j. am. ceram. soc., doi: 10.1111/jace.12150 journal 2013

7
Nanocomposite Derived from CoreShell Nanoparticles via Creep Deformation of an Amorphous Silica Layer Below its Glass Transition Temperature Yoshiaki Kinemuchi, Atsuya Towata, and Masaki Yasuoka National Institute of Advanced Industrial Science and Technology, 2266-98 Anagahora, Shimoshidami, Moriyama, Nagoya 463-8560, Japan The microstructural design of ceramics is relevant to tune their properties. Nanostructuring can drastically modify ceramic properties because of enhanced interfacial effects, although the creation of such structures in ceramics is still challenging because of the interfacial reaction and grain growth at elevated temperatures during sintering. Here, we demonstrate densifica- tion of coreshell nanoparticles consisting of Fe 3 O 4 (core parti- cle, 20 nm diameter) and SiO 2 (shell layer, 2 nm thick) with over 90% of theoretical density below 500°C, which was achieved by facilitating plastic flow of amorphous SiO 2 under high pressure below its glass transition temperature. Thus, grain growth of the core nanoparticles was strongly suppressed, and the core nanoparticles remained separated by an amor- phous layer in the final microstructure reflecting the original coreshell nanostructure. We also analyzed the densification behavior on the basis of a power law creep model, and esti- mated the pressures required to attain full density. I. Introduction S INTERING is a practical way to obtain a bulk form of oxide materials, although it normally requires heating at higher than 1000°C. Such high temperature inevitably leads to grain coarsening during the process, which makes nano- structuring of most ceramics very difficult. However, Chen and Wang pointed out that coarsening during the final sin- tering stage can be suppressed by adopting a heating profile, known as two-step sintering, in which grain-boundary diffu- sion dominates the densification progress. 1 Although this novel sintering profile is effective at maintaining the grain size below 100 nm, the grains still gradually grow to a size more than double than that of the starting nanoparticles. A sintering temperature of typically 0.5 T m (T m : melting point) in two-step sintering is regarded as the lowest bound for the densification of oxides via atomic diffusion, and thus external pressures are thought to be required to reduce the sintering temperature further. In fact, application of external pressure allows densities higher than 90% to be achieved with heating of 200°C for ZnO (0.2 T m ), 2 300°C for In 2 O 3 (0.3 T m ), 3 and 480°C for TiO 2 (0.35 T m ). 4 A remarkable feature of such low temperature processes is their lack of grain growth: the grain size is mostly identical to the size of the starting nanopar- ticles (typically 2040 nm) Furthermore, the grain size is stabilized below a certain threshold temperature, which is several hundred degrees higher than the temperature used for processing. 2 Interestingly, this threshold temperature, 0.4 T m for ZnO as an example, is identical to the onset sintering temperature of the nanoparticles without external pressure, 2 and therefore should be linked with grain-boundary diffusiv- ity. In other words, these results suggest that densification without grain growth is practically difficult unless densifica- tion is achieved below 0.4 T m . Pressure-assisted densification below 0.4 T m is mostly caused by either plastic yield 5 or power law creep. 6,7 Both these phenomena are related to the yield of the material, but the latter shows time dependence. In the above examples of densification of oxides by applying external pressure, 2,3 the densification occurred instantaneously upon pressing. There- fore, plastic deformation is the principal mechanism for the densification. As for the creep deformation of crystalline oxi- des, diffusional creep initiates plastic deformation which brings out superplasticity in some materials. 8 Because the atomic diffusion drives this behavior, it is significant only at high temperature (>0.7 T m ), but the lower bound for defor- mation temperature reduces to 0.4 T m in the case of nano- crystals. 4 This 0.4 T m again coincides with the onset temperature of grain growth in nanoceramics. As expected, grain growth occurred during creep tests, indicating that creep deformation does not freeze grain growth during densi- fication of nanocrystals. In contrast with the creep of nanocrystals, the creep behavior of amorphous phases is relatively well-known, 9,10 yet its feasibility for densification at low temperatures is uncertain. Nevertheless coating an amorphous layer onto nano particles may facilitate densification at low tempera- tures under external pressure. An example of densification of amorphous phases is the hot pressing of fused silica, which has been demonstrated above 1000°C 11 ; this temperature is, however, greater than 0.4 T m for most ceramic materials and thus clearly too high a temperature for the densification of nanocrystals without excessive grain growth. To overcome the limit of 0.4 T m in this study, we first dis- cuss the densification of 30-nm colloidal silica particles under external pressures below 500°C, and then demonstrate densi- fication of coreshell nanoparticles via the creep deformation of the silica shell layer. These coreshell microstructures allow us to both maintain the nanosize of the core particles and space the core particles out. They offer a practical way of obtaining bulk nanostructured materials, and may reduce the interaction between core crystals which occurs in nano- grained magnets, see Ref. [12] for example. II. Experimental Procedure Densification of colloidal silica (Aldrich, Ludox TM-40) was first tested. A transmission electron microscope, TEM (JEOL, JEM-2010, Japan), image of the powder showed that the particles were spherical, ~30 nm in size and had a largely monodisperse particle size distribution. (Fig. 1) The average particle size measured by dynamic light scattering (Malvern, G. Scherer—contributing editor Manuscript No. 31544. Received June 1, 2012; approved November 28, 2012. Author to whom correspondence should be addressed. e-mail: y.kinemuchi@aist. go.jp 429 J. Am. Ceram. Soc., 96 [2] 429–435 (2013) DOI: 10.1111/jace.12150 © 2013 The American Ceramic Society J ournal

