The following article is Open access

Anomalous Codeposition of fcc NiFe Nanowires with 5–55% Fe and Their Morphology, Crystal Structure and Magnetic Properties

, , , and

Published 29 December 2015 © The Author(s) 2015. Published by ECS.
, , Citation Oana Dragos et al 2016 J. Electrochem. Soc. 163 D83 DOI 10.1149/2.0771603jes

1945-7111/163/3/D83

Abstract

The electrodeposition of NiFe nanowires-with the length ∼3.0 μm and diameter 200 nm-using the porous anodized aluminum oxide (AAO) templates on the sputtered Au-back electrode (300 nm) using sulfate/chloride electrolyte solution and potential pulsed deposition. The electrode area of Au-AAO template, determined by the reversible one-electron transfer oxidation of K4Fe(CN)6, used as a probe in CV, was found to be ∼2.4 times larger than Au-thin film electrode. The anomalous codeposition phenomenon known as a "volcano" type curve-with a maximum in Fe-content in NiFe as a function of the applied potential-was observed in the literature. The observed results were explained through the limited mass transport of Fe+2 ions after the peak. This explanation is partially correct, but not complete. The electrodeposition of NiFe nanowire in this work resulted in a similar "volcano" type curve. The alternative explanation of anomalous codeposition-through the surface concentration of H+ dependent adsorption/desorption of FeOH+ and NiOH+ electroactive species-was proposed. The electrodeposition of NiFe nanowire arrays using a designed pulse potential method produced fcc NiFe nanowires with 5–55% Fe with controlled composition, length, and uniformity.

The distinct decrease of parallel coercivity of NiFe nanowire arrays having the same length was observed with increase of Fe-content in NiFe, which also correlates with the increase of their magnetic saturation. The process of magnetization reversal in NiFe nanowires arrays was investigated. We have found two sets of magnetic behavior of NiFe nanowires depending on composition (Ni92Fe8 and Ni79Fe21 vs. Ni60Fe40, Ni56Fe44, and Ni45Fe55).

Export citation and abstract BibTeX RIS

This is an open access article distributed under the terms of the Creative Commons Attribution Non-Commercial No Derivatives 4.0 License (CC BY-NC-ND, http://creativecommons.org/licenses/by-nc-nd/4.0/), which permits non-commercial reuse, distribution, and reproduction in any medium, provided the original work is not changed in any way and is properly cited. For permission for commercial reuse, please email: oa@electrochem.org.

The NiFe alloys are probably the most versatile soft ferromagnetic materials in use today.1 The electrodeposition of soft magnetic NiFe alloys has been extensively used in the manufacturing of modern recording heads in which Ni80Fe20 (Permalloy) and Ni45Fe55 alloys are used as a reader and writer shield materials.2 The use of soft magnetic properties of electrodeposited NiFe alloys in microelectromechanical (MEMS) devices has been reviewed.3

The electrodeposition of magnetic and nonmagnetic nanowires is currently one of the most active areas in advanced materials4,5 due to the potential applications in photochemical synthesis,6 catalytic synthesis,7 cell separation,8 biosensors,9 and medical therapy.10,11 The phenomena and potential applications require a deep understanding of the magnetism of nanowires.12 The studies on magnetic properties of NiFe nanowire arrays were intensified during last 15 years.1320

The electrodeposition of NiFe alloys exhibits a phenomenon known as an anomalous codeposition, which is characterized by the anomaly that less noble metal, i.e. Fe, deposits preferentially.21 The extent of anomalous codeposition can be evaluated by the "selectivity ratio" (SR), which is defined as an atomic ratio of Fe/Ni in the deposit to the molar ratio of Fe+2/Ni+2 in the electrolyte.22 The SR is dependent on solution condition (pH, presence of organic additives, concentration of electrolytes, temperature, etc), method of electrodeposition (constant current density, controlled potential, pulsed current or potential), and the thickness of deposit. The value of SR ranges from 1.0 (for non-anomalous codeposition) to 15 (for anomalous codeposition). The variation of composition of NiFe films affects the following properties of deposit: (i) stress,23 (ii) magnetostriction,24 (iii) crystal structure,25 and (iv) grain size.26 Thereby, the resulting changes profoundly influence the magnetic behavior of electrodeposited NiFe alloys, which was demonstrated recently.27

The present work describes the method of electrodeposition of NiFe nanowires into anodized aluminum oxide (AAO) templates with the controlled elemental composition using designed potential pulse electrodeposition. We report here a new explanation for the observed anomalous codeposition of NiFe in nanowires, through the surface concentration of H+ dependent adsorption/desorption of FeOH+ and NiOH+ electroactive species. We also demonstrate that the variation of Fe-content from 5–55% Fe in NiFe nanowires significantly affect their magnetic properties.

Experimental

The ordered NiFe nanowire arrays have been prepared by electrodeposition into AAO commercially available (Whatmann Inc.) templates. The pore size of Whatman AAO template was 50 μm, the nanowire diameter (dw) 200 nm, and interwire distance (D), i.e. center-to-center, 250 nm. A seed layer consisted of 5.0 nm Ti- used as an adhesion promoter-following by sputtering of 300 nm Au on one side of AAO template, which was used as a back-electrode and contact to the large Cu stage (see Fig. 1).

Figure 1.

Figure 1. The scheme of three electrode cell.

The voltammetry and electrodeposition of NiFe alloys were carried out in the sulfate/chloride plating solution containing: 0.4 M H3BO3, 0.3 M NH4Cl, 0.2 M NiSO4 6H2O, 0.02 M FeSO4 7H2O, 5 mM saccharin as a Na-salt, and 0.3 mM sodium lauryl sulfate (Na-LS). The electrolyte solution was adjusted to pH 3.0. The relatively high concentrations of Fe+2 and Ni+2 salts was used following recently proposed concentration optimization in the case of Fe nanowire electrodeposition.28

The cyanide-based solution for voltammetry and electrodeposition of gold-taken from the literature29 was composed of: KAu(CN)2 (6g/l), K2HPO4(120 g/l) and KH2PO4 (30 g/l) at pH 7.1.

