You are on page 1of 13

Carbon 114 (2017) 334e346

Contents lists available at ScienceDirect

Carbon
journal homepage: www.elsevier.com/locate/carbon

Reinforcement and shape stabilization of phase-change material via


graphene oxide aerogel
Yue Xu, Amy S. Fleischer, Gang Feng*
Department of Mechanical Engineering, Villanova University, Villanova, PA, 19085, USA

a r t i c l e i n f o a b s t r a c t

Article history: Phase change materials (PCMs) are of interest in many applications which may require shape-
Received 10 August 2016 stabilization. In this study, a graphene oxide aerogel (GOxA) reinforced paraffin PCM composite is
Received in revised form developed, effectively reinforcing and shape-stabilizing the paraffin.
22 October 2016
The molecular and diffraction characterizations suggest that the GOxA network potentially affects
Accepted 25 November 2016
paraffin's crystallization. The mechanical characterizations using durometer and nanoindentation show
Available online 28 November 2016
that the composite is 3~7 harder than pure paraffin and maintains significant strength even above
paraffin's melting temperature. Moreover, the composite is much less strain-rate sensitive than paraffin.
The reinforcement via GOxA is much beyond the prediction by the rule-of-mixture, implying a strong
GOxA-paraffin interfacial bonding.
To our best knowledge, this is the first study on the mechanical behavior of paraffin and GOxA-PCM
composite, providing critical insights into their behavior. Additionally, the relationship between the
hardness and durometer index first-ever developed here will enable the quantitative durometer testing
on materials for many other applications at different ambient conditions due to its versatility and
simplicity.
© 2016 Elsevier Ltd. All rights reserved.

1. Introduction [9,16,20e29]. This leads to many potential applications including


energy storage [1,9,16,30,31], super-capacitors [20,23], air and
Phase change materials (PCMs) store and/or release thermal water purification [32,33], and nanocomposites [34].
energy through the solid-liquid phase transition and exhibit the Recently, much research has focused on the integration of
advantages of low cost and high energy density, making them a graphene-based nanomaterials with high thermal conductivities
popular choice for energy storage materials [1e6]. Paraffin is a into PCMs to increase their overall thermal conductivity [35e37]. It
high-performance PCM owing to its high latent heat, low super is also known that graphene aerogel (GA) impregnated with PCM
cooling, and chemical stability; however, its intrinsic low thermal exhibited a rise in thermal conductivity [38].
conductivity can make it difficult to transfer energy effectively into While the use of nano-inclusions can mitigate the effects of low
larger masses, and the possibility of liquid phase leakage can lower thermal conductivity, it is also necessary to address the issue of
the perceived reliability of the systems [1,2,7e9]. liquid leakage. Some methods are available to mitigate this po-
Aerogels are porous materials with extremely low bulk density, tential leakage, such as encapsulation and shape stabilization
composed of nanoparticles, nanowires, nanotubes, or nanosheets [2,39,40]. Shape stabilization of PCMs occurs through the addition
of oxides [10e12], silicon [12], metals [12e15], and carbon mate- of supporting materials into the PCM which can hold the liquid
rials [16e19]. Graphene-based three-dimensional (3D) aerogels phase in place with the matrix, usually by capillary action. Possible
have drawn attention due to their unique properties, including low supporting materials include high density polyethylene (HDPE) [2],
density, high porosity, large specific surface area, excellent elec- opal [8], graphite [41], carbon nanospheres [40], graphene oxide
trical conductivity and outstanding mechanical properties (GOx) [3,4,42], graphene aerogel [30,38], exfoliated graphite
nanoplatelets (xGnP) [43], graphite nanoplatelets [5,44], and gra-
phene nanoplatelets [45]. In some cases, large amounts of sup-
porting materials are needed in order to shape stabilize PCMs. For
* Corresponding author. Mechanical Engineering, Villanova University, 800
Lancaster Ave., Villanova, PA 19085, USA. example, a GO-paraffin composite showed no liquid leakage at
E-mail address: gang.feng@villanova.edu (G. Feng). 51.7 wt% of GO, but the latent heat dropped by 52% [4]. In the case of

http://dx.doi.org/10.1016/j.carbon.2016.11.069
0008-6223/© 2016 Elsevier Ltd. All rights reserved.
Y. Xu et al. / Carbon 114 (2017) 334e346 335

HDPE PCM [2], the shape-stabilization required at least 25e30 wt% nitrate were mixed with 120 ml sulfuric acid in an ice bath under
of HDPE. In this paper instead, the use of graphene oxide aerogel vigorous stirring for 2 h. 15 g of potassium permanganate was then
(GOxA) is explored as both a shape stabilizer and a thermal con- slowly added into the system with constant stirring while the
ductivity enhancer, because its use has been shown to not decrease temperature was kept below 20  C. After one hour of stirring, the
the latent heat [38]. temperature of the mixture was raised to 35  C and then the
Most studies of PCMs have focused on their thermal character- mixture was stirred overnight for complete oxidation yielding a
istics, but their mechanical properties are also vitally important thick paste. 150 ml of deionized water (DI water) was added to this
particularly as related to the shape stabilization behavior. The paste gradually and the mixture was heated to 98  C and stirred for
mechanical properties of graphene-based and GO-based polymer 12 h. 500 ml of DI water and 20 ml of hydrogen peroxide were then
composites have been studied, showing improvements in their added to the mixture. The mixture was then washed with 1 M HCl
strength and modulus [8,46e50] compared to the pure polymer solution to remove metal ions. The resultant graphite oxide was
counterparts. However, to our knowledge, the mechanical proper- washed with DI water and centrifuged for 20 min at room tem-
ties and shape stabilization behavior of GOxA-PCM composite have perature repeatedly until the pH reached 5e6. Finally it was fully
never been fundamentally studied. A comprehensive fundamental dried by heating at 65  C for two days.
understanding of the mechanical behavior of GOxA-PCM composite
is thus vitally important to understand the behavior of GOxA alone, 2.2. Synthesis of graphene oxide aerogel (GOxA)
pure paraffin, and the composite.
Typically, mechanical properties of porous 3D carbon aerogels/ By following the method of Hu et al. [20], the dried graphite
networks have been studied through either compression or nano- oxide powder obtained in last section was first re-dispersed into DI
indentation tests. Table A1 in Appendix 1 summarizes some of the water using a Vibra-cell VC505 ultrasonic processor in order to
recent results, indicating a large discrepancy in the literature form a 3 mg/mL graphene oxide (GOx) suspension. It is critical to
values. For example, for pristine graphene aerogel (GA) with den- have the right concentration of GOx in the suspension in order to
sity less than ~30 mg/cm3, data on strength is in the wide range of fabricate GOxA with the lowest possible density [20]. Then, 400 mL
0.2e40 kPa, while the modulus data shows a range of 2e1200 kPa. of EDA was added into 100 mL of the GOx suspension. This mixture
In general, by increasing the bonding between carbon nano- was sealed in a glass bottle and heated for 12 h at 90  C in order to
materials, e.g., by noble metal [51] or crosslinking [52], the overall synthesize the graphene oxide hydrogel (GOxH). The obtained
strength and modulus should increase. GOxH was immersed in DI water for purification for 3 days. The
In this work, by infiltrating GOxA with paraffin, a PCM com- purified GOxH was then immersed in acetone for 24 h in order to
posite with ~5 wt% of GOxA is formed. The mechanical behavior fully exchange the water in the GOxH with acetone. The purifica-
with respect to both shape stabilization and mechanical properties tion and water exchange process was performed 6 times for com-
will be comprehensively characterized at different strain rates as plete solvent exchange. Finally, the acetone-gel was dried using
well as at a range of temperatures, including above the melting supercritical drying (Leica EM CPD300) to obtain GOxA.
point of the PCM. While the thermal analysis of this GOxA-PCM
composite is also of interest, it will be left for a future paper in 2.3. Preparation of GOxA-PCM composite
order to concentrate here on the mechanical characteristics.
To our best knowledge, this is the first study on the fundamental Previous studies have shown great success in forming polymer-
strain-rate-dependent and temperature-dependent mechanical nanoparticle composites by infiltrating polymer into a porous
behaviors of paraffin and GOxA-PCM composite, providing critical nanoparticle structure [54,55]. GOxA and paraffin were heated to
insights to understand the behavior of graphene-related PCM 80  C (above the melting point of the paraffin) in a vacuum oven.
composites and their application into the design of high- The GOxA samples were immersed in the melted paraffin which
performance energy storage PCM composites with high reliability. fully infiltrated into the GOxA samples under vacuum. The com-
posites were then formed after cooling. The GOxA and composite
2. Material and methods samples were weighed using a scale (Citizen Scale CY 204 with
0.0001 g resolution) to calculate the weight percent of paraffin. The
The graphite powder was purchased from Spectrum Chemical GOxA's weight percentage in the composite samples is (5.2±0.51)
Manufacturing Corp. The sodium nitrate (NaNO3, high purity grade) wt%.
and potassium permanganate (KMnO4, ACS grade) were obtained
from Amresco, Inc. The hydrogen peroxide (H2O2, 29e32% w/w ar. 2.4. Structural and mechanical characterization
soln.) and Ethylenediamine (EDA, 99%) were purchased from Alfa
Aesar. Paraffin wax (IGI 1230A) was purchased from International The bulk density was calculated based on the weight and the
Group, Inc. All were used in as-purchased conditions. apparent sample volume. Scanning electron microscope (SEM)
In general, graphene-based aerogels are prepared by assembling images were produced using a Hitachi S-4800 field emission SEM.
2D graphene sheets into a 3D porous graphene architectures Fourier transform infrared (FTIR) spectra were recorded on a
through a sol-gel process and subsequently freeze-drying or su- Nicolet™ iS™10 FT-IR spectrometer. X-Ray diffraction (XRD)
percritical drying to remove the solvents in the wet gels while spectra were recorded on a Rigaku Miniflex X-ray Diffractometer
maintaining the shape of the 3D network [20,28]. Here, by with Cu Ka radiation. The Brunauer-Emmett-Teller (BET) surface
following the method of Hu et al. [20], GOxAs are formed using the area of GOxA was characterized via a MicroMeritics Gemini VII
ethylenediamine (EDA) - mediated reduction of GOxs followed by 2390 Surface Area Analyzer.
supercritical CO2 drying. The GOxA-reinforced paraffin composite The macroscale mechanical properties at different temperatures
is synthesized through infiltrating paraffin into GOxA. were characterized using a Phase II PHT-960 durometer. The nano/
microscale mechanical properties at different strain rates at room
2.1. Synthesis of graphite oxide