Upload: others

Post on 25-Nov-2021

8 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: J. Am. Ceram. Soc., DOI: 10.1111/jace.12150 Journal 2013

Nanocomposite Derived from Core–Shell Nanoparticles via CreepDeformation of an Amorphous Silica Layer Below its

Glass Transition Temperature

Yoshiaki Kinemuchi,† Atsuya Towata, and Masaki Yasuoka

National Institute of Advanced Industrial Science and Technology, 2266-98 Anagahora, Shimoshidami, Moriyama,Nagoya 463-8560, Japan

The microstructural design of ceramics is relevant to tune their

properties. Nanostructuring can drastically modify ceramic

properties because of enhanced interfacial effects, although thecreation of such structures in ceramics is still challenging

because of the interfacial reaction and grain growth at elevated

temperatures during sintering. Here, we demonstrate densifica-tion of core–shell nanoparticles consisting of Fe3O4 (core parti-

cle, 20 nm diameter) and SiO2 (shell layer, 2 nm thick) with

over 90% of theoretical density below 500°C, which was

achieved by facilitating plastic flow of amorphous SiO2 underhigh pressure below its glass transition temperature. Thus,

grain growth of the core nanoparticles was strongly suppressed,

and the core nanoparticles remained separated by an amor-

phous layer in the final microstructure reflecting the originalcore–shell nanostructure. We also analyzed the densification

behavior on the basis of a power law creep model, and esti-

mated the pressures required to attain full density.

I. Introduction

S INTERING is a practical way to obtain a bulk form ofoxide materials, although it normally requires heating at

higher than 1000°C. Such high temperature inevitably leadsto grain coarsening during the process, which makes nano-structuring of most ceramics very difficult. However, Chenand Wang pointed out that coarsening during the final sin-tering stage can be suppressed by adopting a heating profile,known as two-step sintering, in which grain-boundary diffu-sion dominates the densification progress.1 Although thisnovel sintering profile is effective at maintaining the grainsize below 100 nm, the grains still gradually grow to a sizemore than double than that of the starting nanoparticles. Asintering temperature of typically 0.5 Tm (Tm: melting point)in two-step sintering is regarded as the lowest bound for thedensification of oxides via atomic diffusion, and thus externalpressures are thought to be required to reduce the sinteringtemperature further. In fact, application of external pressureallows densities higher than 90% to be achieved with heatingof 200°C for ZnO (0.2 Tm),

2 300°C for In2O3 (0.3 Tm),3 and

480°C for TiO2 (0.35 Tm).4 A remarkable feature of such low

temperature processes is their lack of grain growth: the grainsize is mostly identical to the size of the starting nanopar-ticles (typically 20–40 nm) Furthermore, the grain size isstabilized below a certain threshold temperature, which isseveral hundred degrees higher than the temperature used forprocessing.2 Interestingly, this threshold temperature, 0.4 Tm

for ZnO as an example, is identical to the onset sinteringtemperature of the nanoparticles without external pressure,2

and therefore should be linked with grain-boundary diffusiv-ity. In other words, these results suggest that densificationwithout grain growth is practically difficult unless densifica-tion is achieved below 0.4 Tm.

