Tabakovic 2018

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 36

Accepted Manuscript

Research articles

Preparation of metastable CoFeNi alloys with ultra-high magnetic saturation


(Bs = 2.4 – 2.59 T) by reverse pulse electrodeposition

Ibro Tabakovic, Venkatram Venkatasamy

PII: S0304-8853(17)32976-1
DOI: https://doi.org/10.1016/j.jmmm.2017.12.003
Reference: MAGMA 63471

To appear in: Journal of Magnetism and Magnetic Materials

Received Date: 18 September 2017


Revised Date: 1 November 2017
Accepted Date: 1 December 2017

Please cite this article as: I. Tabakovic, V. Venkatasamy, Preparation of metastable CoFeNi alloys with ultra-high
magnetic saturation (Bs = 2.4 – 2.59 T) by reverse pulse electrodeposition, Journal of Magnetism and Magnetic
Materials (2017), doi: https://doi.org/10.1016/j.jmmm.2017.12.003

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Preparation of metastable CoFeNi alloys with ultra-high magnetic
saturation (Bs = 2.4 – 2.59 T) by reverse pulse electrodeposition

Ibro Tabakovic*z (a) and Venkatram Venkatasamy, ** (b)


(a)
ECE Department, University of Minnesota, Minneapolis, Minnesota 55435
(b)
Seagate Technology, Research and Technology Development
7801 Computer Avenue South, Bloomington, Minnesota 55435

(EP) a 1
Abstract.
The results of reverse pulse electrodeposition of CoFeNi films with ultra-high magnetic
saturation, i.e. Bs values between 2.4-2.59 T, are presented in this work. Based on valence-bond
theory (Hund’s rule) it was assumed that the electronic configuration of MOH obtained by one
electron reduction of electroactive intermediate (MOH+ads + e MOHads) or oxidation of metal
(M –e + HOH  MOH + H+) would result with larger number of spins per atom for each of
transition metals in MOH-precipitated in CoFeNi deposit- with one more spin than their
respective neutral metal in the order: Fe > Co > Ni. The experimental results showed that the
increase of Bs value above Slater-Pauling curve was not observed for CoFe alloys, thus FeOH
and CoOH compounds were not present in deposit. However, the increase of the Bs values above
the Slater-Pauling curve (Bs = 2.4-2.59 T) was observed, for CoFeNi films obtained by reverse
pulse electrodeposition. Therefore, NiOH as a stable compound is probably formed in a one-
electron oxidation step during anodic pulse oxidation reaction precipitated presumably at the
grain boundaries, giving rise to the ultra-high magnetic saturation of CoFeNi films. The effects
of experimental conditions on elemental composition, magnetic properties, crystal structure, and
thermal stability of CoFeNi films were studied.

z*
E-mail: tabakovicibro@gmail.com
*Electrochemical Society Member
** New address: Ericsson India Global Services.
**E-mail: venkatram.venkatasamy@ericsson.com

(EP) a 2
1. Introduction

The electrodeposition (ED) process has been used for more than five decades in the
industrial production of soft magnetic materials for recording devices and as thin magnetic films
for micrroelectromechanical systems (MEMS) devices [1, 2]. The use of pole materials with high
magnetic saturation (Bs) in writer recording heads is a prerequisite for achieving adequate write-
ability with high areal density in longitudinal (LMR), perpendicular (PMR), and heat-assisted
magnetic recording (HAMR).
The major inherent advantage of ED over sputtering (SP) deposition is in fabrication of
the patterns with high aspect ratio by deposition “through a photolithographic mask”, which
leads to the reduced process content, reduced variance, and therefore reduced cost. The write
pole materials –with increasing Bs values-were introduced into the recording industry for the
first time as longitudinal poles in the sequence: 1.0 T Ni80Fe20 (IBM), 1.6T Ni50Fe50 (IBM), and
1.8T Co64Fe23Ni14 (Seagate), and 2.4T CoFe (Seagate). The increase of magnetic saturation
correlates with areal density (AD), which is a product of bits per inch and tracks per inch at the
inner track of the recording band. Importantly, AD also correlates with the growth of revenue.
The CoFe magnetic alloys with 50-70 % Fe have magnetic saturation: Bs 2.4 T [4]. The
highest magnetic saturation of all transition-metal alloys at room temperature is for Co35Fe65
alloy with Bs = 2.45T, which is communally referred to us as the ‘Slater-Pauling limit” [5].
The development of materials with Bs value higher than 2.45 T has stimulated many new
studies. In 1970, Kim and Takahashi [6] reported a successful fabrication of Fe16N2 film with Bs
of 2.83 T. However, other research groups were unable to reproduce the high Bs of 2.83T, and
these claims have been disputed [7-10]. Recently, J.-P Wang group reinvigorated the old subject
demonstrating that a post annealed FeN film at 150oC/20 hours; possess high magnetic saturation
between 2.6-2.68 T [11, 12]. Generally, according to the localized electron theory, the increase
of magnetization (µB, Bohr magneton per atom) results from the increase of the number of d-