The graphite oxide was prepared using a modified Hummers 1


Throughout this paper, the Dx in the form of “x±Dx” always refers to the
method [53]. First, 5 g of graphite powder and 2.5 g of sodium standard deviation of quantity x.
336 Y. Xu et al. / Carbon 114 (2017) 334e346

temperature were studied using nanoindentation. The detailed microstructure and crystal phases, FTIR spectra and XRD patterns of
procedures and conditions for durometer and nanoindentation pure paraffin and GOxA-PCM composite were measured as in Fig. 4.
testing will be discussed in the related sections later. Fig. 4a shows the FTIR spectra of the GOxA-PCM composite
and paraffin. The GOxA-PCM FTIR spectrum is almost the super-
3. Results and discussion imposition of the spectra of GOxA alone and pure paraffin, with
the addition of three new peaks as shown. The spectra of both
3.1. Morphology and crystal structure characterization of graphene paraffin and GOxA-PCM show the 2915 cm1 and 2848 cm1 CeH
oxide aerogel stretching peaks [63], the 1462 cm1 and 1377 cm1 CeH
bending peaks [64,65], as well as the 719 cm1 CeH rocking peak
Images of the GOx suspension, GOxH and GOxA are shown in that can be seen only in long chain alkanes. The new peaks in the
Fig. 1. The GOxH showed some volume shrinkage after self- 1800 - 2500 cm1 range appear only for the GOxA-PCM com-
assembly (Fig. 1aeb), although this shrinkage is minor compared posite, implying new chemical bonding(s) in the GOxA-PCM
to other chemical reduction methods [20]. Fig. 1bec show that the composite, most likely at the GOxA-PCM interface. For example,
GOxH shrank roughly by 1/3 after the supercritical CO2 drying and the new peak at 2156 cm1 may correspond to C^C stretching
kept the 3D structure nicely. [64,65]. Further analysis on the new peaks and possible associ-
The GOxAs' density r ranges from 17 to 36 mg/cm3 based on 5 ated new phase(s) will be left as future work.
samples. The GOxAs' specific surface area (AS) is determined as high Fig. 4b shows the XRD patterns of the GOxA-PCM composite and
as 476 m2/g. If we assume spherical pores and neglect the gra- paraffin. Except for the new peak labeled, the GOxA-PCM com-
phene's thickness, 3/(rAS) [56] is an estimate for the average pore posite's XRD is almost identical to that of pure paraffin, suggesting
size r, i.e., 0.17e0.37 mm, which is roughly consistent with the SEM that embedding GOxA into paraffin does not significantly change
results (Fig. 2a). the crystallization of paraffin. Here, the XRD pattern of paraffin
Fig. 2a shows a SEM image of GOxA, implying interconnected corresponds to the orthorhombic structure [66]. Due to GOxA's
0.2e1 mm pores formed by thin layers of GOx sheets. Fig. 2bec extremely low density, GOxA alone will not generate significantly
show the SEM images of the as-synthesized and polished surfaces sharp peaks; in fact, previous studies indicate a weak and broad
of GOxA-PCM composite, showing complete infiltration without peak at around 25 for GOxA [9,26]. Fig. 4b also indicates one new
observable voids. peak at ~37.05 for the composite, corresponding to a plane spacing
The oxidation of graphite into GOx and the reduction of GOx of 2.43 Å, implying that a potentially new phase may exist in the
into GOxA were investigated using FTIR, as shown in Fig. 3. No composite, most likely at the GOxA-PCM interface. A study on the
significant peaks were found for graphite, which is consistent with exact structure of this potentially new phase will be left as future
previous work [57,58]. The FTIR spectrum of GOx indicates work.
absorption bands of eOH at 3000-3500 cm1, COOH at 1721 cm1, We also measured the XRD peak broadening to analyze any
CeOH at 1355 cm1, and alkoxy CeOeC at 1047 cm1, which is microstructure changes in the paraffin matrix due to embedding
similar to the GOx spectrum obtained in Han et al.’s work [59]. The GOxA into paraffin, such as the grain-size change [67,68] and/or the
spectrum of GOx also shows a peak at 1588 cm1, representing the crystal-defect density change [68,69]. As shown in Supplementary
skeletal vibrations from unoxidized graphitic domains [60,61]. The Fig. S1b and Table S1, the peak widths (Full-widths-at-half-
GOx's carbonyl COOH peak at 1721 cm1 is not shown in GOxA's maximum) of one typical peak (q ¼ 10.7 ) for both the paraffin and
spectrum, confirming the partial reduction of GOx. The GOx's OeH GOxA-PCM are in fact identical (¼0.414 ). The same peak width
stretching peak at 3000-3500 cm1 and alkoxy CeOeC peak at implies that the microstructure (grain-size and crystal-defect
1047 cm1 are not shown in GOxA's spectrum, again confirming the density) has no significant difference between pure paraffin and
partial reduction of GOx [57,62]. In addition, the GOxA's spectrum the paraffin-matrix in the composite.
shows the epoxy CeOeC peak at 1209 cm1 and the C]C peak at Thus, based on the above analysis of FTIR and XRD, a potentially
1568 cm1, indicating an interaction between graphene sheets and new phase and/or bonding condition at the graphene-paraffin
carbohydrates as well as the assembly of GOx sheets into 3D interface may be generated by embedding graphene into paraffin.
structures [20].
3.3. Shape stabilization and mechanical characterization
3.2. Structure characterization of paraffin and GOxA-PCM
composite As shown in Supplementary Fig. S2, the GOxA-PCM composite
maintains its original shape even above the melting temperature
To investigate the effect of embedding GOxA into paraffin on the without any leakage of liquid paraffin from the system, proving

Fig. 1. Optical images of (a) GOx dispersion, (b) GOxH, and (c) GOxA. (A colour version of this figure can be viewed online.)
Y. Xu et al. / Carbon 114 (2017) 334e346 337

Fig. 2. SEM images of GOxA and GOxA-PCM. (a) GOxA (a single graphene oxide sheet in the center). GOxA-PCM composite in the (b) as-infiltrated and (c) polished conditions.

DSA, appropriate for many common materials from rubber band


with DSA z25 to shoe heel with DSA z80. For elastomeric materials,
durometer hardness D can be correlated to the modulus E roughly
by a power-law relationship [71]. However, this correlation cannot
be easily established in case of plastic indentation, because the
durometer intrinsically tests hardness instead of modulus.
Samples of paraffin, GOxA and GOxA-PCM which were 8 mm in
diameter and 10 mm high were prepared for durometer testing
using a Phase II PHT-960 digital Shore-A durometer. Paraffin and
GOxA-PCM samples were tested at different temperatures in the
range of 22  Ce67  C. The samples were directly put in a water bath
and the sample temperature T was recorded using an E-type ther-
mocouple through a digital OMEGA HH509 thermometer. The dry
GOxA samples were tested only at 22  C. The DSA readings for all
three materials changed with time after increasing the pressing
Fig. 3. FTIR spectra of graphite, GOx and GOxA. (A colour version of this figure can be time, indicating strain rate dependence. For consistency, the DSA
viewed online.) values were read ~3 s after applying the load.
Fig. 5a shows DSA vs. T for all materials. The GOxA-PCM com-
posite is significantly harder than pure paraffin at all temperatures,
excellent shape stabilization. In order to quantitatively analyze the
implying a significant reinforcement effect from embedding GOxA
fundamental mechanism of shape stabilization and investigate the
into the paraffin. It should be mentioned that the standard devia-
effect of embedding GOxA network into GOxA-PCM composite, we
tion in measured DSA for all samples over all temperatures is small
have mechanically tested paraffin, GOxA, and GOxA-PCM at the
(0.4e1.0) which ensures sample uniformity and durometer con-
macroscale (at different temperatures) using a durometer and at
sistency. It is interesting that GOxA-PCM's DSA reaches a plateau at
the nano/microscale using nanoindentation at different strain rates.
10.2 ± 0.7 for T > Tm ¼ 55.4  C (paraffin's melting temperature),
indicating a significant hardness for T > Tm and hence significant
3.3.1. Macroscale mechanical characterization using durometer at shape stabilization. GOxA-PCM's DSA at T > Tm matches well with
different temperatures the hardness (DSA ¼ 13.5 ± 0.8) of dry GOxA alone at 22  C, implying
3.3.1.1. Durometer testing at different temperatures. Paraffin, GOxA that the GOxA-PCM composite's hardness at T > Tm is due to the
and GOxA-PCM were mechanically characterized using a digital support from GOxA network alone as the paraffin is fully melted.
Shore-A durometer at different temperatures. The durometer test is
one common technique for simple, quick and semi-quantitative 3.3.1.2. Method for quantitative analysis of Shore-A durometer
evaluation of materials' resistance to penetration, i.e., hardness. testing. DSA is a hardness index, which cannot be directly compared
The durometer hardness D is standardized by ASTM D 2240 [70]. to the hardness measured by other techniques, such as nano-
One most common durometer scale is Shore A, designated here as indentation. Therefore, in order to have a more quantitative

Fig. 4. (a) FTIR spectra of GOxA, paraffin and GOxA-PCM composite. (b) XRD patterns of paraffin and GOxA-PCM composite. (A colour version of this figure can be viewed online.)
338 Y. Xu et al. / Carbon 114 (2017) 334e346

Fig. 5. (a) Durometer Shore A hardness DSA vs. testing temperature T. (b) Hardness H (calculated from DSA based on Eq. (5)) vs. T. Here, paraffin's Tm (¼55.4  C) and the GOxA's
properties at 22  C are also plotted. (c) The correlation of H and DSA based on Eq. (5). (A colour version of this figure can be viewed online.)

comparison, based on the mechanics of the durometer, we derive,  


for the first time in the literature, a methodology to correlate DSA to 0:55 þ 0:075DSA
H¼ MPa; where a
hardness, H. Here, H is commonly defined as the mean pressure pa2
8
under an indenter. For an axi-symmetric indenter, we have >
< 0:635 mm DSA < 70
   
¼ DSA (5)
>
: 0:395 þ 2:5 1  tan 17:5 mm DSA  70 :
H ¼ P=A and A ¼ pa2c ; (1) 100

where P is the indentation contact load (e.g., in Newton, or N), A is Fig. 5d shows the plot of H vs. DSA, which demonstrates a tran-
the projected contact area (e.g., in mm2), and ac is the contact radius sition at DSA ¼ 70 due to the truncated cone part of the tip (see
(e.g., in mm). Fig. 5c). For DSA<70, Eq. (5) is simply H ¼ 0.434 þ 0.059DSA (MPa),
As illustrated in Fig. 5c, the durometer test is essentially an i.e., a linear correlation (DSA<70). It should be noted that, as advised
indentation test using a truncated axi-symmetric flat punch tip by ASTM D 2240, the absolute value of DSA is reliable and compa-
[70,71], and the contact force P and indentation depth h are related rable between durometers only in the range of DSA ¼ 20e90 [70]
to the Shore-A durometer reading DSA by Ref. [70], due to the relatively large manufacturing tolerance requirement
of durometers (see Fig. 5c) [70].