Pressure-assisted densification below 0.4 Tm is mostlycaused by either plastic yield5 or power law creep.6,7 Boththese phenomena are related to the yield of the material, butthe latter shows time dependence. In the above examples ofdensification of oxides by applying external pressure,2,3 thedensification occurred instantaneously upon pressing. There-fore, plastic deformation is the principal mechanism for thedensification. As for the creep deformation of crystalline oxi-des, diffusional creep initiates plastic deformation whichbrings out superplasticity in some materials.8 Because theatomic diffusion drives this behavior, it is significant only athigh temperature (>0.7 Tm), but the lower bound for defor-mation temperature reduces to 0.4 Tm in the case of nano-crystals.4 This 0.4 Tm again coincides with the onsettemperature of grain growth in nanoceramics. As expected,grain growth occurred during creep tests, indicating thatcreep deformation does not freeze grain growth during densi-fication of nanocrystals.

In contrast with the creep of nanocrystals, the creepbehavior of amorphous phases is relatively well-known,9,10

yet its feasibility for densification at low temperatures isuncertain. Nevertheless coating an amorphous layer ontonano particles may facilitate densification at low tempera-tures under external pressure. An example of densification ofamorphous phases is the hot pressing of fused silica, whichhas been demonstrated above 1000°C11; this temperature is,however, greater than 0.4 Tm for most ceramic materials andthus clearly too high a temperature for the densification ofnanocrystals without excessive grain growth.

To overcome the limit of 0.4 Tm in this study, we first dis-cuss the densification of 30-nm colloidal silica particles underexternal pressures below 500°C, and then demonstrate densi-fication of core–shell nanoparticles via the creep deformationof the silica shell layer. These core–shell microstructuresallow us to both maintain the nanosize of the core particlesand space the core particles out. They offer a practical wayof obtaining bulk nanostructured materials, and may reducethe interaction between core crystals which occurs in nano-grained magnets, see Ref. [12] for example.

II. Experimental Procedure

Densification of colloidal silica (Aldrich, Ludox TM-40) wasfirst tested. A transmission electron microscope, TEM(JEOL, JEM-2010, Japan), image of the powder showed thatthe particles were spherical, ~30 nm in size and had a largelymonodisperse particle size distribution. (Fig. 1) The averageparticle size measured by dynamic light scattering (Malvern,

G. Scherer—contributing editor

Manuscript No. 31544. Received June 1, 2012; approved November 28, 2012.†Author to whom correspondence should be addressed. e-mail: y.kinemuchi@aist.

go.jp

429

J. Am. Ceram. Soc., 96 [2] 429–435 (2013)

DOI: 10.1111/jace.12150

© 2013 The American Ceramic Society

Journal

Page 2: J. Am. Ceram. Soc., DOI: 10.1111/jace.12150 Journal 2013

Zetasizer Nano ZS, U.K.) was 33 nm, in good agreementwith the TEM observation. Nanoparticles were then obtainedvia freeze-drying of the colloid. The X-ray diffraction pattern(XRD, Rigaku, RINT2000, beam source CuKa, Japan) ofthese particles was characterized by a halo peak around 20°,indicating the amorphous phase. The particles were chargedin a mold made of tungsten carbide which was loaded into avacuum chamber equipped with a plunger (Sumitomo, SPS1050, Japan). Subsequently, the particles were densified byuniaxial pressing with the plunger below 500°C at less than10 Pa chamber pressure. The uniaxial pressing was onlyapplied during the constant temperature part of the heatingprogram. Meanwhile, changes in the dimensions of the sam-ple were monitored. To exclude elastic deformation of theexperimental system (i.e., the punches and spacers ratherthan the sample) from the data, pressing was discontinuedfor a short time while the sample dimensions were measured.These data were used for the estimation of density during theprocess. Archimedes method was adopted to check the finaldensity of samples after sintering. The density was calculatedrelative to a theoretical density of 2.2 g/cm3. The shrinkagerate of the nanoparticles without pressure was also analyzedusing a dilatometer (Bruker AXS, TD5200S) in air. Thermo-gravimetric analysis (Rigaku, Thermo plus EVO II) of nano-particles was also performed in air.

