Sie sind auf Seite 1von 10

Carbon 100 (2016) 546e555

Contents lists available at ScienceDirect

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

Absorbate-induced ordering and bilayer formation in propanolgraphite-oxide intercalates


Carlos Cabrillo a, *, Fabienne Barroso-Bujans b, c, Ricardo Fernandez-Perea a,
Felix Fernandez-Alonso d, e, Daniel Bowron d, Francisco Javier Bermejo a
a

Instituto de Estructura de la Materia (CSIC), Serrano 123, 28006 Madrid, Spain


bal 5, 20018 San Sebastian, Spain
Centro de Fsica de Materiales (CSIC-UPV/EHU)-MPC, Paseo Manuel Lardiza
bal 4, 20018 San Sebastian, Spain
Donostia International Physics Center (DIPC), Paseo Manuel Lardiza
d
ISIS Facility, Rutherford Appleton Laboratory, Chilton, Didcot, Oxfordshire, OX11 0QX, United Kingdom
e
Department of Physics and Astronomy, University College London, Gower Street, London, WC1E 6BT, United Kingdom
b
c

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 16 July 2015
Received in revised form
13 January 2016
Accepted 16 January 2016
Available online 18 January 2016

Real-time intercalation of liquid 1-propanol within graphite oxide prepared by Brodie's method has been
monitored by means of state-of-the-art neutron diffraction. We have developed a data analysis based on
the paracrystalline theory which enables the characterization of stacking disorder in terms of the average
number of layers contributing to Bragg scattering and the relative dispersion of the distribution of
interlayer distances. We nd that monolayer intercalation expands the interlayer spacing from 5.63 up
to 9.24 , a process which is accompanied by noticeable improvement of stacking order. Thermally
activated mobility of the molecular adsorbate is observed down to z130 K, followed by the emergence of
a two-dimensional glassy state below such temperature. Heating at a slow rate from this kinetically
arrested state, leads to an adsorbate-driven ordering transition at z180 K, followed by formation of a
stable structure of the propanol-graphite-oxide system. This new phase enables reversible bilayer formation and also is found to persist upon removal of excess propanol in vacuo. The nal propanolgraphite-oxide composite is characterized by an interlayer distance of 8.95 , a much-improved
ordering relative to its graphite-oxide precursor, and a pillar density of one intercalant molecule for
every nine graphene-like carbon hexagons.
2016 Elsevier Ltd. All rights reserved.

1. Introduction
The interest in Graphite Oxide (GO), a well known material in
graphene synthesis, has been reawakened in the past few years
mostly due to its potential for taylor-made design of materials for
separation technologies [1,2], hydrogen storage [3,4], gas capture
[5], or two-dimensional (2D) connement studies [6e10]. In
particular, pillaring of GO by intercalation of different molecules
provides a route to prepare porous materials with tunable pore size
[3,4].
Although the details of the structure of GO remain a matter of
some debate, the widely accepted model due to Lerf [11] portrays it
as consisting of oxidized sheets with the Oxygen-containing groups
located either at the surface of the basal planes (hydroxyl and

* Corresponding author.
E-mail address: ccabrilo@foton0.iem.csic.es (C. Cabrillo).
http://dx.doi.org/10.1016/j.carbon.2016.01.062
0008-6223/ 2016 Elsevier Ltd. All rights reserved.

epoxy) or at its edges (carbonyl and carboxyl). Such chemical


functional groups provide a source of chemical (bonding) interactions, i.e., guest molecules react with these functional groups
in an irreversible (chemisorptive) fashion. Some examples of this
behaviour include the intercalation of boronic acids [4], alkylsilane
coupling agents [12], and organic isocyanates [13].
Guest-host interactions driving the intercalation in GO can also
involve hydrogen bonding (HB) or dispersion interactions as shown
by the pioneering works of MacEwan and collaborators [14]. Since
both, dispersion and HB interactions are relatively weak, molecular
uptake leads to reversible intercalation of guest species, although
the retention of signicant amounts of molecules affecting the
thermal stability of the intercalate has been reported [15,16]. The
results are not only dependent upon the chemical nature of the
intercalant species but also on the method chosen to oxidize the
graphite precursor usually via Hummers' [17] or Brodie's methods
[18]. In particular, GO obtained by Hummers' method (H-GO) retains guest species such as ammonia [19], water [10], or alcohols