P ¼ 0:55 þ 0:075DSA ; in N; (2)


3.3.1.3. Quantitative analysis of durometer testing for GOxA, paraffin,
and GOxA-PCM samples. With this correlation in Eq. (5), the DSA
h ¼ 2:5  0:025DSA ; in mm: (3)
values in Fig. 5a can be correlated to H and replotted vs. T as in
For an Shore-A durometer, the radius a of the cylindrical trun- Fig. 5b. Fig. 5a and b shows same qualitative trends as discussed
cated flat punch tip at a distance x from the tip end is [70] (see before. Fig. 5b shows that, at room temperature, the hardnesses of
Fig. 5c), paraffin and GOxA-PCM based on durometer are 6.06 ± 0.4 MPa
and 11.3 ± 0.8 MPa, respectively. These results match well with the
 corresponding values (6.27 ± 0.4 MPa and 15.7 ± 0.3 MPa, respec-
0:395 þ x tan 17:5 x  1:74mm
a¼ ; in mm: (4) tively) based on nanoindentation as discussed below. Here, based
0:635 x > 1:74mm
on nanoindentation data, the modulus to hardness ratio E/H are 177
When indenting a material with a high ratio of E to H with E/ and 71 for paraffin and GOxA-PCM, respectively, i.e., both are larger
H > ~40 [72e74], the indentation contact depth hc can be well than 40, so Eq. (5) can be used. This newly developed correlation
estimated by the indentation depth h. Consequently, as shown in methodology described by Eq. (5) enables us to quantitatively
Fig. 5c, the contact radius ac in Eq. (1) can be estimated by the analyze the durometer measurements, provided the aforemen-
physical radius a of the indenter tip at the contact boundary. Thus, tioned key assumptions are satisfied (E/H > 40 and DSA ¼ 20e90).
by replacing x in Eq. (4) by h, the final equation to correlate DSA to H, Fig. 5b also indicates that H ¼ 1.23 ± 0.3 MPa for dry GOxA at
is given in Eq. (5). 22  C and H ¼ 1.04 ± 0.3 MPa at T > Tm, again proving excellent
Y. Xu et al. / Carbon 114 (2017) 334e346 339

shape stabilization capability. It should be noted that the GOxA of <0.05 nm/s.
hardness (H ¼ 1.23 ± 0.3 MPa) based on durometer measurements All tests on paraffin and GOxA-PCM were performed using a
is much higher than the hardness (H ¼ 238e400Pa) resulting from diamond Berkovich indenter in the XP head. As demonstrated later,
the nanoindentation measurements discussed later. There may be paraffin and GOxA-PCM are strain-rate dependent, which to our
three major reasons for this difference: (1) for dry GOxA, best knowledge has not been previously studied. Also, as discussed
DSA ¼ 13.5 < 20, so the absolute value of DSA is not reliable; (2) the in Appendix 2, the GOxA-PCM composite can be effectively treated
indentation method is known to overestimate hardness of porous well as a homogeneous material for analyzing the indentation
materials due to indentation induced densification effect [75,76], results.
and here, GOxA is highly porous so that the densification effect is For a coherent and concise discussion, the general nano-
very large; and (3) for rapid compression of porous materials in air, indentation technique is summarized in Appendix 2, and a new
the air pressure built-up because of air trapping in the pores technique of multiple-strain-rate nanoindentation characterization
dominates the effective compression load [77,78]. Here, we believe is first-ever developed and presented in Appendix 3.
the 2nd and 3rd factors may be the dominant ones. In fact, the
readings of durometer measurement were ~3 s for ~2.2 mm 3.3.2.1. Nano/microscale multiple-strain-rate characterization of
penetration, while the flat punch nanoindentation tests took about paraffin and GOxA-PCM. A nanoindenter can control load P and
4e10 min for ~ 7 mm, so the compression/compaction effect under continuously measure the indentation displacement h. The inden-
durometer is probably about 300 times larger than that under tation strain rate ε_ is defined as the displacement-rate-to-
nanoindenter (the densification (2nd) factor), and the durometer displacement-ratio (_ε ¼ h=hz _ _
P=2P) [82,83]. A multiple-strain-rate
tip moved about five orders of magnitude faster than nanoindenter indentation method is used in this study: for one indentation, it
tip (the built-up air pressure (3rd) factor). In fact, when the has three sequentially decreasing strain rates (P=P¼2_ _ ε0 , 0:2_ε0 , and
durometer was sitting on GOxA, its DSA kept rapidly dropping from 0:02_ε0 ) during a single loading (see Fig. 6a). The methodology for
~18 to ~13.5, i.e., ~25%, within ~3s, and then, many GOxA samples multiple-strain-rate nanoindentation characterization is developed
would catastrophically break (similar to the case of GOxA nano- and discussed in Appendix 3 in detail.
indentation as discussed later). Thus, it is difficult to evaluate the Fig. 6a shows the typical load-time (P-t) curves for the two sets
further DSA change (after 3s). On the other hand, for paraffin and of three-strain-rate indentation. One ε_ 0 is equal to 0.05 s1 and the
GOxA-PCM samples, after an initial 5e10% dropping within 1s~2s, other is equal to 0.0025s1; this corresponds to six preset strain
their values of DSA quickly stabilized. _
rates, i.e., P=P¼ (0.025, 0.25, 0.5, 2.5, 5, 50)10e3 se1.
Nevertheless, at room temperature, even considering the Fig. 6b shows the indentation load-displacement (P-h) curves of
perhaps overestimated value (H ¼ 1.23 ± 0.3 MPa) for GOxA, the paraffin and GOxA-PCM. Here, for clear comparison, only one
H ¼ 11.3 ± 0.8 MPa of GOxA-PCM composite is still significantly typical P-h curve is shown for each loading condition for each
larger than the sum of H ¼ 6.06 ± 0.4 MPa of paraffin and _
material. Fig. 6b suggests a transition at P=Pz0.005s 1
: at faster
H ¼ 1.23 ± 0.3 MPa for GOxA, quantitatively. This reinforcement of rates, paraffin (black-dash) behaves almost the same as GOxA-PCM
embedding GOxA into paraffin is much beyond the prediction by (red-dash); at slower rates, paraffin's curve (black-solid) is much
the rule-of-mixture, implying a strong bonding between GOxA and lower. In fact, the 20-fold strain-rate drop induces a minimal
paraffin. Based on the FTIR, XRD, and mechanical characterization, behavior change for GOxA-PCM (red-dash / red-solid), while
since the microstructure of paraffin-matrix is the same as pure inducing a drastic drop for paraffin (black-dash / black-solid).
paraffin (as implied by the same XRD peak broadening (see Thus, paraffin's strain rate dependency is qualitatively much
Fig. S1b)), we hypothesize that a potentially new paraffin inter- more pronounced, especially at slower rates (P=P<~0.005s_ 1
).
phase is created at the surface of graphene oxide (GOx) sheet. This As discussed in Appendix 3, the quantitative characterization of
potentially new phase may act as a bonding agent (through prob- strain-rate dependence requires the steady state (SS) condition,
ably physical bonding, covalent bonding, hydrogen bonding, and/or corresponding to ε_ ¼ P=2P _ (see Eq. (A9)). Fig. 6ced show the
van der Waals bonding) to bond the GOx sheet and the rest of the measured strain rates (_ε ¼ h=h) _ _
and P=P vs. h for GOxA-PCM and
paraffin matrix. A study on the exact nature of the bonding would paraffin, respectively, corresponding to the loading schedules in
be left for a future work. Fig. 6a. The black curves in Fig. 6ced imply that the instrumental
More importantly, the GOxA-PCM composite is much harder control of P=P _ is very accurate at the intended values (P=P¼2_ _ ε0 ,
than pure paraffin at all temperatures and maintains significant 0:2_ε0 , and 0:02_ε0 ) for both samples; in fact, the complete transition
strength even at temperatures much higher than paraffin's melting time between any two preset rates is only 0.01e1.0 s (not shown
temperature. Therefore, this composite shows strong shape stabi- here).
lization capability. On the other hand, the red curves in Fig. 6c show that each
responded strain rate needs significant time to reach its steady-
3.3.2. Nano/microscale characterization state (SS) value (P=2P),_ suggesting the viscoelasticity nature of
A Keysight™ Nano Indenter™ G200, with two indenter heads: GOxA-PCM. The red curves in Fig. 6d demonstrate that, for pure
(1) XP and (2) Advance Dynamics Contact Module (DCM II) heads, paraffin, the strain rates reach the SS for the three larger values of
was used to perform the nanoindentation tests. The XP head has a _ but not for the three smaller ones. The comparison of Fig. 6ced
P=P
higher load capacity, while the DCM II head has a higher load res- shows that pure paraffin needs much longer time to achieve SS,
olution. The continuous stiffness measurement (CSM) technique indicating that paraffin is much more viscous compared to GOxA-
was implemented throughout each test [79e81], with a constant _
PCM, especially at slower loading rates (P=P<~0.005s 1
).
harmonic frequency (50 Hz for XP and 75 Hz for DCM II heads) and
harmonic displacement of 1.5 nm for both heads. 3.3.2.2. Analysis of strain-rate-dependent hardness for paraffin and
Due to GOxA's extremely low hardness and modulus, the GOxA GOxA-PCM. Fig. 7 shows the CSM modulus E and hardness H (see
was tested using two large flat punch indenters: (1) a 456 mm- Appendix 2 for general nanoindentation technique) vs. the
diameter diamond circular punch in DCM II, and (2) a 2 mm- _
measured strain rate (_ε ¼ h=h) as labeled individually. Here the sub-
diameter steel circular punch in XP head. The tests on GOxA were scripts show the materials and calculation method: “OP” designates
run with a constant load-rate-to-load-ratio of 0.05s1. Prior to any the Oliver-Pharr (OP) method, and “True” designates the true
indentation, the nanoindenter was stabilized to a thermal drift rate hardness method as described later. It may be surprising that, while
340 Y. Xu et al. / Carbon 114 (2017) 334e346