(EP) a 3
electrons [13]. For example, the electronic configuration of Fe is [Ar]3d64s2 (4 µB), for Co is
[Ar]3d74s2 (3µB), and for Ni is [Ar]3d84s2 (2 µB). Since the magnetization per atom is almost
entirely due to the spin, the Bohr magnetons per atom (B) should be an integer. However,
localized electron theory cannot explain the fact that-according to the itinerant/band electron
theory-the observed B are non-integer for atoms of transition metals, i.e. 2.2 B for Fe, 1.72 B
for Co, and 0.61 B for Ni. The property, in which the delocalized electrons can move from atom
to atom, is known as itinerant behavior. Notably, the enhanced magnetic moment was observed
for free Fe, Co, and Ni nanoclusters for containing less than 100 atoms, i.e. about 3 B for Fe,
about 2 B for Co, and about 1 B for Ni [14-16]. High proportion of surface atoms over bulk
atoms give a reduced average coordination, i.e. less 3d unpaired spins of surface atoms are used
in the formation of metal bonds [17, 18].
It has been observed in scientific [19-23] and patent literature [24] that electrodeposition
of CoFeNi films results in about 10-30% larger Bs values than those reported for bulk alloys of
the same composition [4]. The increase in Bs value was generally attributed to structural effects
in deposited films.
In this work we present results of the reverse pulse electrodeposition of CoFeNi films
with ultra high magnetic saturation, i.e. Bs values between 2.4-2.59 T. We have assumed that
larger number of spins per atom was obtained through one-electron reduction of the MOH+
electroactive species during cathodic pulse, and one- electron oxidation of deposited metal, M,
during reverse anodic oxidation. Both electrode reactions could produce a monovalent neutral
hydroxide in deposit, MOH, where M = Fe, Co, Ni. Based on valence-bond theory (Hund’s rule)
the electronic configuration of FeOH is [Ar]3d64s1 (B =5.0), for CoOH is [Ar]3d74s1 (B = 4.0),
and for NiOH is [Ar]3d84s1 (B =3.0). Therefore, each of transition metals in MOH possess one
more spin than their respective neutral metal in the order: Fe > Co > Ni.

2. Experimental

The CoFe and CoFeNi films were electrodeposited on 8 inch round alumina coated

AlTiC wafers using 2000 Å sputtered copper seed layers (Cu substrate). The CoFe films were

(EP) a 4
prepared from a CoFe-plating solution containing: H3BO3 (25 g/l), NH4Cl (16 g/l), FeSO4 7H2O

(20 g/l), Na-lauryl sulfate (0.05 g/l) by gradually adding CoSO4 7H20 (0 – 15 g/l). The CoFeNi

films were prepared from a CoFeNi-plating solution containing: H3BO3 (25 g/l), NH4Cl (16 g/l),

FeSO4 7H2O (20 g/l), CoSO4 7H2O (15 g/l), Na-lauryl sulfate (0.05 g/l) by gradually adding

NiSO4 6H20 (5 – 70 g/l). The solution pH was adjusted to 2.3 by adding HCl. Notably; no other

organic additives were used in reverse pulse electrodeposition. The reversed pulse

electrodeposition was carried in this work employing a characteristic on/off profile shown in

Fig.1. The magnitude of cathodic pulse current density ip, pulse anodic/reverse current density ir,

time-on ton, and time-off toff, were optimized experimentally. (See section 3.2). The

Electrochemical Quartz Crystal Microbalance (EQCM) measurements, used in this work for

optimization of the pulse parameters, were the same as described in our previous work [25]

The electrodeposition of CoFe and CoFeNi films were achieved using a semi-automated

industrial tool (Tosetz Co.) having a paddle configuration with a 1000 Oe external magnetic field

to set the easy axis direction in the film. The agitation during electrodeposition was carried out

by using a reciprocating paddle with solution flow rate of 150 mm/s. Total volume of electrolyte

solution was 100 liters. Film thickness was determined by X-ray fluorescence (XRF), taking an

average thickness from 15 point measurement distributed over the entire wafer surface and cross-

checked by a DekTek profilometry ( 9 points). The thickness uniformity is extremely important

in measuring the Bs value, since the major error comes from the average thickness value. (See

text below). The thickness uniformity of films defined as a /mean x 100 was optimized to <

2%, which practically defines the uncertainty in thickness measurements. This was achieved

through tedious optimization of the following tool parameters: Ni-anode height, paddle height,

solution flow rate, and temperature (23oC). The most important parameter in archiving a low

(EP) a 5
thickness uniformity was the current ratio of working electrode (cathode), Iw, vs. holder

(auxiliary electrode), Ih (Iw/Ih = 1.43).

Magnetic properties of the films were determined using a hysteresis loop tracer (BH

looper) and a vibrating sample magnetometer (VSM). The BH looper was calibrated with ~500

nm thick CoFe film prepared by sputtering from Co35Fe65 alloy target deposited on Cu substrate.

Typical magnetization vs. magnetic field (m-H hysteresis curve) is shown in Fig. 2a. Both easy,

Hc, e and hard, Hc, h axis coertivities look alike, i.e. Hc, e = 30.39 Oe, Hc, h = 21.10 Oe, and ms =

227 nW. The main reason for poor anisotropy of these films could be due to the large

magnetostriction of SP Co35Fe65 films (~7.5 x 10-5) [4] combined with measured stress (

~700 MPa). The high applied external magnetic field (H = 2500 Oe)-also used in similar

measurements for ED CoFe and CoFeNi films- was necessary to achieve a full magnetic

saturation. Figure 2b illustrates the effect of the external field, H, on measured Bs value of 0.37

m thick SP Co35Fe65 film. The magnetic saturation Bs in Wb/m2 (or inTesla) was calculated

using the formula Bs = nW/ (t x d), where d is the diameter of deposited wafer (d = 20 cm) and t

is the thickness in microns. For the measurements of ED CoFe and CoFeNi films d = 19.4 cm

was used, due to the edge exclusion area.