The potassium hexacyanoferrate solution used for surface-area measurements contained 2.08 mM K4Fe (CN)6 and 0.1 M Na2SO4 as a supporting electrolyte.

The electrochemical experiments with potential pulse deposition, controlled potential deposition, cyclic voltammetry (CV), linear sweep voltammetry (LSV) and chronoamperometry were carried out at the room temperature in a three electrode cell made from Teflon with a volume of 25 ml, which is schematically shown in Fig. 1. The O-ring silicon polymer with diameter 2r = 1.0 cm, which defines the opening area to the flux of ions, was placed over Au working electrode which is in electrical contact with large Cu-stage. Two different Au electrodes where used as a working electrode, i.e. a Au-AAO template electrode (back-electrode) obtained by sputtering 5 nm Ti plus 300 nm of Au onto one side of the template and a Au-thin film electrode obtained by the simultaneous sputtering of 5 nm of Ti plus of 50 nm Au on silicon wafer. As a reference electrode saturated calomel electrode (SCE) was used and the counter electrode was Pt mesh.

The parameters for potential pulse electrodeposition of NiFe nanowires in quiescent plating solution were chosen on the basis of the CV experiment shown in Fig. 2-top. The cyclic voltammogram shows three regions (a-c), corresponding to: (a) reduction of H+ from the solution, (b) reduction and electrodeposition of Ni+2 and Fe+2 electroactive intermediates and (c) oxidation of NiFe film formed along the wave during the electrodeposition. A typical current-time curve for potential pulse electrodeposition of NiFe alloys is shown in Fig. 2-bottom, using the potential pulsed between −1.1 V/SCE during time-on of 2.5 s and the "rest" potential of −0.7 V/SCE during the time –off. The time-off during 1 s allowed the diffusion layer to "recover" from the deposition during time-on. Importantly, during the time-off the current is zero and thereby only cathodic potential affects the composition of NiFe alloys.

Figure 2.

Figure 2. (Top) Cyclic voltammogram of NiFe solution; Au-thin film electrode; Sweep rate: 20 mV/s, (Bottom) The current-time curve obtained at pulsed potential at Au-thin film electrode.

The potentiostatic electrodeposition of NiFe thin films at the controlled potential of Pt rotating disc electrode (RDE), as well as dissolution of deposit, and analysis by inductively coupled plasma (ICP) were the same as described earlier.30

The morphology of the NiFe nanowires was studied by scanning electron microscopy (SEM). The chemical composition of NiFe nanowires was determined by an energy dispersive X-ray spectrometry (EDS) facility attached to the SEM. The composition of the nanowires represents an average value of three spots along the nanowire arrays with a precision of ±1.0%. The surface roughness of the Au films was measured using atomic force microscopy (AFM) operating ex-situ in trapping mode. The magnetic properties of the samples were investigated by vibration sample magnetometer (VSM) and the crystal structure of the nanowires was characterized by X-ray diffraction (XRD).

The method used to prepare 500 nm NiFe thin films with 10–90% Fe for measurements of stress and magnetostriction was the same as described earlier.25 The resultant stress of the NiFe films -deposited on 6 in round alumina coated AlTiC wafers on Cu-seed layers- was obtained from the measured substrate curvature and calculated via Stoney's equation.31 Magnetostriction was measured by a Lafouda instrument based on a cantilever technique. The magnetic saturation, Bs, of electrodeposited NiFe alloys25 was measured with precision of ±2.0%.

Results and Discussion

Theoretical background

The anomalous codeposition phenomenon known as a "volcano" type curve-with a maximum in Fe-content in NiFe as a function of the applied potential or current density-was observed in the literature for the thin NiFe films3234 and NiFe nanowires.20 The observed results were explained through the limited mass transport of Fe+2 ions after the peak. This explantion is partially correct, but not complete. We propose in this work an alternative explanation of anomalous codeposition-through the surface concentration of H+ dependent adsorption/desorption of FeOH+ and NiOH+ electroactive species. We have used the same concept recently explaining the concentration gradient in thin CoNiFe35 and NiFe36 films, as a special case of anomalous codeposition. In the text below, we will outline two important features of the alternative view on anomalous codeposition, i.e. modified Bocris-Drazic-Despic (BDD) mechanism and the role of H+ in the anomalous codeposition of NiFe alloys.

We have modified the earlier BDD-mechanism and used it to explain some of the observed experimental results in the literature like: decrease of current efficiency in CoFe alloys with the increase of Fe-content in deposit,30 and the increase of tensile stress of CoFe alloys with the increase of Fe-content in deposit.37 The modified BDD-mechanism, which is relevant to all iron group metal electrodeposition in acidic media, is shown in reactions 1-4.

Equation ([1])

Equation ([2])

Equation ([3])

Equation ([4])

Our previous studies demonstrated that the higher concentration of M+2 ions in solution bring about a higher concentration of H+ through the hydrolysis reaction 1.30 The calculated concentrations38 of FeOH+ and NiOH+ species in the plating solution at pH 3.0 are negligible, i.e. [FeOH+] ∼ 10−9 M and [NiOH+] ∼ 10−9 M. Therefore, both species must be formed at the electrode surface through the diffusion and hydrolysis of solvated ions present in relatively high – calculated-concentration, i.e. [Fe+2] = 0.01 M and [Ni+2] = 0.15 M, according to the reaction 1, which is the first step in the Bockris, Drazic, Despic (BDD) mechanism.39

The formed MOH+ads species is reduced in the rate determining step39 in reaction 2 leading to the formation of neutral monovalent metal hydroxide. The key in modified BDD-mechanism is that the proton liberated through hydrolysis reaction1increases the concentration of protons at the electrode surface and can be reduced together with MOH+ads ions in reaction 3at a more negative potential than the H+ from the plating solution. Thereby, the total concentration of protons at the electrode surface, [H+]Total, is the sum of the concentrations of protons from the solution and protons liberated from the reaction 1, i.e. [H+]Total = [H+]Solution + [H+] Reaction.