C. Cabrillo et al. / Carbon 100 (2016) 546e555

[20] in a different way than GO prepared by Brodie's method (BGO). Such differences seem to be attributable to either the presence
of sulphonic groups [19] in H-GO which are absent in B-GO, a
higher degree of oxidation in the former (eventually leading to
exfoliation in liquid media [21]) or higher amounts of carbonyl
(hydroxyl) groups in H-GO compared to B-GO [22]. The corresponding adsorption phenomenology is surprisingly diverse and, in
many cases, far from trivial. Already in the rst systematic studies
in the sixties [23e25], two different interlamellar phases depending on the molecular size of the intercalant were found, a so called
a phase where molecules are preferentially oriented parallel to
the GO layers, and a b phase with perpendicular preferential
orientation in the case of relatively large molecules. Furthermore,
alterations in the liquid/solid transition induced by the 2D
connement in these lamellar lms were also observed [26]. More
recently, it was found that B-GO if immersed in methanol or
ethanol undergoes under compression a remarkable reversible
transition between monolayer and bilayer intercalation. Consistently, both alcohols within B-GO show the same kind of stepwise
monolayer/bilayer transition under cooling while, in stark contrast,
they display huge continuous expansion during cooling when HGO is the host [20]. It seems nowadays established that the stepwise monolayer/bilayer transitions just referred to are characteristic of Brodie's GO hosts and, apart from the mentioned alcohols,
have been observed for solvents as diverse as methanol/water
mixtures (where also selective methanol intercalation is observed)
[27], acetone, dimethylformamide [28], or acetonitrile [29]. This
last example is particularly illustrative since the acetonitrile/H-GO
system does not show any monolayer/bilayer transition but rather
shows reentrancy between a usual layered intercalate and a
disordered gel-like phase within the temperature range 260e200 K
[29].
Within the mono-hydric alcohols, 1-propanol (1-Pr) is the
largest that does not show a b phase [23,24]. Alongside with
methanol and ethanol it intercalates B-GO from both the liquid [30]
and vapour phase [31] leading to a monolayer a-phase with a rather
narrow main diffraction peak and a well developed second Bragg
reection. However, to the best of our knowledge no monolayer/
bilayer transition has been reported on cooling 1-Pr intercalated
within B-GO. Here we report on some results obtained using stateof-the-art pulsed neutron diffraction to explore the kinetics and
temperature dependence of the intercalation and deintercalation of
liquid 1-Pr in Brodie's GO. A detailed analysis of the main and
second-order Bragg reections has enabled a quantitative assessment of the intercalation and deintercalation processes. The results
reveal a rich phenomenology and in particular, no monolayer/
bilayer transition was observed on cooling the intercalate from
room temperature. Instead it was found an ordering transition of
the system during slow heating from the cold (arrested) monolayer
state that allowed for the formation of a bilayer structure at room
temperature. Under removal of the excess propanol in vacuo the
bilayer is easily deintercalated but a monolayer propanol-graphiteoxide composite survives whilst retaining its structural integrity.
2. Experimental
2.1. Materials and GO synthesis
Anhydrous 1-Pr (99.7%) was purchased from Sigma-Aldrich and
used as received. The GO was prepared using natural graphite and
oxidized using a Brodie-based method [32]. In brief, a reaction ask
containing 200 mL of fuming nitric acid (Fluka) was cooled to 0  C
for 20 min using a cryostat bath, followed by the immersion of 10 g
of graphite. Next, 80 g of potassium chlorate (Fluka) was slowly
added over a period of 1 h in order to avoid sudden increases in

