which can be attributed to the strong ferromagnetic exchange-coupling between Co and Fe [25,31]. During the annealing process, the Bs of all alloys increases with increase in Ta, and the alloy with 4 at% Co addition possesses the maximum Bs of 1.80 T at 470 °C and 1.81 T at 490 °C, respectively. Then, the microstructure evolution for Fe79.2Co4P10C6Cu0.8 alloy annealed at different Ta was investigated by XRD, as shown in Fig. 5(b). When Ta is below 410 °C, only one sharp crystallization peak can be detected at 2θ = 44.5°, while annealed at 430 °C and above, three typical crystallization peaks of α-(Fe, Co) phase at 2θ = 44.5°, 64.5° and 82.3° can be detected. This indicates that α-(Fe, Co) phases are fully precipitated once the Ta reaches a certain value and finally, resulting in an increment of crystallization volume fraction (Vcry) of α-(Fe, Co) grains. For nanocrystalline alloys, the relationship between Bs and Vcry can be expressed as the following equation [32]:

Bs = Bsc Vcry + Bsa (1 - Vcry) (1)

where Bsc and Bsa are the saturation magnetic flux densities of the crystalline and amorphous phases, respectively. Moreover, the Bsc ($\widetilde{2}$.1 T) is larger than Bsa ($\widetilde{1}$.5 T) of Fe-based amorphous alloys [33]. Therefore, Bs strongly depends on the Vcry. According to Fig. 5(b), with the increment of Ta, a large amount of α-(Fe, Co) nanocrystals precipitates from amorphous matrix, which gives rise to the Vcry of α-(Fe, Co) grains. The increment in the number density and Vcry promotes the exchange coupling between grains and thus leads to a rapid increase in Bs [31,34]. However, with the Ta up to 490 °C, although the Bs reaches a maximum value of 1.81 T due to the higher Vcry, the Fe3P second crystalline phase is induced by high Ta, both the coarsening and growth of grains, and the precipitation of Fe3P compound inevitably causing a deterioration of Hc. As a result, optimal soft-magnetic properties were obtained by annealing at 470 °C for 2 min, where the Fe79.2Co4P10C6Cu0.8 nanocrystalline alloy exhibits a high Bs of 1.8 T, low Hc of 6.6 A/m and high μe of 15,510.

Fig. 5.   Magnetization performance and microstructure evolution of FeCoPCCu alloys during the annealing process.

Fig. 5(c) shows the hysteresis loops of Fe83.2-xCoxP10C6Cu0.8 alloys annealed at 470 °C for 2 min. With the substitution of Co, Bs increases firstly and then drops. The maximum value of Bs is achieved for the alloy with 4 at.% Co addition. Fig. 5(d) exhibits the microstructure change with Co content for the alloys annealed at optimal conditions (470 °C, 2 min). The alloy with 4 at.% Co shows a higher Vcry ($\widetilde{3}$7%) compared to that of other alloys due to the high integral intensity of diffraction peak of α-(Fe, Co) (the calculation of Vcry is based on Ref. [35]). Thus, the high Vcry of α-(Fe, Co) grains leads to the high Bs. Except for the influence of Vcry, the Bsc of α-(Fe, Co) phase ($\widetilde{2}$.45 T) is higher than that of α-Fe ($\widetilde{2}$.1 T) [24,33]. Thus, compared to the Co-free alloy, Bs can be affected by synergetic influence of Vcry and Bsc for the alloys with Co addition.

As soft-magnetic properties are greatly dependent on structure, microstructure evolution of nanocrystalline alloys was further investigated by TEM. Fig. 6 shows the bright-field TEM images of Fe83.2-xCoxP10C6Cu0.8 (x = 0 and 4) alloys under different annealing temperatures. As shown in Fig. 6(a), a small amount of nanocrystals can be observed in AQ Co-free alloy. The selected area electron diffraction (SAED) pattern (see the inset of Fig. 6(a)) also reveals the presence of crystalline grains, which is consistent well with the XRD measurement and forms a clear contrast with the AQ 4 at% Co-contained alloy, which only shows an amorphous structure feature [see Fig. 6(d)]. Fig. 6(b, c) and (e, f) show nanocrystalline microstructure of the Co-free and contained 4 at% alloys annealed at 410 °C and 470 °C for 2 min, respectively, in which the α-Fe/(Fe, Co) grains are randomly oriented in all the annealed samples according to the SAED patterns. Meanwhile, the average grain sizes (D) are $\widetilde{2}$7 and 20 nm for the annealed 410 °C and 470 °C Co-free alloy and $\widetilde{2}$4 and 18 nm for the Co-contained (4 at.%) alloy, respectively. This result indicates that, with the increase of Ta, both alloys exhibit grain refinement. Especially, the alloy with 4 at% Co represents a more uniform microstructure with much narrower grain distribution from 10 to 30 nm compared with the Co-free alloy, for which the α-Fe grains are distributed inhomogeneously due to the pre-existing α-Fe nanocrystals.