The elemental composition of CoFe and CoFeNi films was determined by inductively

coupled plasma optical emission spectroscopy (ICP-OES) using a Teledyne Leeman Labs

“Prodigy” ICP spectrometer. The small areas at the deposited films were dissolved into

approximately 5 ml solution of 50% (v/v) nitric acid and 1% (v/v) hydrochloric acid. The

solutions were then diluted as necessary for analysis into volumetric flask. The chemical state of

elements was determined using X-ray photoelectron spectroscopy (XPS), performed by Charls

Evans & Associates laboratory and by laboratory of National Institute for R&D in Technical

(EP) a 6
Physics, Iasi, Romania. Films were sputtered for 12 min using 1 kV Ar+ beam and 5 min using

500 eV beam. Sputtering removed about 30-40 nmof material from the surface. The samples

were tilted and analyzed at the take-off angle of 75o.

The morphology was studied using scanning electron microscopy (SEM). The crystal
structure was studied by X-ray diffraction (XRD) and transmission electron microscopy (TEM).
Film stress was calculated from deflection using the Stony equation.

3. Results and Discussion

3.1. Reverse pulse electrodeposition of CoFeNi alloys. Theoretical considerations


The basic idea in this work was to improve Bs value of CoFe and CoFeNi alloys more
than 2.45 T by inclusion of one-electron intermediate MOH product into the deposit.
Theoretically, MOH intermediates could be formed by reduction during the cathodic pulse or by
oxidation during the anodic pulse, in reverse pulse electrodeposition. Here we analyze which of
these two steps could be more favorable for precipitation of MOH into the deposit.
The widely accepted mechanism of reduction of ions of transition elements, M+2 (M = Fe,
Co, Ni) is Bockris-Drazic-Despic (BDD) mechanism [25] described in reactions 1-3.

M+2 + HOH = MOH+ads + H+ (1)


MOH+ads + e  MOHads (2)
+ +
MOH ads + e + H  M + H2O (3)

The key in BDD-mechanism of overall two-electron reduction (M+2 + 2e  M) is that the


electroactive species is not M+2 but MOH+ formed at the electrode surface through diffusion and
hydrolysis of M+2 ion at the electrode surface in reaction 1. This was surprising discovery since
the calculated concentrations of MOH+ species are negligible in acidic solutions [30]. Therefore,
these species must be formed at the electrode surface. The backward reaction (1) is a dissociation
reaction catalyzed by liberated H+ in reaction: MOH+ads + H+  M+2 + H2O. There is a
competition for the protonation between adsorbed MOH+ species in the following order of their
dissociation constants: FeOH+ > CoOH+ > NiOH+. For example due to ~1000 times higher
dissociation constant of NiOH+ (pKa = 4.3) than FeOH+ (pKa = 7.2) [26], NiOH+ is

(EP) a 7
thermodynamically more favored to dissociate into solution while FeOH+ accumulates at the
surface. This explained the kinetic phenomena of anomalous codeposition of NiFe and CoFeNi
alloys. The adsorbed MOH+ species is reduced in the rate determining step [25] in reaction 2,
which is one-electron reduction leading to the formation of neutral monovalent hydroxide, MOH.
The last step (reaction 3) is a fast second electron transfer, which is facilitated by proton
liberated in reaction (1).
We have modified the BDD-mechanism, which has been used to explain the following
experimental facts like: (i) decrease of current efficiency in CoFe alloys with the increase of Fe
content in deposit [27], (ii) the increase of tensile stress of CoFe alloys with the increase of Fe
content in deposit [28] , (iii)the composition gradient in thin CoNiFe [29] and NiFe films [30] as
a special case of the anomalous codeposition, and (iv) anomalous codeposition of NiFe films as a
function of electrode potential [31].
The key of the modified BDD-mechanism is that the proton, liberated in reaction 1,
can be reduced together with MOH+ads species at the more negative potential of -1.0V/SCE than
the half-wave potentials for proton from the solution between -0.46 and -0.93 V/SCE [27, 31]
giving rise to the formation of metal hydrides in reaction: H+ + e + M  M-H [28]. The total
concentration of protons is a sum of protons from the reaction and solution, i.e. [H+] Total = [H+]
Reaction + [H+] Solution.
Taking into the considerations both mechanisms-BDD or modified BDD-one can
conclude that the electrochemical reduction of M+2 ions is not favorable method to obtain a
precipitated one electron product, MOH, in deposit. Two major arguments led us to such
conclusion. First, the results in Fig. 8 demonstrate that the increase of Bs value above the Slater-
Pauling curve was not observed for CoFe alloys, thus FeOH and CoOH compounds were not
present in deposit. Second, the MOH+ads reduction (reaction 2) is the rate determining step
leading to MOHads, which is converted to M in the fast electrode reaction (3).
The increase of Bs value above Slater-Pauling curve was observed only in CoFeNi
alloys by reverse pulse electrodeposition (see results in Fig.8).Therefore, the NiOH compound is
probably formed in an one-electron oxidation step during anodic pulse in presumably an
analogous reaction 8 to the first step found for corrosion mechanism of iron [32-34]

Ni – e + HOH = NiOH + H+ (8)