After testing the SiO2 creep, the feasibility of exploitingSiO2 creep for nanostructuring was studied in a SiO2–Fe3O4

nanocomposite. First, Fe3O4 core particles were prepared byhydrothermal synthesis from FeCl3�6H2O (Wako Pure Chem-ical Industries, Japan) and FeCl2�4H2O (Wako Pure Chemi-cal Industries).13 Here, FeCl3�6H2O (1.39 g) and FeCl2�4H2O(0.57 g) were dissolved in 5 mL of water mixed with 5 g ofethylene glycol to obtain a precursor. Adding aqueousNaOH (5.0 g in 8 mL water) to the precursor led to precipi-tation of the reactants. The reactant mixture was thenplaced in a 25 mL Teflon-sealed autoclave and held at 180°Cfor 4 h to synthesize Fe3O4 nanoparticles. After washingthe particles several times with ethanol and distilled water,the surface of the particles was modified by hydrothermalreaction at 120°C for 1 h in 20 mL water containing 1 g ofaminopropyltriethoxysilane (Shin-Etsu Chemical Co., Ltd,Japan).14 The powder was again washed with ethanol anddistilled water several times. Then tetraethylorthosilicate(Wako Pure Chemical Industries) was coated onto the parti-cles by the St€ober method15 at room temperature. The thick-ness of the coating was adjusted to 2 nm, as shown in Fig. 2.The atomic ratio of Si/Fe measured by energy dispersive X-ray spectroscopy was 25.5/74.5 which corresponded to either

a SiO2/Fe3O4 molar ratio of 50.7/49.3 or volume ratio of38.6/61.4. Assuming the formation of a homogeneous layeron the 30-nm core particles, the ratio of Si/Fe corresponds toa layer thickness of 2.6 nm, which is in reasonable agreementwith TEM observation. For this composition, the theoreticaldensity is estimated to be 4.0 g/cm3. Finally, the core–shellparticles were densified based on the previously evaluatedSiO2 deformation. The microstructure of the sintered materi-als was observed by TEM after a conventional sample prepa-ration procedure: slicing, grinding, dimpling, and ion-milling.Magnetization curves of both the nanocomposites and nan-oceramics prepared from the Fe3O4 used as the core particleswere measured using a Vibrating Sample (Riken Denshi,BVH-10H, Japan).

III. Results and Discussion

(1) Densification of SilicaThe sintering curve of the SiO2 nanoparticles is shown inFig. 3, along with the strain rate. The significant shrinkage at900°C is responsible for the sintering, while a gradual increasein the density is observed below this temperature. The peaksin the strain rate below 900°C are thought to be because ofthe small size of the nanoparticles, which means that diffusionvia the surfaces becomes predominant at relatively low

Fig. 1. TEM image of spherical SiO2 for creep densification test.

(a)

(b)

Fig. 2. Micrographs of core–shell nanoparticles used as anarchitectural unit for nanostructuring. Here, Fe3O4 cores withdiameter of 30 nm (a) are coated with an amorphous SiO2 shell withthickness of 2 nm (b).

430 Journal of the American Ceramic Society—Kinemuchi et al. Vol. 96, No. 2

Page 3: J. Am. Ceram. Soc., DOI: 10.1111/jace.12150 Journal 2013

temperatures. Similar peaks have been found for other nano-sized ceramic powders.2,16 The sintering behavior is identicalwith that of porous silica glass whose glass transition temper-ature, Tg, increased linearly with sintering temperature.17

According to this linear relation, the Tg of our sample is esti-mated to be about 700°C, which is 500°C lower than that ofpure quartz glass.18 The hydroxy groups in the colloidal silicaare one reason for the low Tg, as they are known to disruptthe network structure of the glass.19 Indeed, weight lossrelated to decomposition of hydroxy groups20 was observedin the thermogravimetric analysis (Fig. 4). Interestingly, thedecomposition is mostly governed by temperature, althoughloss of 0.2–0.3 wt% was found during 1 h of isothermal treat-ments at 300°C and 500°C. This means that materials proper-ties change mainly depending on temperature and onlyslightly with time under isothermal conditions. In addition,the viscosity at Tg is normally of the order of 1012 Pa s,implying poor fluidity below this temperature.17 In fact, den-sification below Tg was negligible as shown in Fig. 3.

Densification of the silica nanoparticle compact by externalpressure shows exponential behavior (Fig. 5). The thresholdpressure for densification was largely similar, regardless oftemperature, however, the slope above the threshold reducedwith decreasing temperature. In general, the deformationmode of glass follows Newtonian flow, in which a linear rela-tion of deformation rate with pressure is observed. The densi-fication of fused silica glass obeys Newtonian flow aboveTg,

11 which enables the compact to be densified at ~1000°C inthis system,21 as observed in Fig. 3. Sintering by this mecha-nism is referred to as viscous sintering,22 and theoretical con-sideration of this form of sintering can be found in severalmodels.23,24 However, the densification behavior below Tg isuncertain. The change in glass structure at high pressure,where the density of silica increases from 2.2 g/cm3 above afew GPa of pressure,25 may partly influence the densificationbehavior because such high pressures will exist at the pointcontacts between particles in the initial sintering stage.