547

temperature. The reaction mixture was stirred for 21 h at 0  C. The


mixture was then diluted in distilled water and ltered until the
supernatant had a nitrate content below 1 mg/L (AQUANAL-plus
nitrate, NO3 1e50 mg/L). The resulting GO slurry was dried at 80  C
for 24 h in a vacuum oven (P < 0.1 mbar) and stored in this oven at
room temperature until further use.
Elemental analysis of the GO obtained in this manner obeyed a
stoichiometry C8 H1.05 O2.5. X-ray photoelectron spectroscopy
(XPS), 13C magic-angle-spinning (MAS) NMR data, X-ray diffraction
and scanning electron microscopy (SEM) images of graphite precursor and pristine GO are reported in the Supplementary
Information (SI).
2.2. Pulsed neutron diffraction
Neutron diffraction experiments were performed on the
NIMROD diffractometer at the ISIS facility, Rutherford Appleton
Laboratory, UK. NIMROD is a unique time-of-ight (TOF) neutron
diffractometer [33]. By virtue of its design using pulsed spallation
neutrons, it spans a very wide range of momentum transfers
1
(0:02  Q  100 
A ) with an exceptionally high counting rate.
Such a wide Q range is ideally suited to explore mesoscopic length
scales involved in molecular intercalation processes (~1 nm).
The sample holder was a at bespoke cell of linear internal dimensions 56 mm height, 35 mm width, and 4 mm depth with 1 mm
wall thickness, made of a TiZr null-scattering alloy. Such an alloy
scatters neutrons only incoherently so that do not add Bragg peaks
to the measured signal. The cell was connected to a centerstick
adapted for use with a 4-K closed-cycle refrigerator (CCR) and to a
gas-handling manifold by use of stainless-steel tubing (4 mm inner
diameter). The pressure within the system was monitored with an
MKS baratron. Calibrated thermocouples (RhFe) and heaters were
placed at the top and bottom of the sample cell and along the
stainless-steel tubing to ensure homogeneous heating of the
specimen and to avoid clogging of the circuit connecting the cell to
the gas manifold. This experimental setup allowed for ne temperature control (<0.1 K) over the range 4e400 K.
In addition to neutron-diffraction data on pristine GO and 1-Pr
intercalates, requisite ancillary measurements included those of
the background (empty cryostat), empty cell in the cryostat, and of
a vanadium standard. These data are necessary to perform an absolute calibration of scattered intensities as well as to subtract
contributions not arising from the sample under investigation.
These procedures were implemented with the Gudrun software
package [34]. In addition, Gudrun performs the required transforms
of the raw TOF data to Q-space and makes all the pertinent corrections including self-shielding, absorption, and multiple scattering. For a given stoichiometry and sample mass, Gudrun
computes the Differential Cross Section (DCS) of a given sample in
barns/sr per atom as well as the interference term, hereafter
denoted I(Q). In I(Q), the self-scattering from single nuclei is eliminated after all pertinent data corrections. I(Q) carries all the
structural information encoded in the DCS. Integration of the DCS
over all spatial directions gives the total cross-section (TCS), an
observable that in this work has also been simultaneously and
independently measured using a dedicated transmission detector.
The measured TCS is used in various corrections since recoil effects
render calculated TCSs unreliable if one uses tabulated scattering
lengths, in particular for hydrogen and its isotopes (see Ref. [34]
and also SI). These effects are usually known as inelasticity corrections and, in TOF neutron diffractometers like NIMROD, they
also affect the signal at low Q's. Clever schemes have been developed to deal with such a problem in hydrogenous samples [35]
where a perturbative approach does not work. These effects only
affect the self-scattering at low-Qs, and are of little consequence to

548

C. Cabrillo et al. / Carbon 100 (2016) 546e555

the well-dened Bragg features at higher Qs of relevance to this


work. The TCS can also be used to obtain an absolute measure of the
total number of irradiated nuclei. We have, therefore, used this
observable to follow in real time the evolution of 1-Pr content
during the experiment (see SI).
2.3. Data analysis methodology
The most relevant features in the I(Q) data correspond to the
main (0 0 1) Bragg peak as well as its second reection (0 0 2), both
of which provide direct information on how successive GO layers
are stacked. As such, they convey the most relevant information to
monitor intercalation processes. To model the peaks, we have used
a split-Voigt line shape in the spirit of the Thompson-Cox-Hastings
pseudo-Voigt approximation [36], namely,


8 
 < pV Q ; A; GG< ; GL< ; Q0

<
<
>
>


p Q ; A; GG ; GL ; GL ; GL ; Q0
: pV Q ; A; G > ; G > ; Q
0
L
G

Q  Q0
Q > Q0 ;
(1)

where Q0 is the mode (the location of the maximum) and GG< (GG> )
and GL< (GL> ) are Gaussian and Lorentzian full-width-at-halfmaxima (FWHM) determining the Voigt prole below (above) the
mode. The pseudo-Voigt prole (pV) is here dened as,

2
6
pVQ ; A; GG ; GL ; Q0 A41  hexp

Q  Q0
 ln2
G

2 !

