preparation, characterization and microwave absorption prop

8
Preparation, characterization and microwave absorption properties of electroless Ni–Co–P-coated SiC powder Yongjie Li, Rui Wang, Fengming Qi, Chunming Wang * Department of Chemistry, Lanzhou University, 730000 Lanzhou, China Received 14 November 2007; received in revised form 16 January 2008; accepted 16 January 2008 Available online 20 January 2008 Abstract Silicon carbide particles reinforced nickel–cobalt–phosphorus matrix composite coatings were prepared by two-step electroless plating process (pre-treatment of sensitizing and subsequent plating) for the application to lightweight microwave absorbers, which were characterized by scanning electron microscopy (SEM), X-ray diffraction analysis (XRD), vibrating sample magnetometer (VSM) and vector network analyzer, respectively. The results show that Ni–Co–P deposits are uniform and mixture crystalline of a-Co and Ni 3 P and exhibit low-specific saturation magnetization and low coercivity. Due to the conductive and ferromagnetic behavior of the Ni–Co thin films, high dielectric constant and magnetic loss can be obtained in the microwave frequencies. The maximum microwave loss of the composite powder less than 32 dB was found at the frequency of 6.30 GHz with a thickness of 2.5 mm when the initial atomic ratio of Ni–Co in the plating bath is 1.5. # 2008 Elsevier B.V. All rights reserved. PACS : 41.20.Gz; 61.10.Nz; 68.37.Hk; 78.40.q Keywords: Silicon carbide; Microstructure; Complex relative dielectric permittivity; Complex relative magnetic permeability; Reflection loss 1. Introduction In recent years, microwave absorptive materials have attracted considerable research interest in the materials science [1,2] because of its widespread applications for many electro- magnetic compatibility (EMC) purposes. A number of materials have been described, which are capable of absorbing electro- magnetic radiation. However, the conventional absorptive materials such as metal powders and ferrites are quite heavy, which restricts their usefulness in applications requiring light- weight mass [3]. Moreover, those materials have difficulties in increasing the permeability in GHz region because of Snoek limit for ferrites [4] or eddy current loss for magnetic metals [5]. As one of the ways to overcome these problems, the use of embedding particles (metallic, non-metallic or polymeric) in electroless deposited metals is a convenient method of preparing composite coatings, and the particles increase its mechanical and physical properties [6]. The presence of fine particles as the second phase improves the microhardness, high- temperature inertness, wear and corrosion resistance of the composite coatings [7–9]. High wear resistance, low cost and chemical inertness of ceramic oxides, carbides and nitrides has led to their widespread use as distributed phase [8,10]. Within these ceramics SiC is a kind of useful electronic material, and has high-strength ceramic material with excellent corrosion and erosion resistance. The merits of SiC for high-temperature electronics and short-wavelength optical applications are compared. The outstanding thermal and chemical stability of SiC should enable them to operate at high temperatures and in hostile environments, and also make it attractive for high-power operation [11]. However, their application still suffers from several difficulties. One problem is the low wettability between ceramics and liquid metals. In order to promote wetting between them, the ceramic surface can be modified by deposition of metal coatings using different techniques [12]. So investigations have been carried out on electrodeposited composite coatings comprising of alloy matrixes dispersed with nanoparticles, e.g. Ni–Fe–nano-Si 3 N 4 , Co–Ni–nano-Al 2 O 3 , Zn–Ni–nano-SiC, Ni–P–SiC and Co–P–SiC composite coat- ings are attractive and have been investigated before [13,14]. Ni–Co alloy coatings are of importance, as they possess www.elsevier.com/locate/apsusc Available online at www.sciencedirect.com Applied Surface Science 254 (2008) 4708–4715 * Corresponding author. Tel.: +86 9318911895; fax: +86 9318912582. E-mail address: [email protected] (C. Wang). 0169-4332/$ – see front matter # 2008 Elsevier B.V. All rights reserved. doi:10.1016/j.apsusc.2008.01.076

Upload: momenzi

Post on 02-Nov-2014

32 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Preparation, Characterization and Microwave Absorption Prop

www.elsevier.com/locate/apsusc

Available online at www.sciencedirect.com

4 (2008) 4708–4715

Applied Surface Science 25

Preparation, characterization and microwave absorption properties

of electroless Ni–Co–P-coated SiC powder

Yongjie Li, Rui Wang, Fengming Qi, Chunming Wang *

Department of Chemistry, Lanzhou University, 730000 Lanzhou, China

Received 14 November 2007; received in revised form 16 January 2008; accepted 16 January 2008