Fig. 6. (a) Load-time (P-t) curves for the two sets (_ε0 ¼ 0.05 s1 and 0.0025s1) of three-strain-rate testing. (b) Load-displacement (P-h) curves of paraffin and GOxA-PCM cor-
responding to the loading schedules in (a). (c) P=P _
_ and strain rate (_ε ¼ h=h) vs. h plots for GOxA-PCM and (d) that for paraffin. (A colour version of this figure can be viewed online.)

the CSM hardnesses of paraffin and GOxA-PCM are strain rate (1.570 ± 0.066 GPa) of pure-paraffin. This appears to contradict the
dependent, their CSM moduli are apparently independent of strain- known fact that graphene-reinforced organic materials are nor-
rate across almost four orders of magnitude (105~101s1). This mally stiffer than the pure matrix counterparts [85,86]. Since the
can be understood by the known fact that the CSM modulus is GOxA-PCM composite sample is very compact without any signif-
indeed the dynamic modulus and function of CSM harmonic fre- icant porosity (see Fig. 2c), we believe that this apparent contra-
quency [84] (constant 50 Hz for all tests). The CSM hardness rep- diction may be explained by the significant indentation pile-up on
resents the indentation mean pressure related to the current quasi- paraffin, which is discussed in Appendix 4. It is worth to note that
static strain rate. (see Appendix 4), the highly irregular indent shape for pure paraffin
It may be also surprising to see that, GOxA-PCM's Oliver-Pharr is first-ever observed for any homogeneous materials, probably due
(OP) modulus (1.111 ± 0.032 GPa) is about 40% lower than that to the underlying uneven plastic flow and local back-flow after
unloading of crystalline paraffin. As discussed in Appendix 5, we
could calculate the true hardness HTrue-Paraffin free of pile-up effect
for paraffin using a revised Joslin-Oliver method [87]. Here, since
the indentations on GOxA-PCM show no pile-up (see Appendix
Fig. A1a), the Oliver-Pharr hardness HOP-GOxA-PCM is directly used
for the true hardness of GOxA-PCM.
In Fig. 7, the HOP-GOxA-PCM curve plot two sets of hardness data for
GOxA-PCM: (1) the squares are the results obtained by averaging
over the six steady-states (SSs), and (2) the dense small triangles
are all individual data. We see a good match between the two sets
of data. This match implies that, when the indentation transits from
one SS to the sequential SS, the deformation is still in a quasi-steady
state (QSS) (instead of in a transient state). Therefore, the hardness
results are effectively steady-state results.
In Fig. 7, the HTrue-Paraffin curve also plots two data sets for
paraffin: (1) the squares of averaging over the three steady-states
(SSs) and the tails of the three slower loadings, and (2) the dense
Fig. 7. CSM modulus E and hardness H vs. strain rate for GOxA-PCM and paraffin. (A
individual data. This good matching between the two sets again
colour version of this figure can be viewed online.)
Y. Xu et al. / Carbon 114 (2017) 334e346 341

implies that all indentation data are effectively steady-state results. constraint factor, i.e., the ratio of H to strength s, is ~1.0 instead of
Moreover, Supplementary Figs. S3aeS3b show that, compared to ~3.0 [89].
paraffin, GOxA-PCM has much better indentation data consistency Even when using the large flat punch indenters which are much
and is a material with better-defined solid behavior (instead of bigger than the pore size (0.2~1 mm as shown in Fig. 2a), it is still
viscous behavior). very challenging to perform stable nanoindentations on GOxA due
By comparing paraffin (HTrue-paraffin) and GOxA-PCM (HOP-GOxA- to its extremely high porosity and extremely low mechanical
PCM) in Fig. 7, GOxA-PCM is always much harder than pure paraffin, properties. Fig. 8 shows the H and E calculated by Eq. (A1) and Eq.
and the difference increases by decreasing the strain rate. For (A3) vs. indentation depth h, for two typical indentations (one
example, the composite is 2.7 and 6.7 harder at ε_ ¼ 0.05s1 and indentation for each tip). It should be noted that when h reaches
104s1, respectively, suggesting a strong reinforcement effect due around 7 mm, both indentations become unstable, and the GOxA
to the GOxA. under the indenters were catastrophically collapsed, which
Fig. 7 also indicates that the two materials are strain rate generated large circular holes matching with the flat punch shapes
dependent as expected for organic materials. In this log-log plot of (see Supplementary Fig. S6).
HOP-GOxA-PCM vs. ε_ , the slope (i.e., the strain-rate sensitivity m) is During the stable stages of indentations as shown in Fig. 8, for
nicely a constant (~0.15) for all ε_ , corresponding to a constant both tips, while H (fP) increases continuously and monotonically, E
power-law exponent n z 6.7 for GOxA-PCM (see Appendix 3). (fS) interestingly steps up on multiple plateaus of ~2 mm steps.
However, for paraffin, by increasing ε_ from ~104 to ~101, m de- Since 2 mm is around the upper bound of pore size (see Fig. 2a), each
creases from ~0.45 to ~0.19, i.e., n increases from 2.2 to 5.2. step-up between E plateaus may be associated with collectively
Consequently, the change of n suggests that, the deformation crushing each effective-supporting layer of pores. The increase of H
mechanism for pure paraffin might transit from the molecule and E with increasing h may be because indentation technique is
diffusion (n < 3) at small strain rate to dislocation motion (n > 3) at known to overestimate H and E due to the densification effect for
large strain rates (see Appendix 3). However, due to the space porous materials [75,76]. On the other hand, the increase of E in
limitation, the investigation of deformation mechanism of paraffin plateau steps (instead of continuously increasing as for H) may be
and GOxA-PCM will be left for future work. due to that the CSM stiffness S is much more sensitive to the
Thus, compared to paraffin, GOxA-PCM has a smaller m, and stiffness of the local supporting densified structure. Also, while the
hence is much less strain-rate sensitive, especially at smaller strain values of H (black symbols) match fairly well using the two tips, the
rates. That is to say, by reinforcing paraffin using GOxA, the hard- values of E (red symbols) deviate about 4 times between the two
ness (strength) of the composite is not only much higher but also tips, probably due to the high heterogeneous local fluctuation in the
more stable against strain rate change, especially at smaller strain highly porous GOxA structure as well as the way of how the
rates, which is critical for practical applications. For example, based structure was densified under each indenter.
on Fig. 7 and Table 1, under a 0.27 MPa compression stress, it takes Due to the densification effect, Fig. 8 can be used to estimate the
paraffin only 3 h to deform 10% but would take GOxA-PCM almost upper bounds of E and H for GOxA alone. Table 2 lists the E and H
two centuries. Thus, the shape stabilization capability of GOxA- values at the middle of plateaus based on the number of supporting
PCM is drastically increased compared to pure paraffin. layers collectively penetrated. Moreover, after penetrating the 3rd
Thus, in summary, the paraffin is largely reinforced and shape collective layer, the GOxA under indentation was catastrophically
stabilized by the GOxA-network. In the following, similar to the collapsed, so that the E and H values at the 3rd-layer penetration
discussion based on durometer testing, we want to show that this provides the upper-bound estimates for GOxA's modulus and fail-
reinforcement cannot be explained by the rule-of-mixture, ure strength sf (noting H/s z 1.0 as aforementioned [89]), i.e.,
implying a strong bonding between GOxA and paraffin. E z 75e389 kPa and sf z H z 238e400 Pa for GOxA alone. These
values fall in the ranges for GOxA alone based on previous studies
3.3.2.3. Nano/microscale characterization of graphene oxide aerogel (see Appendix Table A1): 2e1200 kPa for modulus and
(GOxA) alone. In order to test the mechanical properties of GOxA 200e40000Pa for hardness.
alone using nanoindentation, two large circular flat punch in- However, Fig. 7 demonstrates that the GOxA-PCM is at least 2.7
denters were used: a 2 mm-diameter steel tip and a 456 mm- times harder than pure paraffin whose hardness is in the range of
diameter diamond tip. Here, for any flat punch indenter, the contact
area A is a known constant, and the hardness H and modulus E can
be determined using Eq. (A1)e(A3) in Appendix 2. Here, the contact
radius ac is known.
As demonstrated later, since GOxA's modulus E (75e400 kPa) is
6e7 orders of magnitude lower than that (1045 GPa for diamond or
200 GPa for steel) of indenters, Eq. (A3) can be nicely applied.
Equations A1 and A3 indicate that, when a cylindrical flat punch
indenter (of constant ac) is used, H and E are simply proportional to
P and S, respectively. Here, to use Eq. (A3), GOxA's n can be esti-
mated to be ~0.2 [88]. Also, for highly porous materials, the

Table 1
The duration L10 for 10% accumulative strain for paraffin and GOxA-PCM under two
values of uniaxial stress su (See Appendix 3 and Eq. (A10) for the discussion).