pristine GO at room temperature. This corresponds to z 20 wt. %


excess with respect to maximal 1-Pr adsorption reported in previous studies (37 wt. %, see Ref. [31]). The mixture was left to
equilibrate for two hours, placed in the cryostat, and cooled rapidly
below 250 K. The intercalation process was then monitored by
means of diffraction runs of 5 mA h (ca 10 min). Following this step,
the temperature was raised to 304 K in three consecutive steps
(260 K, 280 K, and 305 K), aiming to impregnate and eventually
saturate the GO with the intercalate. This process was considered
complete when the diffraction pattern did not show pristine GO
features and did not change with time. Next, the sample was cooled
down to 100 K and brought back slowly to 300 K in a stepwise
manner (20 K increments at a data-collection rate of 25 mA h per
run).
In the third (and nal) step, we monitored molecular deintercalation, eventually leading to the formation in situ of a nal
dry nanostructure whose thermal response was subsequently
explored in a second thermal cycle. The rst stage of intercalate
evacuation was carried out by gently heating the sample to 305 K
and by connecting the cell to an expansion volume while
measuring diffraction runs of 5 mA h. After four hours of negligible
changes to the diffraction data, a harsher procedure was implemented by applying a continuous vacuum to the sample while
heating to 320 K, and nally to 350 K. After a total of 21 h under
these conditions, the system reached a stationary state where no
noticeable changes were observed in the neutron data. Evacuation
was stopped and the resulting dry intercalate was then subjected
to the same thermal cycle as before, that is, cooling down to 100 K
followed by stepwise heating to 300 K.


1

h
Q Q0
G

3. Results and discussion

7
2 5 ;

3.1. Pristine GO diffraction

(2)
where h and G are determined from GG and GL following the
formulae in Ref. [36]. Fits using this function plus a linear background enable a phenomenological parameterization of the
experimental data in terms of the intensity, mode, and width of
sharp I(Q) features. A limitation of this approach comes from the
admittedly ad-hoc modelling of the smoothly varying background.
If a given diffraction peak does not stand out sufciently above the
underlying background, an accurate background shape is needed
for a reliable estimation of the parameters.

Neutron-diffraction data for pristine GO are displayed in Fig. 1.


The DCS shows, as expected, a progression of relatively sharp reections riding on a strong background corresponding to selfscattering processes dominated by incoherent scattering from
hydrogen. The strong rise in intensity at low-Qs signals the presence of heterogeneous structure at meso and macro length scales.
From the (1 0 0) reection onwards, the I(Q) resembles that of other
graphitic materials (see for instance, [37]). As shown in the gure, it
is possible to assign all main features with the appropriate Miller
indices for graphitic reections. From these data, we nd no signatures associated with (0 0 2) or (0 0 4) Bragg reections from the

2.4. Experimental procedure


Prior to the neutron experiments, the GO sample was kept under vacuum at 353 K for 24 h. The subsequent neutron diffraction
measurements may be subdivided into three sequential steps.
In the rst step, diffraction data were collected on pristine GO.
An amount of 2.7 g was transferred to the sample cell described
above using a glove bag under an argon atmosphere. The cell was
connected to the centerstick and to the gas-handling manifold,
followed by evacuation. The centerstick length was adjusted so that
a 30  30 mm2 neutron beam illuminated the center of the sample
cell during the measurements. Such a beam conguration was used
all throughout the neutron experiments reported here. Diffraction
data were collected at 300 K for a total proton current of 136 mA h,
typically in runs of 20 mA h (ca 40 min).
In the second step, 1-Pr intercalation was monitored in real time,
followed by a study of the thermal behaviour of the intercalate
through a cooling-heating thermal cycle. To this end, 1.5 ml of liquid
1-Pr was gently injected directly into the sample cell containing

Fig. 1. I(Q) for pristine GO at 300 K. The red solid lines correspond to data ts using Eq.
(1) plus a linear background. The inset shows the measured DCS. (A colour version of
this gure can be viewed online.)