Another key for anomalous codeposition is that H+ may act as a catalyst in the dissociation reaction of MOH+ads in reaction5, which is actually the backward reaction6in the first step of BDD-mechanism. The reaction5is an overall reaction which probably proceeds through two steps,i.e. protonation of MOH+ads followed by dissociation of the formed dication.

Equation ([5])

There is a competition for the protonation reaction between two adsorbed species, i.e. NiOH+ and FeOH+. Due to ∼1000 times higher dissociation constant of NiOH+ (pKa = 4.3)44 than FeOH+ (pKa = 7.2),40 NiOH+ is thermodynamically more favored to dissociate into solution while FeOH+ accumulates adsorbed at the surface. The rates of discharge of monohydroxide ions, FeOH+ads and NiOH+ads, are proportional to their respective surface fractions.

The last step in modified BDD-mechanism is reaction 4, i.e. one electron reduction of MOHads to M and OH at the electrode surface. The liberated OH base is protonated in acid-base reaction 6

Equation ([6])

Figure 3-left shows a typical "volcano" type curve obtained at controlled potential electrodepositions of thin film NiFe alloys using a Pt-RDE (125 rpm) with the peak in Fe-content at the potential −1.05 V/SCE. The "selectivity ratio" (SR) increases with the increase of negative potential from −0.7 V/SCE to −1.05 V/SCE (Fig. 3-right) indicating more anomalous codeposition and decreases toward less anomalous codeposition in the range of potentials from −1.05 V to −1.3 V/SCE.

Figure 3.

Figure 3. Effect of electrode potential on composition of NiFe films (Left) and selectivity ratio (Right), Pt-RDE (1 cm2); NiFe solution at pH 3.0; Rotation rate: 125 rpm.

Figure 4 shows the change of the partial current densities of metals (Ni, Fe) and hydrogen (H) during the electrodeposition of NiFe films as a function of the potential of Pt-RDE with the rotation rate of 125 rpm. The total current density iT, (iT = iFe + iNi + iH) shows the current plateau in the range of potentials from −0.7 to −0.95 V/SCE. The potential region of the plateau coincides with the reduction of all four electroactive species: FeOH+, NiOH+, H+reaction and H+solution, giving rise to the gradual increase of plateau. The total current density, iT, at the plateau is higher than the limiting current density, iL, calculated for H+solution for the solution at pH 3.0 under the same hydrodynamic conditions of Pt-RDE (iL = 0.63 mA/cm2).40 The total current density takes off in the range of potentials from −0.95 V to −1.1 V/SCE, where codeposition of NiFe alloys becomes more anomalous. Notably, in the same region of potentials the hydrogen partial current density, iH, is 2–3 times larger than the limiting current density measured for proton in solution at pH 3.0.40 The electrodeposition of Ni and Fe starts at the potential −0.7 V/SCE but with low current efficiency.

Figure 4.

Figure 4. Dependence of partial current densities for hydrogen (H) and metals (Fe, Ni) on potential of Pt-RDE (1 cm2) during potentiostatic electrodeposition; NiFe solution at pH 3.0; Rotation rate: 125 rpm.

Figure 5 shows dependence of the current efficiency and the hydrogen partial current density, iH, on the electrode potential of Pt-RDE. The current efficiency, η, was calculated from the partial current densities using formula: η = (iFe + iNi)/(iFe +iNi + iH) × 100 (%). The current efficiency increases from 5 to 80% in the range of the potentials from −0.7 to −1.2 V/SCE and decreases at the more negative potentials due to the reduction of H2O according to Eq. 7.

Equation ([7])

The produced OH base in reaction 12 is neutralized in acid-base reaction 11. The "parasitic" currents due to the reduction of H+ and H2O generally decrease current efficiency and increase hydrogen partial current density, iH. The rough potential distribution of the dominant hydrogen electroactive species- participating in the increase of hydrogen partial current density-are shown in Fig. 12. It is important to note that the reduction of the proton from solution on Pt electrode starts at potential ∼−0.5 V/SCE.30 It continues to be dominant together with proton from the reaction in the range of potentials from −0.7 V to −0.95 V/SCE. The proton liberated in reaction 6, H+react, is dominant hydrogen elctroactive species, together with FeOH+ and NiOH+, in the potential range of −1.0 V to −1.2 V/SCE.

Figure 5.

Figure 5. Effect of potential on current efficiency and hydrogen partial current density at Pt-RDE electrode (1 cm2) during potentiostatic electrodeposition; NiFe solution at pH 3.0; Rotation rate: 125 rpm.

Figure 12.

Figure 12. Current-time transient on Au-thin film electrode and Au-AAO template electrode run in NiFe solution at the controlled potential during 60 seconds.

Measurements of the surface area of different Au electrodes

The areas of three different Au electrodes, i.e. Au-thin film, Au-AAO template, and Au-electrodeposited into nanowire template at different time/length(Au-ED), were evaluated using K2Fe(CN)6 as a reversible one electron transfer oxidation (Eq. 8).