Material su (MPa) ε_ u (s1) L10 Duration to 10% strain

Paraffin 1.67 5.4  103 19 s


Fig. 8. H and E calculated by Eq. (A1) and Eq. (A3) vs. h, for two typical indentations on
GOxA-PCM 1.67 2.7  106 8.8 h
GOxA (one for each tip). Here, because the intrinsic noise of XP head (2 mm tip) is
Paraffin 0.27 9.0  106 3h
much higher, the associated E data was smoothened and plotted on the background of
GOxA-PCM 0.27 ~1.8  1011 ~190 year
the original unaltered data. (A colour version of this figure can be viewed online.)
342 Y. Xu et al. / Carbon 114 (2017) 334e346

Table 2
The E and H values at the middle of plateaus based on Fig. 8 and the number of layers collectively penetrated.

Indenter Layer# penetrated h at the middle of plateau (mm) E (kPa) H (Pa)

2 mm tip 1st ~1.0 143 43


456 mm tip 1st ~1.0 30 30
2 mm tip 2nd ~4.0 289 200
456 mm tip 2nd ~4.0 52 144
2 mm tip 3rd ~6.0 389 400
456 mm tip 3rd ~6.0 75 238

0.85e6 MPa, while GOxA-PCM is almost four orders of magnitude Acknowledgements


harder than GOxA alone. Thus, the reinforcement due to GOxA in
paraffin is drastically higher than the prediction of the rule-of- The authors appreciate the support of the National Science
mixture, indicating a strong bonding between GOxA and paraffin. Foundation (CBET-1235769). Any opinions, findings, and conclu-
This is qualitatively consistent with the aforementioned results sions or recommendations expressed in this material are those of
based on durometer tests. the author(s) and do not necessarily reflect the views of the Na-
tional Science Foundation. We thank Nancy Peltier in Biology
4. Conclusions Department at Villanova for her help with the supercritical CO2
drying. We also appreciate the comprehensive comments and
A 5 wt% graphene oxide aerogel (GOxA) reinforced paraffin-PCM constructive suggestions from the reviewers of this paper.
composite is developed. The composite maintains its original shape
above melting temperature of the PCM and to hold the molten Appendix 1. Literature survey on the mechanical properties
paraffin without any leakage, indicating superior shape of porous 3D carbon aerogels/networks.
stabilization.
The X-ray diffraction and Fourier transform infrared (FTIR)

Table A1
The mechanical properties of porous 3D carbon aerogels/networks.

Materials Processing/Material condition Density (mg/cm3) Strength s (kPa) Modulus E (kPa)

GA [20] Pristine 3e5 0.2e3 (decreasing with more testing cycles) 10e16
25  C ammonia treated GA [90] Ammonia treatment, but closer to pristine ~16 0.3 2
90  C ammonia treated GA [90] Ammonia treatment, further reinforced ~16 0.7 4.8
3D self-assembly of GO [51] Nobel-metal and glucose assistance ~30 42 260
CNT-graphene hybrid aerogel [91] Acid-treated CNT 32 30 (at 3% strain) 1200
CNT-graphene hybrid aerogel [91] Pristine CNT 41 70 (at 3% strain) 2900
GA [26] Pristine 30.9 40 1200
GA [26] Pristine 47.6 150 2800
GA [26] Pristine 96.1 660 6200
Graphene Assembly [52] Covalently cross-link GO 80e100 10000 51000

spectra imply a potentially new phase at the GOxA-paraffin inter- Appendix 2. Nanoindentation characterization
face in the composite. Durometer testing indicates that the com-
posite is much harder than paraffin at both room temperature and Generally, nanoindentation can be used to determine the
elevated temperatures and maintains significant strength above hardness H and modulus E of the indented material using Eqs.
paraffin's melting temperature. Nanoindentation testing shows (A1)e(A3) [79,92,93].
that the composite is 3e7 times harder and less strain-rate sen-
sitive than paraffin, especially at smaller strain rates. Thus, the P P
H¼ ¼ ; (A1)
composite is not only much harder but also much more stable A pa2c
against strain rate change, which is critical for practical applica-
tions. Furthermore, the durometer and nanoindentation tests both
S 1 1  n2 1  n2i
quantitatively indicate that the reinforcement of paraffin in GOxA is Er ¼ ; where ¼ þ ; (A2)
2ac Er E Ei
beyond the prediction by the rule-of-mixture, implying a strong
GOxA-PCM interfacial bonding.
To our best knowledge, this is the first study on the fundamental   S
E ¼ 1  n2 if Ei > > E; (A3)
mechanical behavior of paraffin and GOxA-PCM composite at 2ac
different temperatures and different strain rates, providing critical
insights into their behaviors. The relationship between the hard- where P is the indentation load; A and ac are the contact area and
ness and durometer index first-ever developed here will enable the radius, respectively. S is the contact stiffness. Er is the reduced
quantitative durometer testing on materials for many other appli- modulus as defined in Eq. (A2), where E and n are the modulus and
cations at different ambient conditions (e.g., temperatures and Poisson's ratio of the indented material, while Ei and ni are the
probably chemical environments) due to its versatile and simple quantities of the indenter.
nature. A nanoindenter can control P and continuously measure the
indentation displacement h. With the continuous stiffness mea-
surement (CSM) technique [79,92,93], a nanoindenter can also
Y. Xu et al. / Carbon 114 (2017) 334e346 343

continuously measure S.
For flat homogenous materials without pile-up, the Oliver-Pharr ε_ ¼ A0 H n ;
method can be used to determine the contact radius ac through P, h, alternatively
S using Eqs. (A4) and (A5) [79,92,93].
H ¼ B0 ε_ m (A7)
P
hc ¼ h  0:75 ; (A4)
S where A' and B' are fitting parameters.
Nanoindenter intrinsically controls P (instead of h). Thus, it is
Ac ¼ pa2c ¼ f ðhc Þ; (A5) much convenient to directly control the loading rate to alter the
strain rate. For a self-similar indenter (e.g., a perfect Berkovich or
where hc is the contact depth. Here, Eq. (A5) is called area function, conical indenter) indenting a homogeneous material, by neglecting
which can be calibrated using standard materials with known indentation size effect [98,99], the contact area A is proportional to
moduli or directly determined by visualizing the indenter shape h2. Thus, A ¼ Ch2, where C is a constant related to indenter geom-
using microscopy [79,92,93]. etry. By differentiating H¼P/A ¼ P/(Ch2), we have [100].
It should be noted that the measured indentation properties
(modulus or hardness) are due to the average responses from the dH dP dh h_ P_ H_
indentation-deformed volume (plastic zone) [72,73,89]. Based on ¼  2 ; i:e; 2_ε ¼ 2 ¼  ; (A8)
H P h h P H
Johnson's expanding cavity model, the plastic zone is normally
estimated by a hemisphere with a diameter 1~2 of the contact
.
2_ε ¼ P_ P if H_ ¼ 0: (A9)
diameter d [72,73,89].
As shown in Fig. 2a, 1.0 mm would be a good estimate for the Because a constant H (z3s) is expected for a material during a
upper bound of the pore-size in GOxA and hence used as the steady state deformation, based on Eq. (A9), we have 2_ε ¼ P=P. _
characteristic microstructure length scale (lu) in the GOxA-PCM Therefore, a constant ε_ test can be approximated by running con-
composite. Our indentation results shown in Fig. 7 are always stant load-rate-to-load-ratio P=P _ (¼2_ε) [100], i.e., exponentially
from h > 5 mm, namely, h > 5lu. More importantly, for a Berkovich increasing the load.
indenter, the contact diameter d z 7h > 35 mm. Thus, any h > 5 mm However, in practice, an indentation with one slow strain rate
indentation corresponds to a plastic zone with diameter >35 mm, (104~105s1) would take multiple days. Consequently, a constant
i.e., >35lu. thermal drift rate - a key assumption for quantitative indentation
Thus, the indented material in the GOxA-PCM composite can be analysis - may not be valid [101]. Therefore, a multiple-strain-rate
effectively treated well as a homogeneous material for the results in indentation technique is developed here, particularly with three
Fig. 7. sequentially decreasing strain rates (P=P¼2__ ε0 , 0:2_ε0 , and 0:02_ε0 )
during a single loading (see Fig. 6a) to minimize its duration (<4 h).
Appendix 3. Multi-strain-rate characterization using The loading portion at each strain rate lasts about 1/3 of the
nanoindentation maximum depth. Two sets of three-strain-rate indentations were
performed with ε_ 0 equals 0.05 s1 and 2.5  103s1 (see Fig. 6a) in
For strain-rate-dependent materials, their strengths depend on this study, corresponding to six preset strain rates, i.e., (0.025, 0.25,
not only the strain rate but also the deformation history. For 0.5, 2.5, 5, 50)103 s1 .
example, after instantaneously applying a constant strain rate (_εu ), An interesting and practical observation should be mentioned:
the required stress would transiently change and finally reach to a the deformation during each entire three-strain-rate indentation is
constant steady-state (SS) value. Thus, only at the steady-state, a almost always at either the steady state (SS) or quasi-steady state
strain rate can be uniquely correlated to the applied stress, namely (QSS). This is because the slow and smooth hardness change during
flow strength (s). In order to quantitatively study the strain-rate- the transition between each pair of preset strain rates. This has
dependence, the mechanical properties should be measured at been demonstrated for both paraffin and GOxA-PCM composite.
the steady state (SS). The SS (and QSS) deformation during the entire multi-strain-rate
A phenomenological power-law relationship is commonly used testing provides a significant benefit: the SS (and QSS) ε_ (in Eq.
to describe the strain-rate dependence of flow strength at the SS, (A7)) spans continuously over a wide range (see Fig. 7) instead of
only the preset values (e.g., 3 strain rates for each three-strain-rate
ε_ u ¼ Asn or alternatively s ¼ B_εm
u; (A6)
testing).
It may also be important in practical considerations to predict
where ε_ u and s are the strain rate and stress during a uniaxial
the life L of a stressed structural part made of strain-rate sensitive
testing at the SS; n, m, A and B are fitting parameters, while m is
material. We may define the expected life L10 under a constant
called the strain rate sensitivity, and n (¼1/m) is called the power-
uniaxial stress su as the duration for reaching a preset critical cu-
law exponent. The value of n normally correlates to the deforma-
mulative strain εc, say 10% (i.e., εc ¼ 0.1), due to the corresponding
tion mechanism, e.g., for crystals, n < 3 normally indicates diffusive
steady-state strain rate ε_ u . Noticing that L10 ¼ εc =_εu ¼ 0:1=_εu , based
deformation, while n > 3 normally indicates dislocational defor-
on Eq. (A7), we have
mation [94].
The indentation strain rate ε_ is defined as the displacement-  
_ 10 B0 1=m
rate-to-displacement-ratio (_ε ¼ h=h) [82,83]. Due to the strong L10 ¼ : (A10)
strain gradient and stress concentration under indentation, the 9 3su
equivalent uniaxial strain ε_ u z0:09_ε [95], while s zH/3 [96,97]. Here, B' and m can be found experimentally by fitting H vs. ε_
Thus, Eq. (A6) can be revised to [94]. 0
using H ¼ B ε_ m (see Fig. 7). For example, for HOP-GOxA-PCM shown in
Fig. 7, the fitting parameters are B’ ¼ 23.35 MPa and m ¼ 0.15, and
Hn
ε_ ¼ A ; alternatively H ¼ 3  0:09m B_εm ; hence L10 e6.0  109 s z 190 year for su ¼ 0.27 MPa based on Eq.
0:09  3n (A10). L10 for paraffin and GOxA-PCM under two values of su are
Or simply, listed in Table 1.
344 Y. Xu et al. / Carbon 114 (2017) 334e346