C. Cabrillo et al. / Carbon 100 (2016) 546e555

p DQ
1
h p2 g 2 ;

2 Qh1 h N

(3)

3.2. 1-Pr intercalation


Fig. 2 displays a general picture of this central part of the
experiment. It shows the 1-Pr load (wt. % relative to GO) irradiated
by the neutron beam obtained from the TCS data, alongside its
corresponding temperature evolution. The vertical dashed line
marks the end of the loading process and the beginning of the
cooling-heating cycle.
Let us rst consider the initial loading of 1-Pr, i.e., the region in
Fig. 2 up to the vertical dashed line. Fig. 3 illustrates the evolution of
the neutron response during the process. Inspection of these data
shows that after a two-hour equilibration period (green diffractogram in the gure), intercalation is far from complete as witnessed
by the presence of a substantial diffraction (0 0 1) peak corresponding to pristine GO. Over six hours were needed for the total
disappearance of GO diffraction features and the emergence of
stationary peaks at ~0.69 1 and ~1.38 1, characteristic of an
expanded intercalated (monolayer) composite structure.
A quantitative analysis of the diffraction patterns along the
whole second part of the experiement, i.e., before de-intercalation
is reported in Fig. 4. This gure shows the evolution of the intensity,
mode, and relative width for both (0 0 1) and (0 0 2) reections.
Peak-integrated intensities are shown in the uppermost frames in
blue. Alongside these data, we also report unnormalized intensities
(i.e., not divided by the number of scattering centres) as this
quantity contains information on the evolution of the 1-Pr load
within the irradiated volume. The middle frames show peak modes
which carry information on interplane spacings given by
d 2p=Q0 . The two lowermost frames in Fig. 4 show peak widths
expressed as % FWHM (100 D Q/Q0), providing a measure of sample
ordering. Again, the vertical dashed line marks the end of the initial
1-Pr load stage as in Fig. 2.
Fig. 4 evidences a noticeable change of regime after z 2.5 h
(T z 300 K). This behaviour is particularly noticeable for the (0 0 1)
intensities, showing a cusp between decreasing (earlier times) and
increasing intensities (later times). In contrast, the (0 0 2) reection
does not show the same initial decrease. This apparent discrepancy
arises from neutron-scattering contrast effects. The intercalation of
1-Pr molecules has diminished the contrast between GO layers,
thereby reducing the intensity of the rst Bragg reection. Since the
300

50
45

250

40
200

where N stands for the average number of Bragg planes (graphene


layers in our case) and g2 is the relative variance of the interplanar
2
distance d, i.e., g 2 d2 =d  1. The left-hand side of Eq. (3) accounts for the integral width of the corresponding Bragg reection.
The rst term on the right-hand side of Eq. (3) corresponds to size
effects described by the usual Scherrer formalism, leading to an
absolute width independent of reection order. The second term
comes from disorder, it is size independent, and depends distinctively on reection order. As expected, this is the case here with
G002 0:144 1 well above G001 0:076 1. Equation (3) provides a convenient quantitative framework to study disorder
directly from line broadening, as shown in more detail in the next
sections. Should the disorder be of the rst kind, we would need to
study the diffuse scattering between Bragg peaks.
Summarizing, our pristine GO exhibits an interlayer distance of
d 2p=Q0001 5:63 , with a well-dened and narrow (0 0 1) peak,
a weak but clearly discernible second-order reection, and no residual features from the graphite precursor. Previous neutron
studies of a Hummers' GO [42] provide the most direct comparison
of our results with existing literature. In this study, the interlayer
distance of GO (d 7:62 ) is substantially larger and the sample is
heterogeneous, as evidenced by the presence of (0 0 2) and (0 0 4)
reections from bulk graphite. In addition, the second GO reection
is not discernible and the data show a substantial amount of diffuse
scattering between peaks. As referred to above, the degree of
oxidation and composition of Hummers' GO is quite different to
those of Brodie's GO, thereby explaining the observed differences in
the neutron-diffraction data. When compared with XRD data of
Brodie's GOs of a higher degree of oxidation [43], our GO specimen
is still considerably more compact (the shortest d in Ref. [43] is
6.91 ). It is consistent, however, with the trend shown in Refs. [43],
where lower degrees of oxidation correlate with shorter d s. In
terms of the C/O ratio, the sample with d 6.91 corresponds to a
C/O 2.04 while in our case it amounts to C/O 3.2. Also, there is
no discernible second reection in any of the samples studied in
Ref. [43]. We are not aware of any diffraction data from GOs with a
discernible second reection of the main peak. On the basis of these

considerations, we conclude that our sample exhibits a low degree


of oxidation and it is remarkably homogeneous, i.e., absence of
graphite precursor, as well as a less expanded and more ordered
layered structure than more oxidized GOs.