Fig. 3. Densification behavior of spherical SiO2 without highpressure. Density (q) and strain rate (de/dt) are shown as a functionof temperature (T). Even below the glass transition temperature(~700°C), gradual densification can be observed as shown in theinset.

(a) (b)

(c) (d)

Fig. 4. Weight change in SiO2 powder from room temperature (RT) to 1000°C at a heating rate of 10 K/min (a), from RT to 300°C withprolonged heating (b), and from RT to 500°C with prolonged heating (c). The change in residual water during prolonged heating (d) is estimatedbased on the total weight loss and weight loss during temperature holding at the corresponding temperature.

February 2013 Nanocomposite from Core–Shell 431

Page 4: J. Am. Ceram. Soc., DOI: 10.1111/jace.12150 Journal 2013

The effect of time on densification is shown in Fig. 6. Inthe all conditions studied, an obvious increase in the densitywith the progress of time was observed, indicating creepdeformation contributed to the densification. The increase indensification rate with temperature suggests variation inviscosity with temperature.

(2) Analysis of Silica DensificationEven below Tg, densification of silica, which was stronglydependent on pressure and temperature, was observed. Fig-ure 7 shows the same data as shown in Fig. 5 but plottedagainst the pressure reduced by the yield stress. Here, theyield stress at different temperatures was estimated fromreported Vickers hardness data10 using Tabor’s indentationlaw26 which describes the relationship between the averagecontact pressure during indentation and the yield stress ofmaterials subjected to the test. The data converge into auniversal curve which increases with roughly the third powerof pressure. This curve corresponds to the initial deformationof silica, thus the contribution of deformation by yield isthought to be rather large. Based on the indentation law,Arzt formulated the densification behavior by yield asfollows5:

ry � 4p3aZq

Pe (1)

Here, ry is the yield stress, a is the contact area, Z is thecoordination number, q is the density, and Pe is the externalpressure. If the above criterion is satisfied, densification

proceeds by yield. It is noted that both a and Z vary withthe progress of densification. For the random packing ofspherical particles,27 Z gradually increases with densification,which can be numerically expressed as a function of q5 asdiscussed below.

a ¼ pðq� q0Þ (2)

Z ¼ Z0 þ 9:5ðq� q0Þ ½q\0:85�Z ¼ Z0 þ 2þ 9:5ðq� 0:85Þ þ 881ðq� 0:85Þ3 ½q� 0:85�

(3)

,Z0 = 7.3 and q0 = 0.64. Substituting these values of Z and ainto Eq. 1 allows us to predict the densification due to yield.

Our starting nanoparticles possessed spherical shape(Fig. 1), but the initial compaction density was 0.60 � 0.01which was slightly lower than that expected for randompacking (0.64),27 suggesting agglomeration of particles. Thisdifference in initial packing causes quantitative disagreement,especially during the initial sintering stage; however, thetrend does not change, regardless of compaction density.Therefore, we adopted the model without modification, asshown in Fig. 7. The model predicts a gradual increase indensity with pressure, indicating that the yield deformationdoes not account for the densification of SiO2. Although thesintering temperature is far below Tg, it is worth mentioningthe possibility of viscous sintering. Under external pressure,with negligible capillary force, densification is predicted toobey the following equation28:

_q ¼ 3Pe

4g1� qð Þ (4)

Here, _q is the rate of densification and g is viscosity. Integra-tion of Eq. 4 results in the following expression:

ln1� q01� q

¼ 3Pe

4gt (5)

Here, t is time. This model also predicts increasing densitywith pressure, although the curve of q against Pe for themodel follows an increasing convex upward function at agiven time; this does not agree with our experimental obser-vations, indicating that viscous sintering is unlikely to beoccurring here.

As the influence of pressure on density shows power lawbehavior, we later discuss the densification based on the

Fig. 5. Densification of spherical SiO2 with applied pressure (P) atvarious temperatures.

Fig. 6. Time (t)-dependent densification of spherical SiO2 at varioustemperatures. The lines indicate the results of model fitting.

Fig. 7. Power law dependence of density on reduced pressure (P/ry,ry: yield stress). Dotted line indicates yield deformation model(Eq. 1).