Available online 20 January 2008

Abstract

Silicon carbide particles reinforced nickel–cobalt–phosphorus matrix composite coatings were prepared by two-step electroless plating process

(pre-treatment of sensitizing and subsequent plating) for the application to lightweight microwave absorbers, which were characterized by

scanning electron microscopy (SEM), X-ray diffraction analysis (XRD), vibrating sample magnetometer (VSM) and vector network analyzer,

respectively. The results show that Ni–Co–P deposits are uniform and mixture crystalline of a-Co and Ni3P and exhibit low-specific saturation

magnetization and low coercivity. Due to the conductive and ferromagnetic behavior of the Ni–Co thin films, high dielectric constant and magnetic

loss can be obtained in the microwave frequencies. The maximum microwave loss of the composite powder less than �32 dB was found at the

frequency of 6.30 GHz with a thickness of 2.5 mm when the initial atomic ratio of Ni–Co in the plating bath is 1.5.

# 2008 Elsevier B.V. All rights reserved.

PACS : 41.20.Gz; 61.10.Nz; 68.37.Hk; 78.40.�q

Keywords: Silicon carbide; Microstructure; Complex relative dielectric permittivity; Complex relative magnetic permeability; Reflection loss

1. Introduction

In recent years, microwave absorptive materials have

attracted considerable research interest in the materials science

[1,2] because of its widespread applications for many electro-

magnetic compatibility (EMC) purposes. A number of materials

have been described, which are capable of absorbing electro-

magnetic radiation. However, the conventional absorptive

materials such as metal powders and ferrites are quite heavy,

which restricts their usefulness in applications requiring light-

weight mass [3]. Moreover, those materials have difficulties in

increasing the permeability in GHz region because of Snoek limit

for ferrites [4] or eddy current loss for magnetic metals [5].

As one of the ways to overcome these problems, the use of

embedding particles (metallic, non-metallic or polymeric) in

electroless deposited metals is a convenient method of

preparing composite coatings, and the particles increase its

mechanical and physical properties [6]. The presence of fine

particles as the second phase improves the microhardness, high-

* Corresponding author. Tel.: +86 9318911895; fax: +86 9318912582.

E-mail address: [email protected] (C. Wang).

0169-4332/$ – see front matter # 2008 Elsevier B.V. All rights reserved.

doi:10.1016/j.apsusc.2008.01.076

temperature inertness, wear and corrosion resistance of the

composite coatings [7–9]. High wear resistance, low cost and

chemical inertness of ceramic oxides, carbides and nitrides has

led to their widespread use as distributed phase [8,10]. Within

these ceramics SiC is a kind of useful electronic material, and

has high-strength ceramic material with excellent corrosion and

erosion resistance. The merits of SiC for high-temperature

electronics and short-wavelength optical applications are

compared. The outstanding thermal and chemical stability of

SiC should enable them to operate at high temperatures and in

hostile environments, and also make it attractive for high-power

operation [11].

However, their application still suffers from several

difficulties. One problem is the low wettability between

ceramics and liquid metals. In order to promote wetting

between them, the ceramic surface can be modified by

deposition of metal coatings using different techniques [12].

So investigations have been carried out on electrodeposited

composite coatings comprising of alloy matrixes dispersed with

nanoparticles, e.g. Ni–Fe–nano-Si3N4, Co–Ni–nano-Al2O3,

Zn–Ni–nano-SiC, Ni–P–SiC and Co–P–SiC composite coat-

ings are attractive and have been investigated before [13,14].

Ni–Co alloy coatings are of importance, as they possess

Page 2: Preparation, Characterization and Microwave Absorption Prop

Y. Li et al. / Applied Surface Science 254 (2008) 4708–4715 4709

high-temperature wear and corrosion resistance. Moreover, the

Ni–Co alloy deposition is an anomalous co-deposition and the

hardness of alloy increases as long as they possess fcc lattice

structure. Thus it was necessary to understand the influence of

the matrix on the properties of the composites. The composite-

coating is believed to combine the advantages of both Ni–Co

alloy and nano-particulate SiC. One of the available report

focuses on studying the variation in SiC content in a given Ni–

Co alloy obtained from Watt’s bath [14].