wt. %

graphite precursor which should appear at ca. 1.85 1 and


3.70 1, respectively.
Quantitatively, the ts using Eq. (1) and shown in the gure as
red thick solid lines are consistent with a mode of the main
reection Q0001 1:116 1 and width G001 0:076 1 (FWHM),
as well as a mode of Q0002 2:240 1 and associated width G002
0:144 1 for the much-weaker (0 0 2) reection. These FWHMs are
above the instrumental resolution (~0.04 1 for Q0001 ), a result
which is in line with previous studies on this class of graphitic
layered materials [38]. Such an intrinsic width is indicative of the
paracrystalline character associated with the stacking of successive
graphene layers and it cannot be ascribed entirely to crystal-size
effects. For a proper crystal, the width is inversely proportional to
an effective grain size typically characterized by a neat number of
Bragg planes, N, as given by Scherrer's formula. Here, the weak
interaction between oxidized graphene layers leads to a stacking
disorder of the second kind [39], where deviations in the distance
between layer pairs are cumulative and scale as the square root of
the number of layers. The presence of such a built-in disorder has
been accounted for by models based upon the paracrystal concept
[40]. Using the theoretical framework developed by Hosemann and
coworkers, the relative width of an h-order peak associated with a
distribution of Bragg planes of the second kind can be expressed as
[41,40]

549

1-Pr Load

35

Temperature

30

150

25
0:00

4:00

8:00

12:00

16:00

20:00

100
24:00

Elapsed time (h:m)


Fig. 2. Time evolution of 1-Pr wt. % relative to GO (left axis) and temperature (right
axis) during intercalation and subsequent thermal cycle. Elapsed time is given in h:m:s
since the rst 10-min run following initial exposure to the neutron beam. (A colour
version of this gure can be viewed online.)

550

C. Cabrillo et al. / Carbon 100 (2016) 546e555

12

1-Pr@GO
(001)
40

35

1-Pr Load
Temperature
Elapsed time (h:m)

25
20
0:00

2:40:13, 304 K
d = 9.194

280

30

3:55:04, 304 K
d = 9.212

wt. %

6:16:13, 304 K
d = 9.238

I(Q) (arbitrary units)

300

45

10

2:00

240

6:00

Pristine GO
(001)

0:08:23, 244 K
d = 9.116

4:00

260

1-Pr@GO
(002)

Q (-1)

0.50

1.0

1.5

Fig. 3. Temporal evolution of neutron-diffraction patterns following 1-Pr loading. The I(Q) corresponding to the pristine GO is also shown as reference (dashed black line). For each
pattern the elapsed time as well as the corresponding interlayer distance are given. The inset shows the evolution of 1-Pr wt. % and temperature, i.e., a zoomed version of Fig. 2. The
arrows indicate the times corresponding to the neutron data shown in the main gure. (A colour version of this gure can be viewed online.)

(0 0 1) peak

0.50

300

0.12

0.40

250

0.10
Unnormalized
intensity
Temperature
Intensity

0.35
0.30
0.25

0.08

200

Arbitrary units

0.45

0.06
150

0.20

0.04

0.15

0.02

300

1.40

0.70

250

Mode
Temperature

0.69

1.39
200

-1

(0 0 2) peak

0.14

1.38
0.68

150

1.37
1.36

0.67
5.8

300

7.6

250

% FWHM

5.6
5.4

7.2

200

Temperature

6.8
150

5.2
6.4
5.0
0:00

4:00

8:00
12:00
16:00
Elapsed time (h:m)

20:00

24:00 0:00

4:00

8:00
12:00
16:00
Elapsed time (h:m)

20:00

100
24:00

Fig. 4. Quantitative analysis of the (0 0 1) and (0 0 2) reections during the rst stages of intercalation. The vertical dashed line marks the end of 1-Pr loading, while the dotted and
the solid gray lines correspond to regime changes as detailed in the text. For comparison, the gure on the right-hand side showing (0 0 2) % FWHMs also shows the (0 0 1) %
FWHMs corresponding to the residual GO (blue crosses). (A colour version of this gure can be viewed online.)