432 Journal of the American Ceramic Society—Kinemuchi et al. Vol. 96, No. 2

Page 5: J. Am. Ceram. Soc., DOI: 10.1111/jace.12150 Journal 2013

power law creep model. This is thought to be adequatebecause of the time dependence of density at constanttemperature (Fig. 6). The basic idea of power law creepdensification relies on the strain rate ( _e) being proportionalto the nth power of pressure:

_e ¼ _e0pn (6)

Here, _e0 is material constant and p is the local pressure atthe contact. In the case of densification, the local pressure isa function of microstructure. If spherical pores are distrib-uted in the material, _q can be expressed by the Wilkinsonmodel6 as follows:

_q ¼ 3

2_e0

qð1� qÞ1� ð1� qÞ1=nh in 3

2nPe

� �n

(7)

Because of the assumption of microstructure, Eq. 7 is appli-cable to the final sintering stage, when q � 0.9. Equation 7reduces to the expression for Newtonian flow densificationfor n = 1, i.e., becomes identical to Eq. 4, thus _e0 equals to1/(3g) . Another expression for power law creep densificationis an equation based on a random packing model (Arztmodel7):

_q ¼ 5:3ðq2q0Þ1=3 _e0ffiffiffia

p

rp

3

� �n

(8)

where:

p ¼ 4pPe

aZq;

Equation 8 is regarded as adequate expression to the initialand intermediate sintering stages: q < 0.9. As both modelsrely on different considerations of the microstructure, therates predicted by the models are not the same at the transi-tion density of 0.9 (the rate given by random packing modelturns out to be faster than that of spherical pore model).This requires a scaling adjustment of the rate curves to applyboth models simultaneously, as shown in Fig. 8. Because ofthis scaling, the physical meaning of _e0 in the models isreduced, while another material property, n, uniquely governsthe shape of the rate curve. From the curve, we can under-stand the sudden drop in densification rate because of theneck growth in the initial sintering stage, the gradual reduc-tion in the rate during the intermediate stage, and the cessa-

tion of sintering in the final stage. The lines in Fig. 6 showthe results of fitting the models to the data. Here, the densityevolution was obtained by the integration of the deformationrate. We found the exponent, n, remained in the rangebetween 2 and 3. This exponent should relate to themicrostructural evolution under pressure (as it also describesanalogous events such as dislocation climb in crystals). How-ever, few studies of the deformation below Tg have beenreported.25 Consequently, we cannot be sure of the physicalmeaning of the number itself. We must also keep in mindthat the change in viscoelastic properties under isothermalconditions, which is expected to occur from the thermogravi-metric analysis, is incorporated in n. Hence, at the moment,we accept it as an empirical number which can be used topredict the deformation.

(3) Application of Creep Deformation to NanostructuringNow, we are ready for practical application of the model tonanostructuring based on the fitted parameters. Figure 9shows the estimated density evolution under various pres-sures. Here, the initial density was evaluated from the datashown in Fig. 5. The computational results indicate that thefinal density is mostly determined by the initial density atlower temperature, and therefore the required pressure forthe same density is higher when the temperature is lower.For example, a pressure of 1.8 GPa is required to attain highdensity at 300°C, whereas two third of that pressure isenough at 500°C. We also have to mention that the estima-tion may include large error when the initial density fallsbelow 0.74 as rearrangement of particle may contribute tothe densification in the initial stage of sintering.

Following the prediction, core–shell nanoparticles consist-ing of Fe3O4 cores and SiO2 shells (Fig. 2) were densified at300°C or 500°C under pressures of 2 or 1.2 GPa,

Fig. 8. Densification rate (dq/dt) curves as a function of density (q)based on the Wilkinson (Eq. 7) and Arzt (Eq. 8) models. The ratescale is normalized to be the same for both models at q = 0.9.

Fig. 9. Estimates of density evolution of SiO2 based on Wilkinsonand Arzt models.

February 2013 Nanocomposite from Core–Shell 433

Page 6: J. Am. Ceram. Soc., DOI: 10.1111/jace.12150 Journal 2013

respectively, for 1 h. Pressure was only applied during theconstant temperature part of the heat treatment. The resul-tant densities (as a fraction of the theoretical maximum den-sity) were 0.954 at 300°C and 0.986 at 500°C, indicating thevalidity of the original prediction. No obvious cracks wereobserved by the naked eye. The X-ray diffraction patternindicates no grain growth of the core particles. (Fig. 10)According to the phase diagram of these phases,29 Fe2SiO4

may form at low oxygen partial pressure. Although we useda graphite heater, which meant that there would be low oxy-gen partial pressure during heating, no traces of the second-ary phase were found in the XRD patterns.