Appendix 4. Irregular localized pile-ups around each pile-up [79,92,93]. In order to determine the true hardness free of
indentation on Paraffin the pile-up effect, it is essential to determine the true contact area
A. Two existing methods to correct for the pile-up effect are (1)
Fig. A1a shows the optical image of a typical indent on GOxA- estimate A at the maximum indentation depth by directly
PCM, displaying flat indent surfaces and straight edges without measuring the residual indent's area using an appropriate micro-
pile-up. Thus, we expect that the Oliver-Pharr method can be used scopy [103,104], and (2) a revised Joslin-Oliver method [87] as
to determine the modulus and hardness well for GOxA-PCM discussed below.
composite. On the contrary, Fig. A1b shows a typical indent on Unfortunately, it is very difficult to have a precise estimate for
paraffin with bumpy indent surfaces and wavy indent edges, which the indent area due to the strong indent irregularity and possible
is the case for every indentation on paraffin (also see the SEM image plastic back-flow. Therefore, we could use a revised Joslin-Oliver
in Supplementary Fig. S4). To our best knowledge, this highly method [87] (see Eq. (A11)) to estimate the true hardness HTrue
irregular indent shape is first-ever observed for any homogeneous free of the pile-up effect, provided we know the true modulus ETrue.
materials. This is probably due to the underlying uneven plastic
flow and local back-flow (resulted from the residual stress field in 2 H
ETrue OP
the permanent indent) after unloading of crystalline paraffin. HTrue ¼ 2
; (A11)
The investigation of this surprising irregular indentation plas- EOP
ticity of paraffin is not the focus of this paper and hence will be left
for a future work. Preliminary investigation using ex-situ optical where HOP and EOP are the Oliver-Pharr hardness and modulus,
microscopy on post-indentation morphology shows that the indent respectively. Here, as aforementioned, the GOxA-PCM is expected
irregularity happens almost spontaneously (within 2 s); thus, the to be slightly stiffer than the paraffin, and the GOxA-PCM com-
recrystallization of paraffin crystal in the indented area may be posite's Oliver-Pharr modulus (EOP-GOxA-PCM ¼ 1.111 ± 0.032 GPa in
ruled out as the key reason as it requires diffusion. Fig. 7) would be free of pile-up effect, which could be a good upper
bound estimate for paraffin's true modulus, i.e., ETrue-paraffin

Fig. A1. Optical images of (a) a typical 30 mm-deep indent on GOxA-PCM and (b) one on paraffin. (c) and (d) The magnified optical images of the boxed region in (b), where the focal
depths are (c) 0 mm and (d) 8 mm, respectively, showing two large ~8 mm-tall localized pile-ups.

Due to the uneven plastic flow, Fig. A1b-A1d interestingly z1.111 GPa. Thus, Eq. (A11) can be used to calculate HTrue-paraffin (a
indicate multiple localized pile-ups (instead of the common form of good upper bound) as plotted in Fig. 7.
single pile-up [93,102]) on each indent edge. For example, Fig. A1b-
A1d show that, an edge of a 30 mm indentation has two large Appendix A. Supplementary data
~8 mm-tall localized pile-ups which did not exist prior to the
indentation (not shown here). Moreover, in order to further Supplementary data related to this article can be found at http://
confirm the existence of pile-ups for paraffin, additional in- dx.doi.org/10.1016/j.carbon.2016.11.069.
dentations were done using a 300 mm sapphire spherical indenter
(see Supplementary Fig. S5). Those indents clearly show multiple
localized pile-ups. Unfortunately, as shown in Fig. A1b and References
Supplementary S4, it is very difficult to have a precise estimate for [1] J.F. Li, W. Lu, Y.B. Zeng, Z.P. Luo, Simultaneous enhancement of latent heat
the indent area due to the strong indent irregularity, and more and thermal conductivity of docosane-based phase change material in the
importantly, the indent may be altered significantly from the con- presence of spongy graphene, Sol. Energy Mater. Sol. Cells 128 (0) (2014)
48e51.
tact under load due to the possible aforementioned plastic back-
[2] R. Ehid, A.S. Fleischer, Development and characterization of paraffin-based
flow (see Fig. A1b and Supplementary S4). shape stabilized energy storage materials, Energy Convers. Manag. 53 (1)
(2012) 84e91.
[3] G.-Q. Qi, C.-L. Liang, R.-Y. Bao, Z.-Y. Liu, W. Yang, B.-H. Xie, M.-B. Yang,
Appendix 5. Calculation of true hardness free of the pile-up Polyethylene glycol based shape-stabilized phase change material for ther-
mal energy storage with ultra-low content of graphene oxide, Sol. Energy
effect for paraffin
Mater. Sol. Cells 123 (2014) 171e177.
[4] M. Mehrali, S.T. Latibari, M. Mehrali, H.S.C. Metselaar, M. Silakhori, Shape-
As discussed in Appendix 4, significant pile-ups are observed stabilized phase change materials with high thermal conductivity based on
around each indentation on paraffin. The Oliver-Pharr method paraffin/graphene oxide composite, Energy Convers. Manag. 67 (0) (2013)
275e282.
based on Eq. (A4) assumes no pile-up, and it underestimates the [5] J.-L. Zeng, S.-H. Zheng, S.-B. Yu, F.-R. Zhu, J. Gan, L. Zhu, Z.-L. Xiao, X.-Y. Zhu,
contact depth hc and hence the contact area A in case of indentation Z. Zhu, L.-X. Sun, Z. Cao, Preparation and thermal properties of palmitic acid/
Y. Xu et al. / Carbon 114 (2017) 334e346 345