As mentioned above, a lot of research work has been carried

out on the effect of operating conditions on the mechanical

properties of composite coatings containing micron size SiC

particles. Reports on the magnetic and microwave-absorbing

properties of nano-composite coatings are scanty. In resonant

absorbers of quarter wavelength, zero-reflection can be

obtained by access to wave impedance mating at the surface

of the absorbing layer, which requires a proper combination of

magnetic permeability and dielectric permittivity at a given

thickness and frequency. Those material parameters in high

frequencies can be controlled by use of magnetic metals

(permittivity control by electrical property and permeability

control by magnetic property) in a single-layered microwave

absorber [15]. In the present work, the thin Ni–Co–P films on

SiC particles were electrolessly synthesized. It was focused on

understanding the influence of Co content on magnetic and

microwave-absorbing properties of composite coatings. The

surface morphology and structure of Ni–Co–P films on SiC

particles were also investigated.

2. Experimental

2.1. Pre-treatment of SiC particles

The materials employed were pure SiC (97% minimum,

Crystolon, Norton Silicon Carbide) which have a true density of

3.2 g/cm3. The electroless plating was performed on the

particle size powders namely 200 grit, and its particle size is

74 mm. The pre-treatment of SiC particles was followed by the

literature of Wang et al. [16].

2.2. Electroless deposition process

The pre-treated SiC particles were plated in an electroless

nickel–cobalt bath with composition listed in Table 1. The

proportion of elemental powders that added to the plating bath

was 10 g/L, and with continuous stirring, all of the particles

were exposed to the electroless nickel–cobalt solution. The

powders were plated at 90 8C and the pH value adjusted to 8.0

using NaOH. The plating time was 30 min, and the samples 1–4

Table 1

Composition of the electroless nickel–cobalt bath

Nickel sulfate, hexahydrate NiSO4�6H2O 0.05–0.1 mol/L

Cobalt sulfate, heptahydrate CoSO4�7H2O 0.05–0.1 mol/L

Sodium citrate Na3C6H5O7�2H2O 40 g/L

Ammonium sulfate (NH4)2SO4 40 g/L

Sodium hypophosphite, monohydrate NaH2PO2�H2O 20 g/L

is plated from the plating solution in which the initial atomic

ratio of Ni–Co is 1:2, 1:1, 1.5:1, and 2:1, respectively.

2.3. Measurements

The surface morphologies of the coatings and the pre-treated

SiC particles were examined using a scanning electron

microscope (SEM, JSM-5600LV/KEVEX Sigma), while

EDX analysis was performed to identify the metal-coating

layer and its components. Phase analysis of the coating was

studied from 308 to 808 using an X-ray diffraction analyzer

(XRD, Rigaku D/max-2400, Cu K-a, l = 0.1514 nm). The

static magnetic properties of sample were investigated by

vibrating sample magnetometer (VSM, Lakeshore 7304). The

sample containing 75 wt.% sample particles was made into

toroidal-shaped samples with an outer diameter of 7.0 mm and

inner diameter of 3.0 mm for microwave measurement. The

complex permeability and permittivity of composite were

measured using a vector network analyzer (Agilent E8363B) in

the frequency range of 0.1–18 GHz.

3. Results and discussion

3.1. SEM/EDX and XRD analysis

Characterization of the Ni–Co–P-coated particles is neces-

sary to determine the amount, chemical nature, morphology,

uniformity and distribution of the metal coating on the powders.

Typical surface morphologies of pre-treated SiC particles and

Ni–Co–P-coated SiC powders that were heat-treated at 400 8Cfor 1 h are shown in Fig. 1. SEM examination of the starting

powders reveals that SiC particles exhibit predominantly

angular shapes with sharp edges and show a clean, deposit-free

surface (Fig. 1a). The chemical components of these particles

as revealed by EDX analysis show that these particles are

mainly composed of SiC. The morphology of sample 1 show

uniformly distributed Ni–Co–P coatings with spicular grains on

activated SiC particle surface (Fig. 1b). Such powders are

consists of a core of ceramic material and a metallic coating. In

Fig. 1c the Ni–Co–P deposits grew larger as the Co content

decreased. In Fig. 1d and e the form of the electroless Ni–Co–P

deposits turned into polyhedrons. Fig. 1 clearly shows that the

electroless deposits with the form of spicular grains were

deposited dispersively on the surface of SiC particles in the

beginning. The size of spicular grains increased more and more

with the decreased Co content until they touched each other in

the direction parallel to the surface of the substrate, which was

attributed to the magnetic attractive force of the coated Ni–Co–

P films. Then the electroless deposits grew only in the direction

vertical to the SiC surface and their form turned into

polyhedrons.

The EDX analysis of samples 1–4 shows that these particles

are primarily comprised of SiC, Ni, Co and P elements

(Table 2). Contrary to expectations, the components of Ni–Co–

P alloy did not show an ideal atomic ratio of Ni–Co compared

with the initial composition of plating bath. The reasons can be

explained as follows. First of all, it can be understood from the

Page 3: Preparation, Characterization and Microwave Absorption Prop

Fig. 1. Surface morphologies of (a) pre-treated silicon carbide, (b) sample 1, (c) sample 2, (d) sample 3 and (e) sample 4.