Equation ([8])

Figure 6 shows a reversible one electron cyclic voltammogram obtained at Au-thin film electrode at a 10 mV/s sweep rate. The cyclic voltammogram shows the peak separation, ΔEp = Epa – Epc = 60 mV and the peak ratio ipa/ipc = 1.0, which is close to the "ideal" reversible voltammogram.41 It is important to note that the results of ΔEp measured at the same sweep rate for Au-ED and Au-AOO template (not shown here) are somewhat higher than that obtained for Au thin film electrode and follow the order: Au-thin film > Au-ED > Au-AAO template. This might indicate the relatively slower electron transfer on Au-ED and Au-AAO template electrodes, compared to the sputtered Au-thin film electrode, possibly due to different crystalline structures and morphology of the Au surfaces. The anodic peak current, ip, is defined by the Randles-Sevcik equation41 (Eq. 9), which is proportional to the bulk concentration, Cb (mol/cm3), area, A (cm2), and square root of the sweep rate, ν1/2 (V/cm).

Equation ([9])

Taking the concentration of K4Fe(CN)6, Cb = 2.08 × 10−6 (mol/cm3), the number of electrons (n = 1), nF/RT = 38.92 (1/V), the Faraday constant (96484 C) and diffusion coefficient, D = 0.63 × 10−5 cm2/s,45 the area of Au-thin film electrode was calculated to be A = 0.754 cm2. The diameter of the O-ring opening is 2r = 1.0 cm, therefore, the calculated geometrical area is A = 0.785 cm2 and very close to the experimentally determined area of the Au-thin film electrode.

Figure 6.

Figure 6. Cyclic voltammogram of 2.08 mM K4Fe(CN)6 in 0.1 M Na2SO4 solution at Au-thin film electrode; Sweep rate: 10 mV/s.

In order to study surface area development during the nanowire electrodeposition and growth of Au we have used a cyanide-based gold solution. Figure 7 shows linear sweep voltammograms obtained at Au-thin film (1) and Au-AAO template electrodes (2) at the sweep rate of 20 mV/s. Two waves with the peaks observed around −0.8 V/SCE and −1.2 V/SCE were attributed to the reduction of adsorbed Au(CN)ads electroactive species at the less negative potential (Eq. 4)-formed in the preceding chemical reaction (Eq. 10,11)-and direct reduction of Au(CN)2 at the more negative potential (Eq. 12).42

Equation ([10])

Equation ([11])

Equation ([12])

The current ratio at the second peak potential obtained at both Au-AAO template electrode (curve 2) and Au-thin film electrode (curve 1), i.e 1.0/0.4 = 2.5, reflects the ratio of electrode area calculated from the Randles-Sevcik equation using K4Fe(CN)6 as a probe (see Fig. 9).

Figure 7.

Figure 7. Linear sweep voltammogram of 2.0 mM KAu(CN)2 at Au-thin film (curve 1) and Au-AAO template (curve 2) electrode; Sweep rate: 20 mV/s.

Figure 9.

Figure 9. (Left) The linear plot of ip1/2 curves obtained by one electron oxidation of Fe(CN)6−4 at Au-thin film, Au-ED, and Au-AAO template electrodes. (Right) The summary of results on measurements of electrode area on all three electrodes.

The electrodeposition of gold nanowires into the AAO-template at the controlled potential of −1.0 V/SCE was carried out every 20 min through 4 hours. After each deposition the formed Au-ED electrode was washed with water and transferred into potassium hexacyanoferrate solution to evaluate surface area. The results of first 20 min deposition produced a uniform 500 nm thick Au nanowire shown in Fig. 8-left. It is clearly visible a rough surface of sputtered 300 nm of Au at the alumina surface.

Figure 8.

Figure 8. (Left) The cross-sectional SEM image of 500 nm Au nanowire obtained after 20 minutes of deposition; (Right) The shape of the current-time curve recorded during 20 min of gold deposition at the controlled potential of −1.0 V/SCE.

The shape of current-time transient recorded during 20 minutes of gold deposition, shown in Fig. 8-right, is very close to the calculated transient from the model for the growth of nanowires published recently.42 The advancing diffusive front is characterized in this model by four stages of mass transport control: linear flux within the pores, localized hemispherical flux at the mouth of the pore, semi-infinite linear diffusion, and the last stage where fluxes inside and outside pores are in balance by the continuity condition giving rise to the current minimum. Notably, the time spent in the first two stages of mass transport is less than 2 seconds.42

The results of CV (Figure 9-left) obtained through one electron oxidation of Fe(CN)6−4 showed a linear plot of ip1/2 curve, which passes through the origin for all three kinds of Au electrodes. The current function, ip1/2, is constant with the increase of the sweep rate. Therefore, both observations suggest that electrode reaction is controlled by diffusion. From the obtained slopes-proportional to the electrode area- all the areas were calculated using Randles-Sevcik equation. The results are summarized in Fig. 9-right. The area for Au-thin film is A = 0.785 cm2, for Au-AAO template is A = 1.91 cm2 or 2.43 times larger than area of Au-thin film electrode, for Au-ED the area is decreasing from A = 1.42 cm2 (after 20 min of deposition) to A = 0.82 cm2 (after 240 min of deposition). A similar increase of the electrode area for ∼2.3 times for Au-AAO template electrode compared to the geometrical area of the cross-section of the porous AAO template with dw = 200 nm was found recently.43 There are two main reasons for the increase of electrochemical area of Au-AAO template electrode. First, the sputtering of 300 nm of Au was found enough to completely close the pore with dw = 200 nm, preventing the leaking of the solution to the Cu-stage. The deposition of Au along the sides of the pores brings about a hemispherical shape of the Au electrode.43 Second, is the increase of the roughness factor by increase of Au thickness, which is defined here as a ratio of rms- roughness of sputtered 50 nm Au thin film to the rms-roughness of the 300 nm Au thin film. Figure 10 shows the images of thin film surfaces of Au sputtered on alumina. It is obvious that grain size on the surface of the 300 nm Au film is higher than on 50 nm Au film. The AFM rms-rougness (5 × 5 μm) measured on the 50 nm Au film is 1.5 nm and on 300 nm is 5.8 nm. Thus, the roughness factor is 3.8. The results obtained on measured electrochemical surface area-which changes with growth of nanowire inside the pores-suggest that the electrodeposition using constant current would produce the composition change along the NiFe nanowire length during the growth due to the changes in current density.