Fig. 6.   Brigh-field TEM images of Fe83.2-xCoxP10C6Cu0.8 alloys with x = 0 (a) AQ ribbons, (b) and (c) corresponding the ribbons annealed at 410 °C and 470 °C for 2 min, respectively; x = 4 (d) AQ ribbons, (e) and (f) corresponding the ribbons annealed at 410 °C and 470 °C for 2 min, respectively. The inset is the corresponding selected area electron diffraction (SAED) patterns and grain size distributions.

4. Discussion

In this study, it is found that Co addition effectively enhances the AFA and Bs of Fe83.2-xCoxP10C6Cu0.8 alloys. Here, we firstly discuss the reasons for the improvement of AFA. When Co is added into the base alloy, the degree of ordering around local Fe and Co atoms could be increased due to the strong bonding nature among the constituent elements [36]. Former research suggests that there is a special chemical short-range order existing in the atomic structure of melting state for amorphous precursor, which can be characterized as the solute-centered atomic cluster in transition metal-metalloid amorphous alloys [37,38]. Therefore, Co addition could lead to a special P/C-centered atomic cluster, which effectively increases the density and viscosity of the molten liquid, thereby inhibiting the long-range diffusion of atoms [17] and finally improving the thermal stability and AFA of alloys. It should be noted that the Co-free alloy exhibits slight crystallization in this work, which seems to be inconsistent with our previous work [22]. The deterioration of AFA may be attributed to the thicker ribbon thickness and different purities of raw materials of the alloys.

Secondly, the increase of Bs by Co addition can be interpreted by the theory of strong ferromagnetism according to magnetic valence theory [31,39]. Based on this theory, the average magnetic valence of $\bar{V}$m can be expressed as $\bar{V}$m=∑Vm,iXi, where Xi is the atomic fraction and Vm,i is the magnetic valence of the ith element, respectively, with Vm=2$N^\uparrow_d$ -Z ($N^\uparrow_d$ is the number of spin-up d-band electrons per atom and Z is chemical valence). The above theory is based on the strong ferromagnetism, i.e. $N^\uparrow_d$ for the magnetic elements (Fe, Co and Cu) and non-ferromagnetic elements (P and C) are 5 and 0, respectively. Therefore, the calculated Vm for Fe, Co, P, C and Cu atoms are 2, 1, -5, -4 and -1, respectively, and finally the $\bar{V}$m of the studied AQ alloys can be obtained. Meanwhile, the magnetic moment per magnetic atom (μm) can be calculated by using μm = VMJs/(NAμB), where VM is the molar volume of ferromagnetic atom, Js is the saturated magnetic polarization, NA and μB denotes the Avogadro constant and Bohr magneton, respectively [40]. Accordingly, the relationship between μm and $\bar{V}$m can be obtained as shown in Fig. 7. As a result, the μm of the base alloy is extremely low while it can be significant strengthened by Co addition and exhibits a linear relationship to $\bar{V}$m, and thus the deviation from linearity without Co addition implies that the Fe(Co)PCCu alloy undergoes a distinct transition behavior from weak ferromagnetism to strong ferromagnetism, which is similar to the previously work [38,39]. Moreover, based on the Bethe-Slater curve [34] (see the inset of Fig. 7), there should exist an optimal value of rab/rd (rab is the interatomic distance and rd is the radius of an unfilled d-shell), where the strongest exchange interacting strength is achieved. For the Fe83.2-xCoxP10C6Cu0.8 alloys, the optimal position may lie in the alloy with composition of x = 4, where it possesses the highest Bs. Noting that the effect of Co concentration on Bs also depends on alloy compositions, different alloy system may exhibit different variation trend of Bs towards Co concentration, such as in the NANOMET-type alloys [29,41], the Bs decreases only when the Co content is over 25 at.%, while in the FeCoPC alloy [42], the maximum Bs is achieved for the alloy with 5 at.% Co addition, which is similar to our work.

Fig. 7.   Relationship between average magnetic moment and magnetic valence of AQ Fe83.2-xCoxP10C6Cu0.8 alloys, the inset is the Bethe-Slater curve with the different elements.