Table 2

Ni, Co, and P content in the Ni–Co–P alloy films coated on SiC powders

Sample number Ni (at.%) Co (at.%) P (at.%)

1 21 51 28

2 75. 19 5.4

3 68 27 5.7

4 86 7.1 6.7

Y. Li et al. / Applied Surface Science 254 (2008) 4708–47154710

thermodynamics and the kinetics of catalytic reaction. The Pd-

activation procedure assured elimination of the passive

condition of the SiC surfaces, making them suitable for

electroless plating due to the presence of chemically deposited

Pd nuclei that initiate nickel–cobalt deposition, which is a key

step in the entire electroless-coating process. For electroless

Ni–Co–P plating, the oxidizing and reducing reactions and their

standard Gibbs free energy changes [17] can be represented as

below

H2PO2� þH2O ¼ H2PO3

� þ 2Hþ þ 2e�; DG� ¼ �96:5 kJ

(1)

2H2PO2� þ 4Hþ þ 2e� ¼ 2P þ 4H2O; DG� ¼ þ48:3 kJ

(2)

Ni2þ þ 2e� ¼ Ni; DG� ¼ þ54:0 kJ (3)

Co2þ þ 2e� ¼ Co; DG� ¼ þ48:3 kJ (4)

From the value of DG8 for each reaction, it is clear that Ni, Co

and P can all reduced by sodium hypophosphite and the trend of

reduction for Ni is larger than Co. Additionally, the catalytic

activity of Ni for the reaction of (1) is better than that of Co [18].

Second, this is attributed to the fact that the electroless bath has

become exhausted due to the limited amount of the Ni–Co

plating solution compared with the large surface to be plated. In

addition, the pH of the alkaline solution was altered with the

progress of the plating process: during deposition, byproducts

of the reaction, orthophosphite (HPO32�) and hydrogen ions

accumulate in the solution, affecting the chemistry of the

plating bath. Such events were characterized by a reduced

gas evolution from the bath and an excessive decoloration of

the electroless solution.

Page 4: Preparation, Characterization and Microwave Absorption Prop

Fig. 2. XRD patterns of (a) pre-treated silicon carbide and heat-treated (b)

sample 1, (c) sample 2, (d) sample 3 and (e) sample 4 at 400 8C for 1 h.

Fig. 3. Hysteresis loops of samples 1–4 measured at room temperature.

Table 3

Magnetic properties of electroless Ni–Co–P ternary alloy deposits

Sample number Mr (emu/g) Ms (emu/g) Hc (Oe)

1 1.7 6.8 622

2 0.9 5.2 200

3 0.8 2.9 503

4 0.7 1.7 211

Y. Li et al. / Applied Surface Science 254 (2008) 4708–4715 4711

In order to confirm the presence of Ni–Co–P ternary alloy

within the bulk of the coating, XRD analysis of the Ni–Co–P-

coated SiC particles is investigated. Fig. 2 shows five X-ray

traces related to the pre-treated SiC particles and heat-treated

samples 1–4 at 400 8C for 1 h. X-ray diffraction patterns of the

heat-treated powders (Fig. 2b–e) show sharp peaks centered at

2u of about 36–538, signifying that this particular deposits

contains a mixture of crystalline Ni3P and a-Co. The figure also

shows sharp peaks corresponding to silicon carbide, as well as a

noisy background, characteristic of an amorphous structure. X-

ray diffraction traces obtained from samples 1 to 4 showed

sharp, well-defined peaks corresponding to a fully crystallized

structure (Fig. 2b–e). The five peaks at 2u = 36.38, 41.88, 42.78,46.68, and 52.88 represent the well-defined peaks correspond-

ing to diffraction from (0 3 1), (2 3 1), (3 3 0), (1 4 1), and

(1 3 2) planes, respectively, of the Ni3P (JCPDS: 34-501). And

the two peaks at 2u = 44.48 and 51.78 represent peaks

corresponding to diffraction from (1 1 1) and (2 0 0) planes,

respectively, of the a-Co (JCPDS: 01-1254).

It is evident that the amorphous nature of the deposits still

remains when the temperature of heat treatment is below

300 8C according to Ref. [19]. After the heat treatment at

300 8C for 1 h, XRD pattern reveals the presence of fcc nickel,

bct nickel phosphide (Ni3P) and hexagonal Ni5P2 phases. It can

be seen that the Ni5P2 phase disappeared, but tetragonal Ni12P5

phase was formed after annealing at 400 8C for 1 h. When the

temperature of heat treatment exceeds 400 8C, the Ni12P5 phase

vanished, the Ni phase and the Ni3P phase increase strongly.