(EP) a 8
The question: Why NiOH, and not CoOH and FeOH is present in deposit- remains unclear and
needs more work to be answered. In pulse reverse (PR) electrodeposition during the cathodic
pulse, ip, deposition of CoFeNi alloy occurs while during the reverse pulse, ir, anodic oxidation
or dealloying of CoFeNi alloy takes place. Dealloying refers to the selective dissolution of one or
more components of an alloy [44].By using thermodynamic arguments based on the standard
electrode potentials of 3 elements in CoFeNi alloy (EoNi = -0.257 V; EoCo = - 0.28 V; EoFe = -
0.44V) one can expect that Fe as the least noble element (LN) should be oxidized selectively
before more noble (MN) Co and Ni elements. The difference in potentials, Eo, between Fe and
Ni (183 mV) is enough for the selective oxidation/dealloying of Fe. However, it seems that
unexpected accumulation of NiOH as the most stable monovalent hydroxide in CoFeNi deposit
is likely to be due to complex kinetic effects such as: possibly higher oxidation potential of
NiOH in comparison to CoOH and FeOH, the rate of chemical reaction following the second
electron-transfer, the rate of the surface diffusion of adsorbed NiOH species, etc.

3.2. Effect of experimental conditions on composition, and magnetic properties of CoFe and
CoFeNi films.

Figure 3 shows the potential dependence during reverse pulse (RP) deposition on Au
electrode during 30 seconds obtained in quiescent CoFe solution under the following
experimental conditions: ip = 22 mA/cm2, ton = 1000 ms, ir = 0.0-5.0 mA/cm2, toff = 1000 ms.
During the cathodic pulse the potential stay around – 1.3 V/SCE (deposition region), while at ir =
0.0 mA/cm2 the potential during toff is stabilized around -0.8 V/SCE (H+ reduction region) and
becomes less negative by the increase of anodic pulse current density and stay at -0.4 V/SCE
(dissolution region-see Fig 3 and 4 from Ref. 25) at ir = 5.0 mA/cm2 . The response of EQCM
during electrodeposition/dissolution of CoFeNi film on the Au electrode indicates the
considerable increase of mass (~80 g) at ir = 0.0 mA/cm2, followed by decrease to ~40 g at ir
=0.0 mA/cm2 and to ~17 g at ir = 5.0 mA/cm2.
The effect of cathodic, ip (Fig. 4a) and anodic, ir (Fig. 4b) pulse current density on Bs
value of CoFeNi films (~550 nm) were studied under constant duty cycle ( =0.66) with ton =
1000 ms and toff = 500 ms in CoFeNi solution with 10/g/l of NiSO4 6H2O. Fig. 4c demonstrates

(EP) a 9
the effect of toff on the Bs value in PR deposition keeping the other parameters constant (ip = 22
mA/cm2, ton = 1000 ms, ir = 1.0 mA/cm2). The results shown in Fig. 3a-3c helped us to define the
following “optimal” working conditions in PR deposition: ip = 50 mA/cm2, ton = 1000 ms, ir = 1.0
mA/cm2, toff = 500 ms.

Figure 5 shows the effect of increase of CoSO4 concentration in CoFe solution with 20
g/l of FeSO4 on composition, Bs value, and coercivity of CoFe films. The thickness of films
increased gradually-by addition of NiSO4 into the solution-from 0.62 to 0.67 m. With a
decrease of Fe and increase of Co content in CoFe deposited films, the Bs values increased from
2.2T for Fe film to 2.38T for Co38Fe62 film (Fig.5a). Both, easy (Hc, e) and hard (Hc, h) axis
coercivities decreased by increase of NiSO4 concentration in solution (Fig.5b)
Figure 6 shows the effect of addition of NiSO4 concentration in CoFe solution on
composition, Bs value, and coercivity of CoFeNi films. The thickness of films increased
gradually by addition of NiSO4 into solution from 0.62 to 0.68 m. The content of Fe in CoFeNi
films remained around 64 %, while Co decreases and Ni increases. The most important effect of
addition of NiSO4 is steady increase of Bs value from 2.35T to 2.59 T (Fig. 6a). The easy and
hard axis coercivity decreased with the increase of Ni content in CoFeNi film (Fig. 6b).
Figure 7 shows typical hysteresis curves of 2.38T Co38Fe62 obtained from a CoFe
solution (Fig. 7a) in the presence of 15 g/l of CoSO4, 2.37T Co32Fe64Ni4, obtained from a
CoFeNi solution in the presence of 10 g/l of NiSO4 (Fig. 7b), and 2.52T Co20Fe64Ni4 obtained
from CoFeNi solution in the presence of 50 g/l NiSO4 (Fig.7c). The values of Hc, e, Hc, h, and Hk
from hysteresis curves and measured grain sizes are shown in Table1.

3.3. Coercivity
Coercivity of CoFe and CoFeNi films, obtained by electrodeposition [20-22, 29, 35-37],
shows general thickness-dependent property of ferromagnetic films. Coercivity, Hc, decreases
with the increase of film thickness, t, according to the relation derived by Neel [38], Hc = C t-n,
where C is the constant depending on morphology, microstructure, and substrate (seed layer).
The n-exponent determined experimentally was found to be 0.3-1.5 [37]. In order to compare Hc

(EP) a 10
values of compositionally different films, we have deposited the films of similar thickness
ranging from 0.62 m to 0.68 m.
Coercivity of ferromagnetic material is a quantitative measure for the magnetic field
required to reverse magnetization direction and overcome the effective anisotropy energy. The
total effective anisotropy energy in the thin film, Kt , includes a contribution of magnetoelastic,
Kme, magnetocrystalline, Kmc, and shape anisotropy energy, Ku (Eq. 9). For thin films with
positive tensile stress, , and magnetostriction, s, the Ku as a demagnetizing factor (-1/2 oMs2)
is taken to be zero for in-plane anisotropy [39]