polyaniline/exfoliated graphite nanoplatelets form-stable phase change Chem. A 2 (23) (2014) 8941e8951.
materials, Appl. Energy 115 (2014) 603e609. [34] F. Irin, S. Das, F.O. Atore, M.J. Green, Ultralow percolation threshold in aerogel
[6] M.M. Farid, A.M. Khudhair, S.A.K. Razack, S. Al-Hallaj, A review on phase and cryogel templated composites, Langmuir 29 (36) (2013) 11449e11456.
change energy storage: materials and applications, Energy Convers. Manag. [35] R.J. Warzoha, A.S. Fleischer, Effect of graphene layer thickness and me-
45 (9e10) (2004) 1597e1615. chanical compliance on interfacial heat flow and thermal conduction in
[7] M. Xiao, B. Feng, K. Gong, Preparation and performance of shape stabilized solideliquid phase change materials, ACS Appl. Mater. Interfaces 6 (15)
phase change thermal storage materials with high thermal conductivity, (2014) 12868e12876.
Energy Convers. Manag. 43 (1) (2002) 103e108. [36] R.J. Warzoha, R.M. Weigand, A.S. Fleischer, Temperature-dependent thermal
[8] M.-Y. Shen, T.-Y. Chang, T.-H. Hsieh, Y.-L. Li, C.-L. Chiang, H. Yang, M.-C. Yip, properties of a paraffin phase change material embedded with herringbone
Mechanical properties and tensile fatigue of graphene nanoplatelets rein- style graphite nanofibers, Appl. Energy 137 (2015) 716e725.
forced polymer nanocomposites, J. Nanomater. 2013 (2013) 9. [37] R.J. Warzoha, A.S. Fleischer, Effect of carbon nanotube interfacial geometry
[9] Y. Zhong, M. Zhou, F. Huang, T. Lin, D. Wan, Effect of graphene aerogel on on thermal transport in solideliquid phase change materials, Appl. Energy
thermal behavior of phase change materials for thermal management, Sol. 154 (2015) 271e276.
Energy Mater. Sol. Cells 113 (0) (2013) 195e200. [38] J. Yang, E. Zhang, X. Li, Y. Zhang, J. Qu, Z.-Z. Yu, Cellulose/graphene aerogel
[10] T.F. Baumann, A.E. Gash, S.C. Chinn, A.M. Sawvel, R.S. Maxwell, J.H. Satcher, supported phase change composites with high thermal conductivity and
Synthesis of high-surface-area alumina aerogels without the use of alkoxide good shape stability for thermal energy storage, Carbon 98 (2016) 50e57.
precursors, Chem. Mater. 17 (2) (2005) 395e401. [39] S. Tahan Latibari, M. Mehrali, M. Mehrali, T.M. Indra Mahlia, H.S. Cornelis
[11] M. Moner-Girona, A. Roig, E. Molins, J. Llibre, Sol-gel route to direct forma- Metselaar, Synthesis, characterization and thermal properties of nano-
tion of silica aerogel microparticles using supercritical solvents, J. Sol-Gel Sci. encapsulated phase change materials via solegel method, Energy 61 (2013)
Technol. 26 (1) (2003) 645e649. 664e672.
[12] S.M. Jung, H.Y. Jung, M.S. Dresselhaus, Y.J. Jung, J. Kong, A facile route for 3D [40] M. Mehrali, S. Tahan Latibari, M. Mehrali, T.M.I. Mahlia, H.S. Cornelis Met-
aerogels from nanostructured 1D and 2D materials, Sci. Rep. 2 (2012) 849. selaar, Effect of carbon nanospheres on shape stabilization and thermal
[13] Y. Tang, K.L. Yeo, Y. Chen, L.W. Yap, W. Xiong, W. Cheng, Ultralow-density behavior of phase change materials for thermal energy storage, Energy
copper nanowire aerogel monoliths with tunable mechanical and electrical Convers. Manag. 88 (2014) 206e213.
properties, J. Mater. Chem. A 1 (23) (2013) 6723e6726. [41] Z. Zhang, X. Fang, Study on paraffin/expanded graphite composite phase
[14] S.M. Jung, D.J. Preston, H.Y. Jung, Z. Deng, E.N. Wang, J. Kong, Porous Cu change thermal energy storage material, Energy Convers. Manag. 47 (3)
nanowire aerosponges from one-step assembly and their applications in (2006) 303e310.
heat dissipation, Adv. Mater. 28 (7) (2016) 1413e1419. [42] Z. Sun, W. Kong, S. Zheng, R.L. Frost, Study on preparation and thermal en-
[15] W. He, G. Li, S. Zhang, Y. Wei, J. Wang, Q. Li, X. Zhang, Polypyrrole/silver ergy storage properties of binary paraffin blends/opal shape-stabilized phase
coaxial nanowire aero-sponges for temperature-independent stress sensing change materials, Sol. Energy Mater. Sol. Cells 117 (0) (2013) 400e407.
and stress-triggered joule heating, ACS Nano 9 (4) (2015) 4244e4251. [43] J.-N. Shi, M.-D. Ger, Y.-M. Liu, Y.-C. Fan, N.-T. Wen, C.-K. Lin, N.-W. Pu,
[16] A. Mikhalchan, Z. Fan, T.Q. Tran, P. Liu, V.B.C. Tan, T.-E. Tay, H.M. Duong, Improving the thermal conductivity and shape-stabilization of phase change
Continuous and scalable fabrication and multifunctional properties of carbon materials using nanographite additives, Carbon 51 (2013) 365e372.
nanotube aerogels from the floating catalyst method, Carbon 102 (2016) [44] S. Wi, J. Seo, S.-G. Jeong, S.J. Chang, Y. Kang, S. Kim, Thermal properties of
409e418. shape-stabilized phase change materials using fatty acid ester and exfoliated
[17] T. Horikawa, J.i. Hayashi, K. Muroyama, Size control and characterization of graphite nanoplatelets for saving energy in buildings, Sol. Energy Mater. Sol.
spherical carbon aerogel particles from resorcinoleformaldehyde resin, Cells 143 (2015) 168e173.
Carbon 42 (1) (2004) 169e175. [45] M. Mehrali, S.T. Latibari, M. Mehrali, T.M. Indra Mahlia, H.S. Cornelis Met-
[18] M.A. Worsley, J.H. Satcher, T.F. Baumann, Synthesis and characterization of selaar, M.S. Naghavi, E. Sadeghinezhad, A.R. Akhiani, Preparation and char-
monolithic carbon aerogel nanocomposites containing double-walled carbon acterization of palmitic acid/graphene nanoplatelets composite with
nanotubes, Langmuir 24 (17) (2008) 9763e9766. remarkable thermal conductivity as a novel shape-stabilized phase change
[19] T.Q. Tran, Z. Fan, P. Liu, H.M. Duong, Advanced morphology-controlled material, Appl. Therm. Eng. 61 (2) (2013) 633e640.
manufacturing of carbon nanotube fibers, thin films and aerogels from aer- [46] X. Zhao, Q. Zhang, D. Chen, P. Lu, Enhanced mechanical properties of
ogel technique, in: Asia Pacific Confederation of Chemical Engineering graphene-based poly(vinyl alcohol) composites, Macromolecules 43 (5)
Congress 2015, Melbourne: Engineers Australia, 2015, pp. 2444e2451. (2010) 2357e2363.
[20] H. Hu, Z. Zhao, W. Wan, Y. Gogotsi, J. Qiu, Ultralight and highly compressible [47] C.-C. Ji, M.-W. Xu, S.-J. Bao, C.-J. Cai, Z.-J. Lu, H. Chai, F. Yang, H. Wei, Self-
graphene aerogels, Adv. Mater. 25 (15) (2013) 2219e2223. assembly of three-dimensional interconnected graphene-based aerogels and
[21] L. Qiu, D. Liu, Y. Wang, C. Cheng, K. Zhou, J. Ding, V.-T. Truong, D. Li, Me- its application in supercapacitors, J. Colloid Interface Sci. 407 (0) (2013)
chanically robust, electrically conductive and stimuli-responsive binary 416e424.
network hydrogels enabled by superelastic graphene aerogels, Adv. Mater. [48] J.A. King, D.R. Klimek, I. Miskioglu, G.M. Odegard, Mechanical properties of
26 (20) (2014) 3333e3337. graphene nanoplatelet/epoxy composites, J. Appl. Polym. Sci. 128 (6) (2013)
[22] Y. Xu, K. Sheng, C. Li, G. Shi, Self-assembled graphene hydrogel via a one-step 4217e4223.
hydrothermal process, ACS Nano 4 (7) (2010) 4324e4330. [49] N. Norhakim, S.H. Ahmad, C.H. Chia, N.M. Huang, Mechanical and thermal
[23] K. Zhang, L.L. Zhang, X.S. Zhao, J. Wu, Graphene/polyaniline nanofiber properties of graphene oxide filled epoxy nanocomposites, Sains Malays. 43
composites as supercapacitor electrodes, Chem. Mater. 22 (4) (2010) (4) (2014) 603e609.
1392e1401. [50] D. Cai, M. Song, A simple route to enhance the interface between graphite
[24] G. Tang, Z.-G. Jiang, X. Li, H.-B. Zhang, A. Dasari, Z.-Z. Yu, Three dimensional oxide nanoplatelets and a semi-crystalline polymer for stress transfer,
graphene aerogels and their electrically conductive composites, Carbon 77 Nanotechnology 20 (31) (2009) 315708.
(0) (2014) 592e599. [51] Z. Tang, S. Shen, J. Zhuang, X. Wang, Noble-metal-promoted three-
[25] Y. Qian, I.M. Ismail, A. Stein, Ultralight, high-surface-area, multifunctional dimensional macroassembly of single-layered graphene oxide, Angew.
graphene-based aerogels from self-assembly of graphene oxide and resol, Chem. Int. Ed. 49 (27) (2010) 4603e4607.
Carbon 68 (0) (2014) 221e231. [52] M.A. Worsley, S.O. Kucheyev, H.E. Mason, M.D. Merrill, B.P. Mayer, J. Lewicki,
[26] X. Zhang, Z. Sui, B. Xu, S. Yue, Y. Luo, W. Zhan, B. Liu, Mechanically strong and C.A. Valdez, M.E. Suss, M. Stadermann, P.J. Pauzauskie, J.H. Satcher, J. Biener,
highly conductive graphene aerogel and its use as electrodes for electro- T.F. Baumann, Mechanically robust 3D graphene macroassembly with high
chemical power sources, J. Mater. Chem. 21 (18) (2011) 6494e6497. surface area, Chem. Commun. 48 (67) (2012) 8428e8430.
[27] M.A. Worsley, P.J. Pauzauskie, T.Y. Olson, J. Biener, J.H. Satcher, T.F. Baumann, [53] W.S. Hummers, R.E. Offeman, Preparation of graphitic oxide, J. Am. Chem.
Synthesis of graphene aerogel with high electrical conductivity, J. Am. Chem. Soc. 80 (6) (1958) 1339.
Soc. 132 (40) (2010) 14067e14069. [54] Y.-R. Huang, Y. Jiang, J.L. Hor, R. Gupta, L. Zhang, K.J. Stebe, G. Feng,
[28] M.B. Lim, M. Hu, S. Manandhar, A. Sakshaug, A. Strong, L. Riley, K.T. Turner, D. Lee, Polymer nanocomposite films with extremely high
P.J. Pauzauskie, Ultrafast solegel synthesis of graphene aerogel materials, nanoparticle loadings via capillary rise infiltration (CaRI), Nanoscale 7 (2)
Carbon 95 (2015) 616e624. (2015) 798e805.
[29] H. Huang, P. Chen, X. Zhang, Y. Lu, W. Zhan, Edge-to-edge assembled gra- [55] Y. Cho, S.Y. Lee, L. Ellerthorpe, G. Feng, G. Lin, G. Wu, J. Yin, S. Yang, Elas-
phene oxide aerogels with outstanding mechanical performance and su- toplastic inverse opals as power-free mechanochromic sensors for force
perhigh chemical activity, Small 9 (8) (2013) 1397e1404. recording, Adv. Funct. Mater. 25 (38) (2015) 6041e6049.
[30] S. Ye, Q. Zhang, D. Hu, J. Feng, Core-shell-like structured graphene aerogel [56] S. Lowell, J.E. Shields, M.A. Thomas, M. Thommes, Characterization of Porous
encapsulating paraffin: shape-stable phase change material for thermal en- Solids and Powders: Surface Area, Pore Size and Density, Springer Science &
ergy storage, J. Mater. Chem. A 3 (7) (2015) 4018e4025. Business Media, 2012.
[31] S.T. Nguyen, H.T. Nguyen, A. Rinaldi, N.P.V. Nguyen, Z. Fan, H.M. Duong, [57] L. Zhang, G. Chen, M.N. Hedhili, H. Zhang, P. Wang, Three-dimensional as-
Morphology control and thermal stability of binderless-graphene aerogels semblies of graphene prepared by a novel chemical reduction-induced self-
from graphite for energy storage applications, Colloids Surfaces A Phys- assembly method, Nanoscale 4 (22) (2012) 7038e7045.
icochem. Eng. Aspects 414 (0) (2012) 352e358. [58] R. Kumar, M. Mamlouk, K. Scott, A graphite oxide paper polymer electrolyte
[32] J. Liang, Z. Cai, L. Li, L. Guo, J. Geng, Scalable and facile preparation of gra- for direct methanol fuel cells, Int. J. Electrochem. 2011 (2011) 7.
phene aerogel for air purification, RSC Adv. 4 (10) (2014) 4843e4847. [59] D. Han, L. Yan, Supramolecular hydrogel of chitosan in the presence of gra-
[33] Q. Fang, B. Chen, Self-assembly of graphene oxide aerogels by layered double phene oxide nanosheets as 2D cross-linkers, ACS Sustain. Chem. Eng. 2 (2)
hydroxides cross-linking and their application in water purification, J. Mater. (2014) 296e300.
346 Y. Xu et al. / Carbon 114 (2017) 334e346