The higher intensity for annealed samples pattern indicates to

the higher degree of crystallinity in these samples. In this paper,

during the annealing process, the higher phosphorus regions

were supposed to be further increased by extraction of P

dissolved in nickel grains, which in turn gave rise to

precipitation of the hard inter-metallic Ni3P phase, the most-

prevalent phosphide in the deposits. These transformations

support the earlier observation obtained by Sankara et al. [19]

and Hur et al. [20].

3.2. Magnetic properties of Ni–Co–P deposits

The magnetic properties of electroless Ni–Co–P ternary

alloy deposits were studied in their as-deposited condition. The

hysteresis loop obtained for as-plated electroless Ni–Co–P

deposits (samples 1–4) is shown in Fig. 3. The magnetic

properties, viz., specific saturation magnetization (Ms),

remanence (Mr) and coercivity (Hc), derived from the hysteresis

loop, are given in Table 3. The shape of the hysteresis loops

seems to be very similar to that exhibited by partially or totally

amorphous materials and by an amorphous Ni–Co–P film

[21,22]. It is well known that magnetic characteristics of

amorphous materials are significantly smaller than those of

crystalline materials [23]. Being amorphous in nature, the

electroless Ni–Co–P deposits of the present study exhibit soft

magnetic characteristics. According to Tarozaite et al. [24] the

formation of certain crystallites shape, which determines high

coercivity, takes place only at definite phosphorus content. The

phosphorus content, necessary for the formation of grains

boundaries, differs for the films, deposited under different

conditions, because of the grains with different size and shape

formation. According to them, electroless Ni–Co–P deposits

exhibit a high coercivity when the phosphorus content of the

deposit lies between 4 and 6 wt.%. When the phosphorus

content is lower than 4 wt.%, films of coarse crystallites are

deposited, resulting in lower coercivity. If phosphorus content

exceeds 6 wt.%, films consisting of very fine grains are

deposited which also leads to a decrease in coercivity.

Additionally the phosphorus content in Ni–Co–P alloys is

Page 5: Preparation, Characterization and Microwave Absorption Prop

Fig. 4. Frequency dependence of the real (a) and imaginary (b) part of the

complex permittivity of pre-treated silicon carbide and samples 1–4.

Y. Li et al. / Applied Surface Science 254 (2008) 4708–47154712

17, 2.9, 3.1 and 3.7 wt.% corresponding to samples 1, 2, 3 and

4, respectively. Thus the coercivities of all samples are low and

they are found to decrease and then increase with cobalt content

of the deposits (Table 3). A comparison of the magnetic

characteristics of electroless Ni–Co–P deposits of the present

study reveals that the specific saturation magnetization and

remanence are found to increase firstly and then decrease and

increase finally with increased cobalt content in the deposits

(Table 3), which is due to higher magnetic moment of Co atom.

However, Rivero et al. [25] report a linear increase in specific

saturation magnetization with increase in cobalt content for

electrodeposited Ni–Co–P amorphous ribbons and Matsubara

and Yamada [26] also report an increase in specific saturation

magnetization of electroless Ni–Co–P deposits with increased

cobalt content. Therefore detailed magnetic properties need to

be studied further.