Kt = Kme + Kmc + Ku (9)


The dominant anisotropy in CoFe and CoFeNi alloys seems to be magnetoelastic anisotropy
energy from coupling between stress, , and magnetostriction,s, given by Eq. 10.
Kme = (3/2) s  (10)
For example, the calculated magnetoelastic energy density for PR electrodeposited Co38Fe62
alloy (Fig. 7) is about 4.7 x 104 J/cm3 where s = 63 x 10-6 taken from literature [36 ] for CoFe
alloy with 61% Fe, and using measured average stress, = 500 MPa, of Co38Fe62 film. This is
about an order of magnitude higher than the magnetocrystalline energy anisotropy, Kmc, of bulk
CoFe alloy of the same composition [40]. However, easy axis coercivity is doubled, Hc, e, ~20
Oe, for the Co37Fe63 film obtained by DC electrodeposition [see Fig.11 from Ref. 36], compared
to the Co38Fe62 film obtained by PR electrodeposition in this work (Fig. 7a, Table 1). The larger
average grain size (~35 nm) correlates with higher Hc, e value in Co37Fe63 film [36]. That means
that smaller grain size has great impact on coecivities of CoFe and CoFeNi films (Fig. 5b and
Fig.6b). Notably, the measured grain size from the deconvoluted bcc (110) peak, by using the
Scherrer formula was very small, i.e. 12.8-18.7 nm (Table 1). Generally, in pulse
electrodeposition (PP or PR) the population of adatoms on the surface during pulse deposition is
higher than during constant current density (CD) electrodeposition, resulting in an increased
nucleation and therefore in a finer grained structure [41]. The decrease of coercivity observed for
CoFeNi films can be attributed to the decrease of magnetoelastic anisotropy energy, Kme, due to
the decrease of magnetostriction. It was shown experimentally [21,22, 42] that in CoFeNi alloys
by decreasing of Co and increasing of Ni in deposit-as we have also observed in this work
(Fig.6)-coercivity decreases.

(EP) a 11
The random anisotropy model [43] provides a guiding principle through magnetics of
materials with a grain size diameter, D, lower than the exchange length, Lex = (A/K1)1/2 , where
A denotes the exchange constant and K1 denotes the first term in the angular expansion of the
magnetocrystalline anisotropy energy, Kmc. If the grain size diameter is below the magnetic
exchange length (D < Lex), a domain wall can overlap a number of grains which leads to the
decrease of coercivity.

3.4. Magnetic saturation


Figure 8 shows the Slater-Pauling curve for CoFe alloys-Bohr magnetron per atom
(B/atom) vs. number of (3d + 4s) electrons-with the superposition of the experimental values
obtained for magnetic saturation, Bs, in this work with electrodeposited CoFe, CoFeNi and SP
Co35Fe65 alloys. The Slater-Pauling curve (S-P curve) was constructed for CoFe alloys with 0 to
80 %.Co. The saturation magnetization of CoFe alloy is a linear combination of the partial
magnetization for Co (in B/atom) and Fe (in B/atom) multiplied by fractions of their
compositions, XCo and YFe, respectively. Due to the fact that Fe is a weak ferromagnet rather
than a strong ferromagnet such as Co and Ni, for pure Fe the value of mFe = 2.2 B/atom was
used. The increase in saturation moment of Fe to 2.8 B/atom-through substitution of Co atom-is
attributed to the effect of rearrangement of the outer shell electron configuration [5]. Taking this
value for Fe and using Co = 1.8 B/atom for Co as a strong ferromagnet we have constructed S-P
curve and calculated saturation moment for Co35Fe65 alloy as mCoFe = 2.45 B/atom, which is
equivalent to the measured Bs value-of SP Co35Fe65 alloy-or oMs = 2.45T.
The results clearly demonstrate that Bs values for CoFe alloys did not cross above
S-P curve and did not break the 2.45T barrier, while CoFeNi alloys with 10-17.5% Ni (see Fig.
6a) showed an ultra high magnetic saturation with Bs values from 2.4T to 2.59T.The XPS
analysis (see Fig.10) and measured Bs values seems to verify our assumption that the increase of
magnetization is a result of the increase d-electron numbers due to the presence of NiOH in
thedeposit-presumably at the grain boundaries.