Figure 10.

Figure 10. SEM images of thin film surfaces of sputtered gold on alumina. (Left) 50 nm Au-thin film with rms-roughness of 1.5 nm, (Right) 300 nm of Au-thin film with rms-roughness of 5.8 nm.

Comparison of electrochemical behavior of NiFe solution on Au-thin film and Au-AAO template electrodes

Two cyclic voltammograms for NiFe solution obtained on Au-thin films and Au-AAO template electrodes, respectively-in the range of potentials from +0.4 V to −1.4 V/SCE at the sweep rate of 20 mV/s-are overlaid in Fig. 11-left. Both CV curves showed the same characteristic regions (a-c), which were already discussed for the case of Au-thin film electrode shown in Fig. 2. The major difference between two CV voltammograms is the appearance of a reduction new peak at −0.19 V/SCE, which is a product of the oxidation at the starting anodic potential of +0.4 V/SCE. In the double-sweep cyclic voltammogram (Fig. 11-right) run from +0.4 V/SCE to −0.45 V/SCE, the peak at −0.19 V appeared which on returned sweep produce chemically reversible peak at +0.08 V/SCE. The chemical nature of this redox couple on Au-AAO template electrode is not known at the present time. In region (a) the reduction of H+ diffusing from the solution starts at the less negative potential in the case of Au-AAO template electrode. In the region (b)-i.e. reduction and deposition of Ni+2 and Fe+2 electroactive species-the cathodic current in the case of Au-AAO template electrode is higher than the current obtained for Au-thin film electrode, due to the larger surface area at Au-AAO template.

Figure 11.

Figure 11. (Left) Cyclic voltammetry curves on Au-thin film electrode (black) and Au-AAO template electrode (red) in NiFe solution at 20 mV/s sweep rate, (Right) Double-sweep cyclic voltammograms run from +0.4 V/SCE to −0.45 V/SCE on Au-AAO template electrode at 20 mV/s sweep rate.

Figure 12 shows the chronoamperometry curves for electrodeposition of NiFe on Au-thin film and Au-AAO template electrodes at two controlled potentials during 60 s, i.e. at −1.1 and −1.4 V/SCE, respectively. It is shown that the cathodic steady-state current obtained at the potential of −1.1 V/SCE is 2.4–3.0 times higher for Au-AAO template electrode due to the larger surface area. The electrodeposition of NiFe at −1.4 V/SCE into the AAO template gave higher current which decreased after 10–20 s by ∼30%. Importantly, the potential region between −1.2 V and −1.4 V/SCE is a region where H2O is reduced producing H2 bubbles which block the surface area of the growing nanowires, giving rise to the lower current and generally nonuniformity in the nanowire length.

Effect of potential on composition, morphology and crystal structure of NiFe nanowires

Figure 13 shows a typical potential-dependent "volcano" curve with a peak in Fe-content at −1.1 V/SCE for NiFe nanowires deposited in quiescent solution on Au-thin film electrode (Fig. 13-left) and Au-AAO template electrode (Fig. 3-right) using a three electrode cell described in Fig. 1. The "selectivity ratio" increases to the maximum value of ∼13-with increase of negative potential from −0.9 V/SCE to −1.1 V/SCE- indicating more anomalous codeposition in this potential range. Note that the potential at which a maximum in Fe-content in NiFe is obtained is less negative in quiescent solution (Fig. 13) compared to that obtained in agitated solution (Fig. 3), which is in agreement with the previous studies32 on the effect of agitation in anomalous codeposition. The selectivity ratio in NiFe nanowires (Fig. 13-right) decreases to the value ∼1.0 in the range of more negative potentials where codeposition of NiFe alloys becomes non-anomalous.

Figure 13.

Figure 13. Effect of potential on composition and selectivity ratio obtained in the quiescent NiFe solution at pH 3.0. (Left) NiFe films deposited on Au-thin film electrode. (Right) NiFe nanowire arrays deposited on Au-AAO template electrode.

The stable composition of Ni56Fe44 nanowire arrays with lengths of 3.0, 5.0 and 9.0 μm, respectively was obtained (Fig. 14-left, open circles). The composition along the length of 9.0 μm nanowire (Fig. 14-left, filled circles) was in the range of ±1% of the mean composition of Ni56Fe44. The Ni56Fe44 nanowires with a length of 9.0 μm shown in Fig. 14-right were uniformly deposited into Au-AAO template. For magnetic studies we used NiFe nanowires with the same length (L ∼ 3 μm) and the nanowire diameter (dw = 200 nm). An example of uniformly deposited Ni79Fe21 nanowire with L ∼ 3.0 μm is shown in Fig. 15-right with a focus on a single nanowire with dw ∼ 200 nm shown in Fig. 15-left, which had grown with a convex shape.

Figure 14.

Figure 14. (Left) The dependence of composition of NiFe nanowire arrays with the length 3.0, 5.0 and 9.0 μm (open circles) and along the length of the single nanowire (filled circles). (Right) The cross-sectional SEM image of Ni56Fe44 nanowire arrays with the length of 9.0 μm.

Figure 15.

Figure 15. Cross-section SEM image of Ni79Fe21 nanowire arrays with L ∼ 3.0 μm (Right), and with a focus on a single nanowire with dw = 210 nm (Left).