3.3. Microwave absorptive properties

Fig. 4 illustrates the real and imaginary parts of permittivity

of samples 1–4, and SiC particles–olefin composites. It can be

seen from Fig. 4 that electroless Ni–Co–P ternary alloy deposits

of various composition show a similar variety trend and it is

apparent that both e0 and e00 are sensitive to metallic behavior of

Ni and Co in the deposits. The powder–paraffin composites

exhibit that the values of real part e0 of complex permittivity

decrease with increasing frequency except a resonance peak

around 8 GHz. It is found that the values of both real and

imaginary parts of complex permittivity of samples are higher

along with increasing metallic content of Ni and Co in the

deposits, which is attributed to better metallic behavior of Ni

and Co atoms than P atom, but they are comparatively smaller

than those observed for the metal powder or fibre composites as

reported in the literatures [27–29]. The lower real part value of

complex permittivity is a great advantage to strike a balance

between permeability and permittivity, thus decreasing the

reflection coefficient of the absorber compared with other metal

magnetic materials for microwave-absorbing application. It is

interesting that the values of imaginary part of complex

permittivity increase slightly with increasing frequency and the

curves of the real part of permittivity exhibit the abrupt

decrease, while the imaginary part of permittivity show the

sharp peak at the corresponding frequency. This suggests a

resonance behavior, which is expected when the composite is

highly conductive and skin effect become significant. In our

studies, the resonant frequency of the samples is related to the

high conductivity of nickel and cobalt. Additionally, the

decreasing amplitude of e0 and the increasing amplitude of e00

are enhanced with the increasing metallic content of Ni and Co

in the Ni–Co–P films, and they all shift to lower frequency. In

general, the permittivity originates from orientation polariza-

tion, atomic polarization and electronic polarization. Normally,

the resonance that originated from vacancy or pores usually

dominates in the low-frequency regions, provided that there

exist space charges in the materials. High-frequency resonance

is attributed to atomic and electronic polarization [30]. So we

observed a resonance peak in the curves of both the real and

imaginary parts of complex permittivity can be interpreted as

different chemical compositions of Ni–Co–P deposits that is the

intrinsic characteristics of our prepared materials and atomic

polarization.

The real and imaginary parts of permeability of samples 1–4,

and SiC particles–olefin composites are shown in Fig. 5. One

feature of the data is that the values of real part m0 of complex

permeability for all samples all decrease with increasing

frequency as shown in Fig. 5a, this is due to both eddy current

loss and ferromagnetic resonance [31]. The values of imaginary

part m00 of complex permeability for all samples also show a

decrease with increasing frequency as shown in Fig. 5b and

exhibit two distinct resonance peaks around 1 and 8 GHz,

respectively. Usually, for ferrite magnetic materials, the

microwave magnetic loss of magnetic materials originates

mainly from hysteresis loss, domain wall resonance, natural

ferromagnetic resonance, and the eddy current effect. The

hysteresis loss was caused due to irreversible magnetization

and was negligible in a weak applied field. It is possible to

separate the permeability spectra into the spin rotational

component and the domain wall motion contribution, using

numerical fitting [32]. Resonance due to domain wall move-

ment normally occurs at low-frequency region (<2 GHz);

Page 6: Preparation, Characterization and Microwave Absorption Prop

Fig. 5. Frequency dependence of the real (a) and imaginary (b) part of the

complex permeability of pre-treated silicon carbide and samples 1–4.

Fig. 6. Frequency dependence of the reflection loss of sample 3 at various

sample thicknesses.

Y. Li et al. / Applied Surface Science 254 (2008) 4708–4715 4713

however, resonance due to spin rotational component occurs as

high-frequency region. So the first resonance peal (around

1 GHz) may be due to domain wall resonance. It has been

reported that the natural resonance frequency for Co (hcp)

particles was 6.5 GHz [33–35]; that is to say, the magnetic loss

that peaked around 8 GHz could not be explained by natural

ferromagnetic resonance. Thus, it could be concluded that the

magnetic loss around 8 GHz was mainly caused by the eddy

current effect.

It is reasonable that both the dielectric loss and the magnetic

loss can be influenced by the ‘‘core–shell’’ microstructure of

microwave absorbent. In general, the dielectric loss is attributed

to the lags of polarization between the core/shell interfaces as

the frequency is varied. In addition, the ‘‘core–shell’’

microstructure of microwave absorbent has something to do

with eddy current loss that is one of the contributors to magnetic

loss. In our study, the SiC particles may act as a magnetic

inactive layer. The magnetic inactive layer causes the

demagnetizing field and the cut-off of the magnetic connection

between the magnetic components. We believed that the core of

SiC particles increased electromagnetic energy dissipation.

However the contribution of ‘‘core–shell’’ microstructure of

electroless Ni–Co–P-coated SiC powders on the microwave

properties is complicated in this paper, and the exact

mechanism yet is not ascertained, so more experiments need

to be done in the future.

For composites with n magnetic particles embedded in a

non-magnetic matrix, we can obtain the relationship between

theoretical effective permeability of the composites and the

total volume fraction of magnetic particles in the matrix.

Basically, the effective permeability meff increases for

composites with high-volume fraction p. A linear relationship

is also observed between the effective permeability and volume

fraction when the volume fraction p in the composites is small.

The SiC particles are non-magnetic, however, the electroless

Ni–Co–P-coated SiC powders are magnetic, therefore the

weight gain of electroless Ni–Co–P-coated SiC powders in the

matrix increases will increase the effective permeability meff of

composites. Similarly, the SiC particles are semiconductor, and

the thin film of Ni–Co–P is conductor, so the volume fraction p

of electroless Ni–Co–P-coated SiC powders in the matrix

increases will also increase the permittivity of composites.