(EP) a 12
3.5. XPS analysis of oxygen in 2.52T Co20Fe66Ni14 film

The chemical composition of CoFe and CoFeNi films obtained in this work were
analyzed using ISP on total transition metals without analyzing any of light elements present as
“impurities” in the deposit. The chemical analysis of oxygen present presumably in the form of
NiOH in 2.52T Co20Fe66Ni14 ~620 nm thick film, was obtained from cross-sectional surface
which was sputtered for 12 min using 1 kV Ar+ beam and 5 min 500 eV beam to remove surface
oxides and hydroxides compounds before analysis. Notably, all experiments were performed in
industrial electroplating tool equipped with an additional apparatus that enabled continuous
reduction of Fe+3 ions, formed by oxidation of Fe+2 ions by oxygen dissolved in solution. It has
been demonstrated that the parallel electrochemical reduction of Fe+3 ionic species are the major
source of accumulation of oxide and hydroxide compounds in CoFe deposit [45,46]
Figure 9 shows low resolution XPS spectra for 2.52T Co20Fe66Ni14 ~620 nm thick film.
The “bulk” concentration of elements were analyzed from XPS spectra and calculated in atomic
percentage (see results attached in Table). The standard binding energy of different chemical
states of each element was analyzed according to literature data [47, 48]. Therefore, based on
XPS analysis the ternary alloy-characterized as 2.52T Co20Fe66Ni14 in this work-is actually
quaternary alloy containing oxygen with composition in atomic %:Co19.06Fe50.86Ni15.51O14.51.
Importantly, the atomic percentage ratio of Ni/O = 1.07, which indicates the presence of oxygen
in the form of NiOH or NiO compounds.
Figure 10 shows the convoluted high-resolution XPS spectra of the oxygen 1s core
level. The peak at 530.13 eV can be attributed to NiO species, while a high intensity peak at
532.13 eV could be assigned to NiOH species. A weak peak at 533.9 eV indicates the presence
of H2O in the deposit. We believe that the presence of NiO and H2O compounds in deposit might
be due to the electron-transfer reaction of two NiOH species in close proximity (low reaction
entropy) according to Eq. 11
2 NiOH [ET reaction]NiO + Ni + H2O (11)

(EP) a 13
3.6. Crystal structure of 2.52T Co20Fe66Ni14 film

Figure 11 shows XRD scan of 2.52T Co20Fe66Ni14 film with thickness of ~600 nm
obtained by PR electrodeposition on Cu seed layer. The XRD patterns of crystalline films reveal
the existence body centered cubic phase (bcc) with (110), (200), (211) and (222) peaks. The
average grain size was estimated from deconvoluted bcc (110) peak with preferred crystal
orientation, using the Scherrer formula. The estimated grain size of this film was 12.8 nm.
Figure 12 shows the crystal microstructure examined by transmission electron
microscopy (TEM). The low-resolution bright-field TEM obtained from a cross-section sample
of the film deposited on Cu seed layer (Fig. 12 a) showed randomly oriented small grains grown
on Cu substrate. There was no indication of layered structure, expected to be induced by reverse
pulse deposition. High resolution TEM sample (Fig. 12b) shows the presence of small bcc
metallic grains (dark spots) which were surrounded with the amorphous phase (bright spots). The
amorphous phase contained NiOH and NiO compounds located presumably at the grain
boundaries.

3.7. Thermal stability of 2.52T Co20Fe66Ni14 alloy

The ~600 nm thick 2.52T Co20Fe66Ni14 film deposited on Cu seeded, 8-inch AlTiC
wafer was pre-sputtered for 5 min using 1kV Ar+ beam to remove surface oxides followed by
sputtering of 10 nm Ru layer. The Ru layer served as a protective layer from air oxidation of the
sample film during annealing process. The wafer, protected by Ru, was kept at room temperature
for 3 weeks. During this time occasional magnetic measurements using a BH looper were carried
out. Notably, the magnetization of the film remained constant at room temperature during 3
weeks with Bs = 2.52 T. The annealing of the same wafer sample for 2 hours at higher
temperatures showed a decrease of Bs values from 2.52 T at 23oC to 2.27 T at 400oC (Fig. 13).
We characterize 2.52T Co20Fe66Ni14 alloy as metastable due to its thermal instability. The
possible reason for instability could be the enhanced rate of reaction described in Eq. 11 at the
higher temperature and longer time, which transforms NiOH species to NiO, Ni, and H2O.
In conclusion, 2.52 T Co20Fe66Ni14 alloy and other CoFeNi alloys with 10-17.5% Ni
cannot be applied in fabrication of recording PMR heads due to the various high temperature

(EP) a 14
annealing steps during wafer fabrication process post deposition of these materials as a writer
pole, or for recording HAMR heads due to the high working temperature of HAMR heads.
However, due to their stability at room temperature, we see an open possibility for these
materials with ultra-high magnetic saturation in biomedical applications- for CoFeNi nanowires
to improve heating efficiency for magnetic hyperthermia, [50] and for CoFeNi nanoparticles to
improve MRI contrast agents.[51]

4. Conclusions
CoNiFe films with Bs values between 2.4-2.59 T were obtained by reverse pulse
electrodeposition. The boost in magnetic saturation was enabled by the entrapment of spin rich,
MOH deposits in the films, during reverse oxidation step. XPS analysis of the films indicated
the presence of O, possibly, in the form of NiOH, further supporting the assumption of creating
and trapping MOH films during reverse anodic pulse. The films displayed poor thermal stability
beyond room temperature, possibly due to the decomposition of NiOH species into NiO, Ni, and
H2O at higher temperature. XRD analysis of the films indicated the presence of body centered
cubic (bcc) phase, with a small average grain size around 12.8 nm. The metastable nature of the
films, preclude it from being considered for hard drive, writer pole application. However, due to
their stability at room temperature, we see an open possibility for these materials with ultra-high
magnetic saturation in biomedical applications.

Acknowledgment

We are grateful for the early support by Seagate Technology. We acknowledge the continual
help of our colleagues from Seagate Technology: Steve Riemer, Jie Gong, Venkat Inturi, Paul
Jallen, Pat Ryan, and Mark Kief, during our work on electrodeposition of CoFe and CoFeNi
magnetic alloys.