The crystalline structure of electrodeposited NiFe nanowire arrays into AAO template was determined by XRD measurements. As expected, the NiFe nanowires have fcc structure shown in Fig. 16. The NiFe (111) and (200) diffraction peaks were present in all electrodeposited NiFe nanowires with 5–55%Fe. The average grain size was determined to be 9–11 nm based on (111) diffraction peak using the Scherer's equation, which is close to the value found in Ni-rich NiFe nanowires.16 The Au (111) diffraction peak shown in Fig. 16 is from the Au-seed layer deposited at the bottom of AAO template.

Figure 16.

Figure 16. XRD patterns of NiFe nanowires witgh 5–55 At.% Fe.

Magnetic properties of NiFe nanowire arrays with 8–55%Fe

Figure 17 shows the normalized hysteresis loop for the NiFe nanowire arrays with the same length (L∼3 μm), diameter (dw = 200 nm), interwire distance (D = 250 nm) and diameter/interwire distance ratio of 0.8 as a function of their composition. All of the hysteresis curves are tilted with a low sureness (Mr/Ms). When the value of magnetic field is high enough to overcome the effective magnetic anisotropy the magnetization of nanowire arrays switches to its saturation value. A similar behavior was observed for Ni79Fe21 nanowire arrays with dw = 70 nm and ratio of diameter/interwire distance of 0.7.18

Figure 17.

Figure 17. Normalized hysteresis loops of NiFe nanowire arrays with L ∼3.0 μm, dw = 200 nm, and dw/D = 0.8 measured in parallel and perpendicular directions to the wire axis.

The values of coercivities (Hc), and squareness (Mr/Ms), when field is applied parallel and perpendicular to the nanowire axes are given in Table I. The parallel coercivity, Hc||, is higher than perpendicular, Hc⊥, for Ni92Fe8 and Ni79Fe21 alloys, but it is smaller for Ni60Fe40, Ni44Fe56 and Ni45Fe55 alloys. The observed sureness (Mr/Ms < 0.1) is close to zero revealing better orientation along the wire axis. We will discuss the two sets of magnetic behavior of NiFe nanowires depending on composition (Ni92Fe8 and Ni79Fe21 vs. Ni60Fe40, Ni56Fe44, and Ni45Fe55) including the difference of coercivities, ΔHc = Hc, parallel – Hc, perpendicular, and the process of magnetization reversal.

Table I. The effect of composition of NiFe nanowire arrays on coercivity and squreness.

Composition (At. %) Parallel coercivity (Oe) Perpendicular coercivity (Oe) Squereness, Parallel (Mr/Ms) Squereness Perpendicular (Mr/Ms)
Ni92Fe8 95.7 13.7 0.062 0.063
Ni79Fe21 92.7 43.6 0.051 0.109
Ni60Fe40 73.3 132.8 0.019 0.061
Ni56Fe44 66.5 116.8 0.051 0.109
Ni45Fe55 53.7 75.5 0.024 0.030

The effective anisotropy energy, Keff, (Eq. 13) is determined by shape anisotropy (Ksh = 6 × 106 erg/cm3)45 and magnetostatic anisotropy (Kms) that overcome magnetocrystalline (Kmc = −7 to 5 × 103 erg/cm3)1,46 and magnetoelastic anisotropy (Kme = 4.3 to 75 × 103 erg/cm3) (see Table II)

Equation ([13])

The coercivity is a quantitative measure of magnetic field required to reverse the magnetization direction in the thin film or nanowire arrays. The impact of two dominant anisotropy energies (Ksh and Kms) in nanowire arrays is the best illustrated if one compare the coercivities obtained in thin films in which the shape anisotropy is zero.48 Notably, the thin films of 500 nm NiFe with the same composition as two samples of electrodeposited nanowire arrays, i.e., Ni79Fe21 and Ni45Fe55 showed easy axis coercivities as 0.8 and 2.1 Oe, respectively. The easy axis coercivity, Hce,, obtained with the thin films is about 25–120 times lower than the easy axis coercivity, Hc, parallel, in nanowire arrays. It is important to note, that in the studies of magnetic properties of NiFe nanowires1320 a magnetocrystalline anisotropy was taken into consideration as very small or "negligible"-which is true if one compares with the values of Ksh and Kmc. However, although the Kme is larger than Kmc, the magnetoelastic, Kme, anisotropy of NiFe nanowires-obtained and measured at the room temperature-was completely ignored.

Table II. Some properties of NiFe nanowires arrays with 8–55 At. % Fe.

  Ni92Fe8 Ni79Fe21 Ni60Fe40 Ni56Fe44 Ni45Fe55 Units
Ms 558 813 1156 1196 1276 emu cm−3
2π Ms 3504 5105 7460 7510 8013 Oe
σ 6.7 100 192 190 200 MPa
λs × 10−6 −20 3.0 16.0 21.0 25.0  
Kme −1,34 × 103 4.5 × 103 4.6 × 104 5.9 × 104 7.5 × 104 erg cm−3
Kmc - −2 × 103 9 × 103 1.2 × 104 2.3 × 104 erg cm−3
Hms* 3013 4390 6415 6458 6891 Oe
Heff** 3022 4402 6152 6425 6842 Oe
ΔHc*** 82.0 49.1 −59.3 −50.3 −21.8 Oe
Hc|| /Hms 0.0316 0.021 0.0116 0.010 0.0077  
ΔHc/Heff 0.271 0.011 −0.00963 −0.00783 −0.00318  

*Eq. 14 **Eq. 16 ***ΔHc = Hc|| -Hc⊥

In Figure 18 are presented results of stress and magnetostriction measurements obtained for 500 nm thin films of NiFe alloys with 5–95%Fe prepared by galvanostatic electrodeposition (5 mA/cm2) under the applied magnetic field in order to induce in-plane anisotropy, using acidic solution (pH 2.8).25 Notably, saccharin was added into the plating solution as a stress reliever.