Thirdly, we discuss the abnormal variation of soft-magnetic properties towards microstructure for Co-contained nanocrystalline alloys. Here, the typical alloy with 4 at.% Co addition was also used for investigation. According to the XRD and TEM measurement, there is an obvious decrease in D due to the grain refinement with increasing Ta, but the Hc exhibits an increasing variation (see the light green area in Fig. 8(a)). The result seems to be contrary with Herzer's random anisotropy model [43], which points out that the value of Hc follows the D6 proportional rule. The first reason is that the average D for 4 at.% Co-contained alloy is 18 nm, which may be larger than ferromagnetic exchange correlation length (Lex, only 5-10 nm for Co-based alloy [44]). Secondly, the higher Vcry of α-(Fe, Co) grains also induces larger magneto-crystalline anisotropy (<K>) as the magneto-crystalline anisotropy constant of α-(Fe, Co) is greater than that of α-Fe [45]. Considering the above two factors, the ferromagnetic exchange-coupling interaction may not be too strong to effectively average out the large < K> in Co-contained alloy, hence leading to a larger Hc. As μe generally takes an opposite towards Hc, μe should be theoretically small when < K> is large. But, actually, there is a significant increase in μe before 470 °C, where large < K> is induced, indicating that < K> is not dominant. Fig. 8(b) shows the μe, Vcry and Bs dependent on Ta. It can be clearly seen that the variation of μe is quite in accordance with that of Vcry and Bs. Meanwhile, the variation tendency of μe is also consistent well with the opposite change of D. Accordingly, for the FeCoPCCu nanocrystalline alloys, it is believed that the value of μe is mainly affected by the average saturation polarization (Js) and D, while the value of Hc is mainly determined by < K>.

Fig. 8.   (a) Change of coercivity (Hc) and grain size (D) with annealing temperature (Ta), (b) the variation trend of effective permeability (μe), crystallization volume fraction (Vcry) and saturation magnetization (Bs) dependent on Ta for Fe79.2Co4P10C6Cu0.8 alloy.

Finally, the microstructure evolution of Fe(Co)PCCu alloys is summarized as illustrated in Fig. 9. There exist a few α-Fe nanocrystals embedded in amorphous matrix for Co-free alloy. Under heat treatment, except for the pre-existing α-Fe crystalline grains growth, a large number of new α-Fe grains are separated out and continuous to grow, leading to the slight increase in Hc. When annealed at a proper temperature (e.g. 470 °C), the maximum nucleation rate may occur, where the highest number density of nanocrystals with smallest D is formed, leading to a non-uniform refined nanocrystalline structure. On the contrary, the alloy with 4 at.% Co addition displays a typical amorphous structure in AQ state. It has been reported that the simultaneous additions of P and Cu to Fe-based alloys could induce some inhomogeneities, such as α-Fe nano-clusters during the quenching process, hence the separated Cu/P-Cu clusters might act as nucleation sites for α-Fe [[46], [47], [48]]. Although, in our work, it was not found due to its very small size and low Vcry, the influence of these clusters must be considered due to the large positive mixing enthalpy of 13 and 6 kJ/mol between Fe/Co and Cu atoms [36]. Thus, with Co addition, the primary crystallization phase changes from α-Fe to α-Fe and α-(Fe, Co) combined phases. The competing formation of different phases is beneficial for the AFA. Meanwhile, with the increase in Ta, the nucleation, growth and interaction of α-(Fe, Co) grains adhering to Cu/P-Cu clusters leads to a more uniform nanostructure with a higher number density of nanocrystals. As listed in Table 1, compared to the reported nanocrystalline Fe83.25P10C6Cu0.75 [22] and Fe83.25P9C7Cu0.75 [23] alloys, the Bs is significantly enhanced from 1.65 T to 1.74-1.8 T, which promises the application at higher excitation magnetic field. Meanwhile, the Fe79.2Co4P10C6Cu0.8 nanocrystalline alloy annealed at proper condition also combines with a relative low Hc of 6.6 A/m and high μe of 15,510. The present work reveals the mechanism of Co addition on tuning AFA and soft-magnetic properties of FeCoPCCu alloys, and promotes the potential application as magnetic functional materials.

Fig. 9.   Sketch map of microstructure evolution of the AQ alloys with x = 0 (A) and x = 4 (B) at different Ta.

5. Conclusion

In this work, Fe83.2-xCoxP10C6Cu0.8 (x = 0, 4, 6, 8 and 10) alloys with good soft-magnetic properties and higher AFA were obtained. The soft-magnetic properties towards microstructure evolution of FeCoPCCu alloys were investigated. It is found that Hc of Co-contained alloys exhibit different variation compared with that of Co-free alloy. For Co-contained alloys, Hc is mainly determined by < K> while μe is dominated by both D and Js. The value of Bs is effectively enhanced for both amorphous and nanocrystalline alloys with proper Co addition. Due to the high Vcry of α-(Fe, Co) grains and refined uniform nanocrystalline microstructure, the Fe79.2Co4P10C6Cu0.8 alloy annealed at 470 °C for 2 min was successfully developed combined with high Bs of 1.8 T, low Hc of 6.6 A/m and high μe of 15,510.

Acknowledgments

This work was supported by the National Key Research and Development Program of China (Grant No. 2016YFB0300502) and the National Natural Science Foundation of China (Grant Nos. 51631003, 51401052, 51871237 and 51501037).

The authors have declared that no competing interests exist.


/