Observation by TEM also supported the above results.TEM images (Fig. 11) confirmed the homogeneous distribu-tion of amorphous phase between core particles. The high-resolution TEM image shows Fe3O4 grains with latticefringes and an amorphous layer in between them. The SiO2

layer is roughly 2–3 nm thick, which is similar to the thick-ness of the shell layer of the starting particles. In otherwords, about half the volume of the SiO2 layer moved tofill the pores. By comparing with images of the startingnanoparticles (Fig. 2), it is found that the shape of the

core particles is maintained, which strongly suggests thatthe creep deformation of the shell layer governs the entiredensification. The magnetization curve of the nanocompos-ite indicated a decrease in saturation magnetization from77.8 to 58.1 emu/g because of the net reduction in theamount of magnetic material compared with monolithicnano-Fe3O4 ceramics, whereas an increase in coercive forcefrom 118 to 141 Oe was also observed, indicating decou-pling of the interaction between the core crystals by theSiO2 layer.

IV. Conclusion

Nanostructuring of bulk oxides is of interest in the presentwork. Specifically, we focused on an architecture based oncore–shell nanoparticles; this architecture allows the modifi-cation in the interface phase while retaining the dimensionsof the core phase. However, prevention of core-grain growthand phase mixing was a challenge. Such issues are connectedwith the high-temperature process required for densification;thus, a significant reduction in the process temperature isessential. Here, we have reported the feasibility of low tem-perature densification via creep deformation of SiO2 glass.Although the process temperature was below the glass tran-sition temperature, obvious densification was measured whenan external pressure higher than 300 MPa was applied to30-nm-sized SiO2 particles. This densification was reasonablyexplained by a model based on power law creep deformation.Analysis based on the model revealed that external pressuresof 1.8 and 1.2 GPa are required to reach high density at300°C and 500°C, respectively. Following this computationalanalysis, we demonstrated densification of core–shell nano-particles (consisting of 30-nm diameter Fe3O4 cores sur-rounded homogeneously by a 2-nm SiO2 layer). The finaldensities reached corresponded well with those predicted bythe original analysis. The microstructure showed homoge-neously distributed SiO2 phase between Fe3O4 nanograins.Furthermore, no core-grain growth or phase mixing wasobserved. Thus, we believe our demonstration paves a newway to design and control the nanostructure of bulk oxidematerials.

Acknowledgments

YK is grateful to Dr. K. Takagi of AIST for his help with the setup of thehigh-pressure apparatus and also to Ms. R. Suzuki of AIST for her supportwith TEM observations. This study was supported by New Energy andIndustrial Technology Development Organization (NEDO) project of“Research and development of alternative new permanent magnetic materialsto Nd–Fe–B magnets”.

References

1I. W. Chen and X. H. Wang, “Sintering Dense Nanocrystalline CeramicsWithout Final-Stage Grain Growth,” Nature, 404 [9] 168–71 (2000).

2Y. Kinemuchi, H. Nakano, M. Mikami, K. Kobayashi, K. Watari, andY. Hotta, “Enhanced Boundary-Scattering of Electrons and Phonons in Nano-grained Zinc Oxide,” J. Appl. Phys., 108 053721 (2010).

3Y. Kinemuchi, E. Guilmeau, O. I. Lebedev, and A. Maignan, “Tuning ofDimensionless Figure of Merit via Boundary Scattering in In2O3-d,” J. Appl.Phys., 110, 124304 (2011).

4H. Hahn and R. S. Averback, “Low-Temperature Creep fo NanocrystallineTitanium[IV] Oxide,” J. Am. Ceram. Soc., 74 [11] 2918–21 (1991).

5E. Arzt, “The Influence of an Increasing Particle Coordination on theDensification of Spherical Powders,” Acta Metall., 30, 1883–90 (1982).

6D. S. Wilkinson and M. F. Ashby, “Pressure Sintering by Power lawCreep,” Acta Metall., 23, 1277–85 (1975).

7E. Arzt, M. F. Ashby, and K. E. Easterling, “Practical Applications ofhot-Isostatic Pressing Diagrams: Four Case Studies,” Metall. Trans. A, 14A,211–21 (1983).

8F. Wakai, S. Sakaguchi, and Y. Matsuno, “Superplasticity of Yttria-StabilizedTetragonal ZrO2 Polycrystals,” Adv. Ceram. Mater., 1, 259–63 (1986).

9A. S. Argon, “Delayed Elasticity in Inorganic Glasses,” J. Appl. Phys., 39[9] 4080–6 (1968).