Figure 18.

Figure 18. Effect of Fe-content in 500 nm thin films of NiFe prepared by galvanostatic electrodeposition (5 mA/cm2) in acidic NiFe solution at pH 2.8 on stress (Left), and magnetostriction (Right).

The measured stress of fcc NiFe films with 5–55 At. % Fe increases with the increase of Fe-content in deposit, but is relatively low in the range from compressive −50 MPa to +200 MPa tensile stresses, which is typical for NiFe films deposited in the presence of saccharin as a stress relieving agent. The electrodeposition of the same films without the presence of saccharin could result in 2–3 times higher stresses.49 The linear increase of tensile stress in fcc NiFe electrodeposited films was observed earlier.23,50 Our results (Fig. 18) show the linear increase of tensile stress in both fcc and bcc NiFe alloys. We believe that the stress mechanism in these NiFe films could be due to adsorption/desorption of hydrogen similar to the described earlier in the case of CoFe,37 i.e. through incorporation of hydrogen as a metal hydride at the grain boundaries, and subsequent recombination and diffusion out of hydrogen molecule leaving a grain vacancies at the boundaries. In the range of fcc/bcc and metastable (MS) NiFe alloys-in the so called Invar region25 - other stress mechanisms possibly operate.

The measured values for magnetostriction, λs, agree, within error with the literature values.1,47 The accepted value for λs = 0 is around Ni80Fe20 elemental composition. Our results show that the zero magnetostriction is obtained with the Ni81Fe19 elemental composition.. At the composition of NiFe film with Fe < 19%, magnetostriction is negative, above 19% of Fe in NiFe magnetostriction increases monotonously to the positive direction. Therefore, the calculated magnetoelastic anisotropy energy shown in Table II might be expected to exceed even more magnetocrystalline anisotropy of NiFe alloys with higher Fe-content and/or without the presence of saccharin in plating solution as a stress relieving agent.

Figure 19-left shows the decrease of the parallel coercivity, Hc||, with the increase of Fe-content in NiFe nanowires. The decrease of Hc|| correlates with the increase of the magnetic saturation, Bs, which follows the Slater-Pauling curve.25 This can be attributed to the combined effect of shape anisotropy and magnetostatic anisotropy between the nanowires.45 For the arrays of nanowirws the effective anisotropy containing only the magnetostatic contribution is given by Equation 14.45

Equation ([14])

where Ms is the magnetic saturation and P is the packing factor or porosity given by r2/3.5D2, which depends on diameter of nanowire, dw = 2r, and the interwire distance, D. Thus, the increase in Hms is due to the increase of saturation magnetization or due to the increase of Fe-content in NiFe alloys. The increase of Hms enhances the stray field in nanowire arrays-resulting in lower Hc||. This can be also achieved by increasing the length of NiFe nanowire alloy, L, with the same composition or the same magnetic saturation value according to the Equation 15,51 due to the net contribution from the perpendicularly aligned dipole field, Hd, which reduces the effective anisotropy. The results in Fig. 19-right illustrate the effect of length (3 vs. 9 μm) of Ni40Fe60 nanowire arrays on coercivity measured at different angles.

Equation ([15])
Figure 19.

Figure 19. (Left) Effect of Fe-content in NiFe nanowires on parallel coercivty and magnetic saturation measured in thin films. (Right) Parallel coercivity dependence on the length of Ni60Fe40 nanowire arrays (3.0 vs. 9.0 μm) measured at different angles.

Our results clearly show the decrease of coercivity, Hc||, as a function of Fe-content (5–55%) in NiFe nanowire arrays electrodeposited in AAO templates with large diameter (dw = 200 nm) and dw/D ratio is 0.8. This is somewhat different from the results in the literature, obtained with NiFe nanowire arrays containing 15–35%Fe deposited in AAO templates with smaller diameters (d = 35–41 nm) and dw/D ratio ∼0.4, which showed small increase of coercivity, Hc||,51 or no significant dependence of coercivity15 as a function of alloy composition. The values of Hc||, obtained with NiFe nanowires deposited in AAO templates with small diameters (dw = 35–41 nm) are much higher15,51 than that obtained in this work (dw = 200 nm). The decrease of coercivity with increase of nanowire diameter is a general trend for the various ferromagnetic materials52 indicating that the shape anisotropy is dominant.

Figure 20-left shows dependence of coercivity, Hc, measured parallel and perpendicular to the nanowire axis as a function of Fe-content in NiFe. The difference ΔHc = Hc||, -Hc⊥ is positive for Ni92Fe8 and Ni79Fe21 alloys and negative for other three NiFe alloys with Fe > 30%. At the Fe-content in NiFe around the composition of 30 At. % Fe there is a crossing of parallel and perpendicular curves where ΔHc = 0. Since the magnetocrystalline, Hmc, and magnetoelastic, Mme, anisotropies are acting in-plane direction, or perpendicular to the nanowire axis, the effective anisotropy can be expressed as:

Equation ([16])

The positive ΔHc difference is observed with NiFe nanowires having a "negligible" both Hmc and Hme anisotropies. The negative ΔHc was observed with the NiFe nanowires, having a higher Ms-values and increasing Hmc and Hme anisotropy fields, which are acting simultaneously on both shape, Hsh, and magnetostatic, Hms, anisotropy fields giving rise to the decrease of easy axis coercivity, Hc||, in nanowire arrays. Figure 20-right shows decrease of ΔHc, as a function of Fe-content in NiFe nanowire arrays reaching the minimum at 40% Fe and increase followed by increase of Fe content in NiFe alloys.

Figure 20.

Figure 20. Effect of Fe-content in NiFe nanowires on parallel and perpendicular coercivity (Hc) (Left), and on ΔHc = Hc∥. –Hc⊥. (Right).