The normalized input impedance Zin of a single metal-

backed microwave-absorbing layer is given by [36]:

Z in ¼ffiffiffiffiffimr

er

rtanh

2p fdffiffiffiffiffiffiffiffimrerp

c(5)

where mr and er are the relative complex permeability and

permittivity, respectively, of the composite medium, c is the

velocity of electromagnetic waves in free space, f is the

frequency of microwaves, and d is the thickness of the absorber.

The reflection loss is related to Zin by [36]:

RL ¼ 20 log

���� Z in � 1

Z in þ 1

���� (6)

Thus, the surface reflectance of an absorber is a function of six

characteristic parameters, viz, m0r, m00r , e0r, e00r , f , and d. Fig. 6

shows the calculated reflection loss as a function of frequency

for sample 3 at different thicknesses. The calculations use the

Page 7: Preparation, Characterization and Microwave Absorption Prop

Fig. 7. Frequency dependence of the reflection loss of samples 1–4 at thickness

of 2.5 mm.

Y. Li et al. / Applied Surface Science 254 (2008) 4708–47154714

actual values er and mr as shown in Figs. 4 and 5. The reflection

loss is found to depend sensitively on the thickness of the

absorber and the maximum attenuation of the incident wave is

observed with a thickness of 2.5 mm for sample 3, and the

values of reflection loss of the two peaks are �32.6 and �27.3,

respectively. As shown in Fig. 6, when it is related to same

sample the values of reflection loss become minus and they all

shift to lower frequency with increasing sample thickness.

Fig. 7 shows the calculated reflection loss for samples 1–4

at thickness of 2.5 mm. It is found that when it is related to same

sample thickness the values of reflection loss become minus

and they all shift to lower frequency with decreasing weight of

Co in the Ni–Co–P films. It is suggested that the reflection loss

is sensitive to a matching thickness and the composition of

samples.

The difference in microwave absorption properties of the

plated powders is resulted from the different composition

contents of the plated Ni–Co–P layers. The Bohr magnetons

(MB) of Co atom and Ni atom are 1.7 and 0.6, respectively. So,

the magnetism of cobalt is better than the nickel. P atom is non-

magnetic. Although cobalt content in sample 1 is the largest, the

phosphorus content in it exceeds 8 wt.%, which means poor

magnetism of sample 1. For comparison, the Co content of

sample 3 is larger and the P content of it is lower, so sample 3

possesses the better magnetism than the other samples. For

magnetic microwave absorbers, magnetic loss is the main

microwave loss tunnel. The powder with better magnetism

results in larger microwave loss. The frequency of the strongest

loss peak of the powders is different, which may be resulted

from the different magnetism of the powders. In a word, the

prepared samples exhibit good absorption performance in the

2–18 GHz frequencies and appear to be a potential microwave-

absorbing material.

4. Conclusions

In summary, the Ni–Co–P alloys were successfully

deposited on SiC particles by two-step electroless plating

process for the application to lightweight microwave absorbers.

Characterization techniques show that Ni–Co–P deposits are

spicular grains and comprised of a mixture of crystalline Ni3P

and a-Co, and exhibit low-specific saturation magnetization

and low coercivity. Due to the conductive and ferromagnetic

behavior of the Ni–Co thin films, high dielectric constant and

magnetic loss can be obtained in the microwave frequencies. It

is concluded that the higher the Co content in the alloy films, the

higher absorption rates for electromagnetic radiation. Reflec-

tion loss less than �32 dB were found at 6.30 GHz with a

thickness of 2.5 mm. The proposed absorber is well advanced in

both mass and thickness in comparison with conventional

ferrite absorber.

Acknowledgement

This work was supported by the National Natural Science

Foundation of China (Grant No. 20577017).

References

[1] R.C. Che, L.-M. Peng, X.F. Duan, Q. Chen, X.L. Liang, Adv. Mater. 16

(2004) 401–405.

[2] A. Wadhawan, D. Garrett, J.M. Perez, Appl. Phys. Lett. 83 (2003) 2683–

2685.

[3] K. Hatakeyama, T. Inui, IEEE Trans. Magn. 20 (1984) 1261–1263.

[4] J. Smit, H.P.J. Wijn, Ferrites, Phillips Technical Library, Eindhoven, 1959,

p. 271.

[5] D. Rousselle, A. Berthault, O. Acher, J.P. Bouchaud, P.G. Zerah, J. Appl.

Phys. 74 (1993) 475–479.

[6] G.O. Mallory, J.R. Hajdu, Electroless Plating, AESF, USA, 1990.