10J. H. Westbrook, “Hardness-Temperature Characteristics of Some SimpleGlasses,” Phys. Chem. Glasses, 1 [1] 32–6 (1960).

11T. Vasilos, “Hot Pressing of Fused Silica,” J. Am. Ceram. Soc., 43 [10]517–9 (1960).

Fig. 10. X-ray diffraction pattern of nanocomposite densified at500°C. The pattern of the core–shell nanoparticle starting powder isalso shown as a reference. dXRD indicates the correspondingcrystallite size based on Scherrer’s equation.

Fig. 11. TEM image of nanocomposite. Inset indicates Fe3O4

grains with lattice fringes and an amorphous layer in between them.The amorphous layer exists with a thickness of ~2 nm.

434 Journal of the American Ceramic Society—Kinemuchi et al. Vol. 96, No. 2

Page 7: J. Am. Ceram. Soc., DOI: 10.1111/jace.12150 Journal 2013

12R. Fischer and H. Kronm€uller, “Static Computational Micromagnetism ofDemagnetization Process in Nanoscaled Permanent Magnets,” Phys. Rev. B,54 [10] 7284–94 (1996).

13S. Li, G. W. Qin, T. W. Pei, Y. Ren, Y. Zhang, C. Esling, and L. Zuo,“Capping Groups Induced Size and Shape Evolution of Magnetite ParticlesUnder Hydrothermal Condition and Their Magnetic Properties,” J. Am.Ceram. Soc., 92 [3] 631–5 (2009).

14Y. Kobayashi, H. Inose, T. Nakagawa, K. Gonda, M. Takeda, N. Ohuchi,and A. Kasuya, “Control of Shell Thickness in Silica-Coating of Au Nanoparticlesand Their X-ray Imaging Properties,” J. Colloid Interface Sci., 358, 329–33 (2011).

15W. St€ober and A. Fink, “Controlled Growth of Monodisperse SilicaSpheres in the Micron Size Range,” J. Colloid Interface Sci., 26, 62–9 (1968).

16Y. Kinemuchi and K. Watari, “Dilatometer Analysis of Sintering Behaviorof Nano-CeO2 Particles,” J. Eur. Ceram. Soc., 28, 2019–24 (2008).

17T. Hiroshi, Y. Tetsuo, and E. Kiyoshisa, “Sintering Temperature ofPorous Glass and Transition Temperature of High Silica Glass,” Yogyo-Kyokai-Shi, 94 [6] 30–6 (1986). (in Japanese)

18G. W. Morey, “The Properties of Glass”, pp. 275–8, Reinhold publishingcorporation, New York, 1954.

19G. W. Scherer, C. J. Brinker, and E. P. Roth, “Sol->Gel->Glass: III.Viscous Sintering,” J. Non-Cryst. Solids, 72, 369–89 (1985).

20G. W. Scherer, “Dilation of Porous Glass,” J. Am. Ceram. Soc., 69, 473–80 (1986).

21M. D. Sacks and T. Tseng, “Preparation of SiO2 Glass From ModelPowder Compacts: II, Sintering,” J. Am. Ceram. Soc., 67, 532–7 (1984).

22C. J. Brinker and G. W. Scherer, “Sol-gel Science, The Physics andChemistry of sol-gel Processing”, Academic press, San Diego, Ch. 11, (1989).

23J. K. Mackenzie and R. Shuttleworth, “A Phenomelogical Theory ofSintering,” Proc. Phys. Soc. London, Sect. B, 62, 833–52 (1949).

24G. W. Scherer, “Sintering of low-Density Glass: I, Theory,” J. Am.Ceram. Soc., 60, 236–9 (1977).

25J. D. Mackenzie, “High Pressure Effects on Oxide Glasses: I, Densificationin Rigid State,” J. Am. Ceram. Soc., 46 [10] 461–70 (1963).

26D. Tabor, “The Hardness of Metals,” Clarendon press, Oxford (1951).27G. Mason, “Radial Distribution Functions From Small Packing of

Spheres,” Nature, 217 [24] 733–5 (1968).28P. Murray, E. P. Rodgers, and A. E. Williams, “Practical and Theoretical

Aspects of hot Pressing of Refractory Oxides,” Trans. Brit. Ceram. Soc., 53 [8]474–510 (1954).

29L. S. Darken, “Melting Points of Iron Oxides on Silica; Phase Equilibriain the System Fe-Si-O as a Function of gas Composition and Temperature,”J. Am. Chem. Soc., 70, 2046–53 (1948). h

February 2013 Nanocomposite from Core–Shell 435