The comparison of some properties of thin films and nanowires of NiFe alloys with 8–55 At. % Fe are given in Table II. The results demonstrate that the magnetism of nanowires is dominated by shape and magnetostatic anisotropy. The shape anisotropy for an infinite cylinder is 2π Ms,12 where Ms is the saturation magnetization determined by measurements on NiFe thin films.25 The corresponding values for Hms and Heff were calculated using the Equations 14 and 16, respectively. The values for Hmc (not shown) were calculated using literature data for Kmc, while Hme values (not shown) were calculated from Kme based on experimental data presented in Fig. 18 and Table II. The parallel coercivity, Hc||, is only the small fraction fraction of Hms. The ratio Hc||,/Hms decreases with increase of Ms value or Fe-content in NiFe alloys. The ratio ΔHc/Heff vs. Fe-content in NiFe exhibit the same shape as the ΔHc vs. Fe-content in NiFe curve shown in Fig. 20-right.

The process of magnetization reversal in NiFe nanowire arrays was investigated. We have obtained two sets of behavior depending on the composition of NiFe nanowires. In the case of Ni92Fe8 and Ni79Fe21 nanowire coercivities decrease in the 0–90° and increase in 90–180° angular range (Fig. 21). In the case of Ni40Fe60, Ni56Fe44, and Ni45Fe55, the coercivities increase in the 0–90° and decrease in 90–180° angular range.

Figure 21.

Figure 21. Dependence of parallel coercivity on the composition of NiFe nanowires with 8–55 At. %Fe measured at different angles.

Two of the most common magnetization reversal mechanisms are coherent rotation (CR) and curling mode (C).53 For a specific magnetic material, the critical diameter, dcr,-given by Equation 17 where A is the exchange stiffness and Ms is magnetic saturation54 - has been considered as a criteria for transition between CR and C mode.

Equation ([17])

For the small nanowire diameters (dw < dcr) the preferred mode is coherent rotation (CR), while for dw > dcr the curling mode (C) is expected.54 By assuming A = 10−6 erg/cm and taking Ms-values in emu/cm3 from Table II-the critical diameters for Ni92Fe8, Ni79Fe21 Ni40Fe60, Ni56Fe44, and Ni45Fe55 nanowires were calculated as: 37.2, 25.6, 18.0, 17.4 and 16.3 nm, respectively. Since dw > dcr is characteristic for all of NiFe nanowires studied one could expect that C-mode would be a dominant reversal mechanism. We have observed experimentally an angular dependence of Ni92Fe8 and Ni79Fe21 nanowire coercivity, with the lowest value at the 90° angle, which theoretically corresponds to the curling reversal mechanism (C-mode).54 For Ni40Fe60, Ni56Fe44, and Ni45Fe55 nanowire coercivity, we have observed the highest value at the 90° angle, which theoretically corresponds to the coherent rotation (CR).55

Conclusions

In this study we have measured the electrochemical area on three Au electrodes (sputtered thin film, electrodeposited gold into nanowires, and AAO template). They were found to be about 1–2.4 time larger compared to the smooth Au-thin film electrode. These results suggested that the electrodeposition using a constant current would produce composition changes along NiFe nanowire due to the changes in current density. Ordered fcc NiFe nanowires with 5–55% Fe and with the uniform length and composition along the nanowire length were obtained using designed potential pulse electrodeposition.

The anomalous codeposition of NiFe nanowire resulted with a typical "volcano" type curve with a peak in Fe-content as a function of applied potential. Similar results observed in the literature were explained through the limited mass transport of Fe+2 ions after the peak. This explanation is partially correct, but not complete. An alternative explanation of anomalous codeposition was proposed by the increased surface concentration of H+-in a modified BDD mechanism-and its preferred catalytic reaction with NiOH+ads species which is thermodynamically favored to dissociate into solution while FeOH+ads accumulates at the surface. The rate of discharge of adsorbed FeOH+ and NiOH+ ions are proportional to their respective surface concentrations.

Our results clearly show a decrease of coercivity, Hc||, as a function of Fe-content (5–55%) in NiFe nanowire arrays electrodeposited in AAO templates with L∼3.0 μm, dw = 200 nm and dw/D = 0.8, with the increase of Fe-content in NiFe nanowire. The decrease of Hc|| correlates with the increase of the magnetic saturation, Bs, which follows the Slater-Pauling curve. Our results demonstrate that the effective anisotropy is dominated by shape and magnetostatic anisotropy, while the magnetocrystalline and magnetoelstic anisotrpopy are only a small fractions of the shape anisotropy.

We have found two sets of magnetic behavior of NiFe nanowires depending on composition (Ni92Fe8 and Ni79Fe21 vs. Ni60Fe40, Ni56Fe44, and Ni45Fe55). The parallel coercivity, Hc||, is higher than perpendicular, Hc⊥, for Ni92Fe8 and Ni79Fe21 alloys, but it is smaller for Ni60Fe40, Ni44Fe56 and Ni45Fe55 alloys. Also, we have observed angular dependence of Ni92Fe8 and Ni79Fe21 nanowire coercivity with the lowest value at a 90° angle which theoretically corresponds to the curling reversal mechanism. For Ni40Fe60, Ni56Fe44, and Ni45Fe55 nanowire coercivity, we have observed the highest value at a 90° angle, which theoretically corresponds to the coherent rotation.

Acknowledgments

We thank Dr. Steve Riemer (Seagate Technology, Bloomington, USA) for helpful discussions and Marieta Porcescu (National Institute of R&D for Technical Physics, Iasi, Romania) for help in the experimental work.

This work was supported by European Commission through FP7-REGPOT-2012-2013-1 NANOSENS project (grant Agreement No. 316194).

Please wait… references are loading.
10.1149/2.0771603jes