[60] E.-Y. Choi, T.H. Han, J. Hong, J.E. Kim, S.H. Lee, H.W. Kim, S.O. Kim, Non- measurement technique and its applications, Mater. Charact. 48 (1) (2002)
covalent functionalization of graphene with end-functional polymers, 11e36.
J. Mater. Chem. 20 (10) (2010) 1907e1912. [81] J. Hay, P. Agee, E. Herbert, Continuous stiffness measurement during
[61] N.G. Sahoo, H. Bao, Y. Pan, M. Pal, M. Kakran, H.K.F. Cheng, L. Li, L.P. Tan, instrumented indentation testing, Exp. Tech. 34 (3) (2010) 86e94.
Functionalized carbon nanomaterials as nanocarriers for loading and de- [82] B.N. Lucas, W.C. Oliver, Indentation power-law creep of high-purity indium,
livery of a poorly water-soluble anticancer drug: a comparative study, Chem. Metallurgical Mater. Trans. A 30 (3) (1999) 601e610.
Commun. 47 (18) (2011) 5235e5237. [83] J.M. Wheeler, V. Maier, K. Durst, M. Go€ken, J. Michler, Activation parameters
[62] M. Chen, C. Zhang, X. Li, L. Zhang, Y. Ma, L. Zhang, X. Xu, F. Xia, W. Wang, for deformation of ultrafine-grained aluminium as determined by indenta-
J. Gao, A one-step method for reduction and self-assembling of graphene tion strain rate jumps at elevated temperature, Mater. Sci. Eng. A 585 (2013)
oxide into reduced graphene oxide aerogels, J. Mater. Chem. A 1 (8) (2013) 108e113.
2869e2877. [84] E.G. Herbert, W.C. Oliver, G.M. Pharr, Nanoindentation and the dynamic
[63] N. Shukla, C. Liu, P.M. Jones, D. Weller, FTIR study of surfactant bonding to characterization of viscoelastic solids, J. Phys. D Appl. Phys. 41 (7) (2008)
FePt nanoparticles, J. Magnetism Magnetic Mater. 266 (1e2) (2003) 074021.
178e184. [85] M. Bhattacharya, Polymer nanocompositesda comparison between carbon
[64] J. Coates, Interpretation of Infrared Spectra, a Practical Approach, Encyclo- nanotubes, graphene, and clay as nanofillers, Materials 9 (4) (2016) 262.
pedia of Analytical Chemistry, John Wiley & Sons, Ltd, 2006. [86] J. Du, H.-M. Cheng, The fabrication, properties, and uses of graphene/polymer
[65] G. Hollis, D.R. Davies, T.M. Johnson, L.G. Wade, Organic Chemistry, in: composites, Macromol. Chem. Phys. 213 (10e11) (2012) 1060e1077.
L.G. Wade Jr (Ed.), Test Item File, sixth ed., Pearson Prentice Hall, Upper [87] G. Feng, W.D. Nix, Indentation size effect in MgO, Scr. Mater. 51 (6) (2004)
Saddle River, N.J, 2006. 599e603.
[66] G.M. Miha ly Freund, Paraffin Products: Properties, Technologies, Applica- [88] N. Bheekhun, A.R. Abu Talib, M.R. Hassan, Aerogels in aerospace: an over-
tions, 1982. view, Adv. Mater. Sci. Eng. 2013 (2013) 18.
[67] R. Yogamalar, R. Srinivasan, A. Vinu, K. Ariga, A.C. Bose, X-ray peak broad- [89] K.L. Johnson, Contact Mechanics, Cambridge University Press, 1985.
ening analysis in ZnO nanoparticles, Solid State Commun. 149 (43e44) [90] Z. Han, Z. Tang, P. Li, G. Yang, Q. Zheng, J. Yang, Ammonia solution
(2009) 1919e1923. strengthened three-dimensional macro-porous graphene aerogel, Nanoscale
[68] Y.H. Zhao, X.Z. Liao, Z. Jin, R.Z. Valiev, Y.T. Zhu, Microstructures and me- 5 (12) (2013) 5462e5467.
chanical properties of ultrafine grained 7075 Al alloy processed by ECAP and [91] Z. Sui, Q. Meng, X. Zhang, R. Ma, B. Cao, Green synthesis of carbon nanotube-
their evolutions during annealing, Acta Mater. 52 (15) (2004) 4589e4599. graphene hybrid aerogels and their use as versatile agents for water purifi-
[69] T. Ungar, A. Borbely, The effect of dislocation contrast on x-ray line broad- cation, J. Mater. Chem. 22 (18) (2012) 8767e8771.
ening: a new approach to line profile analysis, Appl. Phys. Lett. 69 (21) [92] K.O. Kese, Z.C. Li, B. Bergman, Method to account for true contact area in
(1996) 3173e3175. soda-lime glass during nanoindentation with the Berkovich tip, Mater. Sci.
[70] T.K. Gupta, B.P. Singh, R.K. Tripathi, S.R. Dhakate, V.N. Singh, O.S. Panwar, Eng. A 404 (1e2) (2005) 1e8.
R.B. Mathur, Superior nano-mechanical properties of reduced graphene ox- [93] J.D. Gale, A. Achuthan, The effect of work-hardening and pile-up on nano-
ide reinforced polyurethane composites, RSC Adv. 5 (22) (2015) indentation measurements, J. Mater. Sci. 49 (14) (2014) 5066e5075.
16921e16930. [94] G. Feng, A.H.W. Ngan, Creep and strain burst in indium and aluminium
[71] H.J. Qi, K. Joyce, M.C. Boyce, Durometer hardness and the stress-strain during nanoindentation, Scr. Mater. 45 (8) (2001) 971e976.
behavior of elastomeric materials, Rubber Chem. Technol. 76 (2) (2003) [95] W.H. Poisl, W.C. Oliver, B.D. Fabes, The relationship between indentation and
419e435. uniaxial creep in amorphous selenium, J. Mater. Res. 10 (08) (1995)
[72] G. Feng, S. Qu, Y. Huang, W.D. Nix, An analytical expression for the stress 2024e2032.
field around an elastoplastic indentation/contact, Acta Mater. 55 (9) (2007) [96] P. Zhang, S.X. Li, Z.F. Zhang, General relationship between strength and
2929e2938. hardness, Mater. Sci. Eng. A 529 (2011) 62e73.
[73] G. Feng, S. Qu, Y. Huang, W.D. Nix, A quantitative analysis for the stress field [97] D. Tabor, The Hardness of Metals, Oxford University Press, 2000.
around an elastoplastic indentation/contact, J. Mater. Res. 24 (3) (2009) [98] W.D. Nix, H. Gao, Indentation size effects in crystalline materials: a law for
704e718. strain gradient plasticity, J. Mech. Phys. Solids 46 (3) (1998) 411e425.
[74] G. Feng, The Application of Contact Mechanics in the Study of Nano- [99] G.M. Pharr, E.G. Herbert, Y. Gao, The indentation size effect: a critical ex-
indentation, 2006. amination of experimental observations and mechanistic interpretations,
[75] K. Vanstreels, C. Wu, M. Gonzalez, D. Schneider, D. Gidley, P. Verdonck, Annu. Rev. Mater. Res. 40 (1) (2010) 271e292.
M.R. Baklanov, Effect of pore structure of nanometer scale porous films on [100] B.N. Lucas, W.C. Oliver, Indentation power-law creep of high-purity indium,
the measured elastic modulus, Langmuir 29 (38) (2013) 12025e12035. Metallurgical Mater. Trans. A 30(3) 601e610.
s, C. Yacou, M. Verdier, R. Dendievel, A. Ayral, Mechanical properties
[76] D. Jauffre [101] V. Maier, K. Durst, J. Mueller, B. Backes, H.W. Ho € ppel, M. Go€ken, Nano-
of hierarchical porous silica thin films: experimental characterization by indentation strain-rate jump tests for determining the local strain-rate
nanoindentation and Finite Element modeling, Microporous Mesoporous sensitivity in nanocrystalline Ni and ultrafine-grained Al, J. Mater. Res. 26
Mater. 140 (1e3) (2011) 120e129. (11) (2011) 1421e1430.
[77] B. Barabadi, R. Nathan, K.-p. Jen, Q. Wu, On the characterization of lifting [102] A. Pakzad, J. Simonsen, R.S. Yassar, Elastic properties of thin poly(vinyl
forces during the rapid compaction of deformable porous media, J. Heat alcohol)ecellulose nanocrystal membranes, Nanotechnology 23 (8) (2012)
Transf. 131 (10) (2009) 101006. 085706.
[78] R. Crawford, R. Nathan, L. Wang, Q. Wu, Experimental study on the lift [103] N.X. Randall, Direct measurement of residual contact area and volume
generation inside a random synthetic porous layer under rapid compaction, during the nanoindentation of coated materials as an alternative method of
Exp. Therm. Fluid Sci. 36 (2012) 205e216. calculating hardness, Philos. Mag. A 82 (10) (2002) 1883e1892.
[79] W.C. Oliver, G.M. Pharr, Measurement of hardness and elastic modulus by [104] L. Charleux, V. Keryvin, M. Nivard, J.P. Guin, J.C. Sanglebœuf, Y. Yokoyama,
instrumented indentation: advances in understanding and refinements to A method for measuring the contact area in instrumented indentation
methodology, J. Mater. Res. 19 (01) (2004) 3e20. testing by tip scanning probe microscopy imaging, Acta Mater. 70 (2014)
[80] X. Li, B. Bhushan, A review of nanoindentation continuous stiffness 249e258.

You might also like