(EP) a 15
References

1. T. Osaka, Electrochim. Acta, 45 (2000) 3311.

2. E. I. Cooper, C. Bonhote, J. Heidmann, Y. Hsu, P. Kern, J. W. Lam, R. Ramasubramanian,

N. Robertson, L. T. Romankiw, H. Xu, IBM J. Res.&Dev., 49 (2005) 103.

3. C. Larson, J. R. Smith, Trans. Inst. Metal Finish., 89 (2011) 333.

4. R. M. Bozorth, Ferromagnetism, D. Van Nostrand Co. Inc., New York ,(1951), p. 194.

5. G. Scheunert, O. Heinonen, R. Hardeman, A. Lapicki, M. Gubbins, R. M. Bowman, Appl.

Phys. Rev., 3 (2016) 492.

6. T. K. Kim, M. Takahashi, Appl. Phys. Lett., 20 (1972) 492.

7. Y. Sugita, K. Mitsonka, M. Komuro, H. Hoshiya, Y. Komuro, H. Hoshiya, Y. Kozono, M.

Hanzono, J. Appl. Phys., 70 (1991) 5977.

8. M. Takahashi, H. Shoji, H. Takahashi, T. Wakiyama, M. Kinoshita, W. Ohitta, IEEE Trans.

Magn., 29 1993) 3040.

9. A. Sakuma, J. Appl. Phys., 79 (1996) 8.

10. M. S. Patwari, R. H. Victora, Phys. Rev. B, 64 (2001) 214417.

11. N. Ji, L. F. Allard, E. L-P. Wang, Appl. Ohys. Lett., 98 (2011) 092506.

12. N. Ji, Y. Wu, J.-P. Wang, J. Appl. Phys., 109 (2011) 07B767.

13. B. D. Cullity, Introduction to Magnetic Materials, Addison-Wesley, Pulishing Co., London,

1972, p. 151.

14. I. M. L. Billas, A. Chatelaun, W. A. DeHeer, Science, 265 (1994) 1682.

(EP) a 16
15. K. W. Edmonds, C. Binns, S. H. Baker, M. G. Maher, J. Magn. Magn. Mater., 231 (2001)

113.

16. S. Binns, S. Louch, S. H. Baker, R. W. Edmonds, M. G. Maher, S. C. Thornton, IEEE Trans.

Magn., 38 (2002) 147.

17. R. H. Victora, L. M. Falicov, Phys. Rev., B 30 (1984) 3896.

18. J. Tersoff, L. M. Falicov, Phys. Rev., B 26 (1982) 6186.

19. A. Krakauer, A. J. Freeman, E. Wimmer, Phys. Rev., B 28 (1983) 610.

20. T. Osaka, M. Takai, K. Hayashi, K. Ohashi, M. Saito, K. Yamada, Nature, 392 (1998) 796.

21. X. Liu, P. Evans, G. Zangari, J. Magn. Magn. Mater., 226-230 (2001) 2073.

22. X. Liu, G. Zangari, M. Shamsuzzoha, J. Electrochem. Soc., 150 (2003) C159.

23. N. G. Chechenin, E. V. Khomenko, D. I. Vainchatein, J. Th. M. Hosson, J. Appl. Phys., 103

(2008) 07E738.

24. K. Kudo, G. Oikawa, Y. Maruyama, H. Shiina, US Patent, US 7,679,860 B2 (2010).

25. J. O’M. Bockris, D. Drazic, A. R. Despic, Electrochoim. Acta, 4 (1961) 235.

26. S. Hessami, C. Tobias, J. Electrochem. Soc., 136 (1989) 3611.

27. I. Tabakovic, S. Riemer, N. Jayaraju, V. Venkatasamy, J. Gong, J. Electrochim. Acta, 58

(2011) 25.

28. I. Tabakovic, J. Gong, S. Riemer, V. Venkatasamy, M. Kief, Electrochim. Acta, 25 (2010)

9035.

29. J. Gong, S. Riemer, A. Morrone, V. Venkatasamy, M. Kautzky, I. Tabakovic, J.

Electrochem. Soc., 159 (2012) D447.

30. I. Tabakovic, J. Gong, S. Riemer, M. Kautzky, J. Electrochem. Soc., 162 (2015) D102.

(EP) a 17
31. O. Dragos, H. Chiriac, N. Lupu, M. Grigoras, I. Tabakovic, J. Electrochem. Soc., 163 (2016)

D83.

32. J. O’M. Bockris, H. Kita, J. Electrochem. Soc., 108 (1961) 676.

33. J. O’M. Bockris, D. Drazic, Electrochim. Acta, 7 (1962) 293.

34. E. Kelly, J. Electrochem. Soc., 112 (1965) 124.

35. J. Gong, S. Riemer, M. Kautzky, I. Tabakovic, J. Electrochem. Soc., 156 (2016) D439.

36. S. Riemer, J. Gong, M. Sun, I. Tabakovic, J. Electrochem. Soc., 156 (2009) D439.

37. I. Tabakovic, V. Inturi, S. Riemer, J. Electrochem. Soc., 149 (2002) C18.

38. L. Neel, J. Phys. Radium, 17 (1956) 250.

39. D. C. Cronemryer, J. Appl. Phys., 70 (1991) 2911.

40. H. S. Jung, W. D. Doyle, J. E. Wittig, J. F. Al-Sharab, Appl. Phys. Lett., 13 (2002) 81.

41. J.-C. Puippe, F. Leaman, Eds., Theory and Practice of Pulse Plating, American

Electroplaters, Orlando, 1986, p.5.

42. B. Y. Yoo, S. C. Hernahdez, D.-Y. Park, N. V. Myung, Electrochim. Acta, 51 (2006) 6346.

43. G. Herzer, IEEE Trans. Magn., 25 (1989) 3327.

44. J. Erlebacher, J. Electrochem. Soc., 151 (2004) C614.

45. I. Tabakovic, S. Riemer, K. Tabakovic, M. Sun, M. Kief, J. Electrochem. Soc., 153 (2006)

C586.