[7] E.A. Pavlatou, M. Stroumbouli, P. Gyftou, N. Spyrellis, J. Appl. Electro-

chem. 36 (2006) 385–394.

[8] P. Gyftou, M. Stroumbouli, E.A. Pavlatou, P. Asimidis, N. Spyrellis,

Electrochim. Acta 50 (2005) 4544–4550.

[9] N.K. Shrestha, M. Masuko, T. Saji, Wear 254 (2003) 555–564.

[10] K.H. Hou, M.D. Ger, L.M. Wang, S.T. Ke, Wear 253 (2002) 994–1003.

[11] H. Morkoc, S. Strite, G.B. Gao, M.E. Lin, et al. J. Appl. Phys. 76 (1994)

1363–1398.

[12] R. Asthana, J. Mater. Sci. 33 (1998) 1959.

[13] R. Tarozaite, M. Kurtinaitiene, A. Dziuve, Z. Jusys, Surf. Coat. Technol.

115 (1999) 57–65.

[14] L. Shi, C. Shun, P. Gao, F. Zhou, W. Liu, Appl. Surf. Sci. 252 (2006) 3591–

3599.

[15] S.-S. Kim, S.-T. Kim, J.-M. Ahn, K.-H. Kim, J. Magn. Magn. Mater. 271

(2004) 39–45.

[16] R. Wang, et al., Preparation and characterization of nanodiamond

cores coated with a thin Ni–Zn–P alloy film, Mater. Charact. 59 (2008)

108–111.

[17] O.M. Glenn, B.H. Juan, Electroless Plating: Fundamentals Applications,

American Electroplaters and Surface Finishers Society, Orlando, 1990.

[18] H.H. Chiang, W. Shen, The Fundamentals and Practice of Electroless

Plating, National Defense Industry Publisher, Bejing, 2001.

[19] T.S.N. Sankara Narayanan, S. Selvakumar, A. Stephen, Surf. Coat.

Technol. 172 (2003) 298–307.

[20] K.H. Hur, J.H. Jeong, D.N. Lee, J. Mater. Sci. 26 (1991) 2037.

[21] M.R. Khan, E.L. Nicholsan, J. Magn. Magn. Mater. 54–57 (1986) 1654–

1656.

[22] N. Fenineche, A.M. Chaze, C. Coddet, Surf. Coat. Technol. 88 (1996)

264–268.

[23] K. Huller, M. Sydow, G. Dietz, J. Magn. Magn. Mater. 53 (1985) 269–274.

[24] R. Tarozaite, G. Stalnionis, A. Sudavicius, M. Kurtinaitiene, Surf. Coat.

Technol. 138 (2001) 61–70.

Page 8: Preparation, Characterization and Microwave Absorption Prop

Y. Li et al. / Applied Surface Science 254 (2008) 4708–4715 4715

[25] G. Rivero, M. Multigner, J.M. Garcia, P. Crespo, A. Hernando, J. Magn.

Magn. Mater. 177–181 (1998) 119–120.

[26] H. Matsubara, A. Yamada, J. Electrochem. Soc. 141 (1994) 2386.

[27] M.Z. Wu, H.H. He, Z.S. Zhao, X. Yao, J. Phys. D: Appl. Phys. 33 (2002)

2927–2930.

[28] M.Z. Wu, H.H. He, Z.S. Zhao, J. Phys. D: Appl. Phys. 33 (2002) 2398–

2401.

[29] P. Chen, R.X. Wu, T.E. Zhao, F. Yang, J.Q. Xiao, J. Phys. D: Appl. Phys.

38 (2005) 2302–2305.

[30] A. Verma, A.K. Saxena, D.C. Dube, J. Magn. Magn. Mater. 263 (2003)

228–234.

[31] D. Wan, X. Ma, Physics of Magnetism, Press of University of Electric

Science and Technology of China, Chengdu, 1994, pp. 406–459.

[32] T. Tsutaoka, J. Appl. Phys. 93 (2003) 2789–2796.

[33] L. Olmedo, G. Chateau, C. Deleuze, J.L. Forveille, J. Appl. Phys. 73

(1993) 6992–6994.

[34] G. Viau, F. Ravel, O. Acher, F. Fievet-Vincent, F. Fievet, J. Appl. Phys. 76

(1994) 6570–6572.

[35] G. Viau, F. Ravel, O. Acher, F. Fievet-Vincent, F. Fievet, J. Magn. Magn.

Mater. 140–144 (1995) 377–378.

[36] S.S. Kim, S.B. Jo, K.K. Choi, J.M. Kim, K.S. Churn, IEEE Trans. Magn.

27 (1991) 5467.