46. S. R. Brankovic, S,-E. Bae, D. Litvinov, Electrochim. Acta, 53 (2008) 5934.

47. J. F. Mondler, W. E. Stickle, P. E. Sobel, K. D. Bomben, “Handbook of X-ray Photoeletron

Spestroscopy”, Perkin-Elmer, Eaden-Praire, MN, 1992.

48. A. Briggs, M. P. Seak, Eds., “Practical Surface Analysis” , Vol. 1,John Willey&Sons, New

York, 1990.

(EP) a 18
49. C. D. Wagner, D. A. Zatko, R. H. Raymond, Anal Chem., 52 (1980) 1445.

50. J. Alonso, H. Khushid, V. Sankar, Z. Nemati, M. H. Phan, E. Garayo, J. A. Garcia, H.

Srikanth, J. Appl. Phys., 117 (2015) 17D113.

51. B. H. Bin Na, I. C. Song, T. Hyeon, Adv. Mater., 21 (2009) 2133.

Table 1 and Figures Captions:

Fig.1. Current - time profile in the reverse pulse electrodeposition.

Fig.2. (a) Hysteresis curve of ~500 nm SP Co35Fe65 film, (b) Effect of external magnetic field on

measured Bs value of 0.37 m SP Co35Fe65 film.

Fig.3. Potential dependence during reverse pulse electrodeposition of CoFe films on Au

electrode during 30 seconds; ip = 22 mA/cm2 (ton = 1000 ms); ir = 0.0-5.0 mA/cm2 (toff = 1000

ms).

Fig.4. Effect of reverse pulse conditions on Bs value: (a) Cathodic pulse current density, ip,

(b)Anodic pulse current density, ir, and (c) Time-off, toff, under other constant parameters: ip =

22 mA/cm2, ton = 1000 ms, ir = 1.0 mA/cm2.

Fig.5. Effect of CoSO4 7H2O concentration in solution with 20 g/l FeSO4 7H2O on (a)

composition and Bs value, and (b) coercivity of CoFe films.

Fig.6. (a) Effect of NiSO4 6H2O concentration in solution containing 20 g/l FeSO4 7H2O and 15

g/l CoSO4 7H2O on composition and Bs value, and (b) coercivity of CoFeNi films.

Fig.7. Typical hysteresis curves of (a) 2.38T Co38Fe62, (b) 2.37 T Co32Fe64Ni4, and (c) 2.52 T

Co20Fe66Ni14

(EP) a 19
Fig.8. Superposition of the experimental values obtained for magnetic saturation, Bs, of ED

CoFe, ED CoFeNi, and SP CoFe films with the Slater-Pauling curve for CoFe alloys.

Fig.9. XPS spectra obtained from cross-sectional surface of ~600nm thick 2.52 T Co20Fe66Ni14

film.

Fig.10. High resolution of O 1s XPS spectra from ~600nm thick 2.52 T Co20Fe66Ni14 film.

Fig.11. X-ray diffraction patterns (2,  = 3o) of ~600nm thick 2.52 T Co20Fe66Ni14 film

obtained by reverse pulse electrodeposition.

Fig.12. Transmission electron microscopy spectra (TMS) obtained from ~600nm thick 2.52 T

Co20Fe66Ni14 film, (a) Low resolution BF TEM image of the film deposited on Cu substrate, (b)

High resolution TEM plain view image of the same film.

Fig.13. Effect of annealing temperature on Bs value of 2.52 T Co20Fe66Ni14 film.

(EP) a 20
Fig. 1

(EP) a 21
Fig.2

(EP) a 22
Fig. 3

(EP) a 23
Fig.4

(EP) a 24
Fig. 5

(EP) a 25
Fig.6

(EP) a 26
Fig.7

(EP) a 27
Table 1

Table 1. The values of Hc, Hk, and grain size of selected alloys from Fig. 7

Alloy Hc,e Hc,h Hk Grain size*


(Oe) (Oe) (Oe) (nm)
2.38T Co28Fe62 16.9 11.0 5.4 18,7

2.37T Co32Fe64Ni4 26.7 25.3 8.1 13.9

2.52 T Co20Fe66Ni14 14.4 6.4 10.4 12.8

*Measured by XRD

(EP) a 28
Fig. 8

(EP) a 29
Fig. 9

(EP) a 30
Fig. 10

(EP) a 31
Fig. 11

(EP) a 32
Fig. 12

(EP) a 33
Fig. 13

(EP) a 34
Highlights

CoNiFe films with Bs values between 2.4-2.59 T were obtained by reverse pulse
electrodeposition. The boost in magnetic saturation was enabled by the entrapment of spin rich,
MOH deposits in the films, during reverse oxidation step. XPS analysis of the films indicated
the presence of O, possibly, in the form of NiOH, further supporting the assumption of creating
and trapping MOH films during reverse anodic pulse. The films displayed poor thermal stability
beyond room temperature, possibly due to the decomposition of NiOH species into NiO, Ni, and
H2O at higher temperature. XRD analysis of the films indicated the presence of body centered
cubic (bcc) phase, with a small average grain size around 12.8 nm. The metastable nature of the
films, preclude it from being considered for hard drive, writer pole application. However, due to
their stability at room temperature, we see an open possibility for these materials with ultra-high
magnetic saturation in biomedical applications.

(EP) a 35

You might also like