unnormalized intensity shows a at trend at these initial stages, it


follows that below some temperature Ta (300 K, 1-Pr molecules
can only enter previously intercalated GO galleries. Once the temperature increases above Ta, the intercalation of empty GO galleries
quickly overcomes the aforementioned contrast effects, as evidenced by a sharp rise in both intensities. The relatively rapid

intercalation above Ta also induces a spatial redistribution of the 1Pr liquid, as attested by the signicant increase in 1-Pr intensity
shown in Fig. 2.
In terms of peak widths (% FWHM), it is clear that 1-Pr intercalation greatly improves the ordering of the sample, as evidenced
by a clear decrease of (0 0 1) widths during loading. This ordering

C. Cabrillo et al. / Carbon 100 (2016) 546e555

also translates into an increasingly intense (0 0 2) reection. This


reection also displays a % FWHM comparable to that of (0 0 1) in
the residual GO (see blue crosses in the bottom right panel of Fig. 4).
To gain further insights, we can now use Eq. (3) on both (0 0 1)
and (0 0 2) reections to solve for N and g. For a proper application
of Eq. (3), the (non-negligible) instrumental resolution need to be
accounted for. Under the conditions of our experimental setup on
NIMROD, the relative FWHM resolution for both Bragg reections is
estimated to be a ~3.5% [33,34]. The line shape of the resolution
kernel is Gaussian-like with a linearly Q-dependent FWHM.
Otherwise, our peaks are narrow enough for the folding with the
kernel be very well approximated by a convolution with a Gaussian
of constant FWHM (and given by 3.5% of the corresponding Q0).
With these assumptions, and using the convolution properties of
the Voigt prole, incorporating
resolution effects boils down to
q
redening GG in Eq. (2) as 0:035 Q0 2 G2G .
The results of this procedure are shown Fig. 5. The procedure is
not precise enough for a fully edged quantitative assessment but it
sufces for a comparative analysis. Focussing on the initial loading
stage, N remains essentially constant below Ta while g diminishes
signicantly. This result is in line with the ability of 1-Pr to enter
only previously intercalated GO below Ta z 300 K, as discussed
earlier. During loading and above this temperature, we also observe
an increase in the number of effective Bragg planes and the associated variance in interplanar distance, as one would expect on the
basis of the discussion presented above.
3.2.1. Thermal cycle
Cooling down to 100 K and subsequent heating up to 300 K
display a fairly well-dened plateau in (0 0 1) width. The

100

300

80
250

40

200

60

150
N
Temperature

20

300
9.5
250

551

boundaries of this thermal cycle mark two regime changes, one


taking place during cooling at around Tg2D x130 K (dashed lines in
Figs. 4 and 5) and another during heating at (200 K (solid-gray line
in Figs. 4 and 5). Closer inspection of N and g provides a means of
examining their underlying nature.
During the rst stage of the thermal cycle (relatively rapid
cooling), the ensuing thermal contraction redistributes some of the
1-Pr out of the irradiated volume (see Fig. 2). Concomitantly, N
strongly decreases while g increases mildly. Altogether, these results signal a partial de-intercalation mainly driven by the differential thermal contraction of the composite system. This trend is
observed down to Tg2D x130 K where peak widths, N, and g all
reach stable values, signalling a transition to a kinetically arrested
state. Since molecular mobility freezes at Tg2D and there is no evidence of qualitative changes in molecular rearrangements, we may
assign a glassy nature to this transition. This quenching of the
disorder happens well above the glass transition of bulk 1-Pr at ~
98 K [44] and somewhat below the bulk melting temperature
(Tm 147 K). A further decrease in temperature does not affect
signicantly the quality of the paracrystal, only leading to an
additional thermal contraction of the material as witnessed by a
modest increase in the mode of the primary reection.
Consistent with a partial de-intercalation, during the subsequent slow heating, crossing Tg2D induces an increase in both N and
g, resembling the initial loading stage. However, at a Tt z 180 K, this
trend suddenly stops and N drops to the glassy value, followed by
an abrupt increase. Most notably, g at Tt reverses the trend and
starts decreasing. This behaviour indicates a structural rearrangement (a kind of annealing of the paracrystal) leading to an
improved packing which enables a further intercalation of 1-Pr
molecules (increasing N) in such a way that there is also an
improvement in layer order (decreasing g). As clearly shown in
Fig. 6, the new (thermally induced) ordering facilitates so much the
intercalation process that a second interspersed molecular layer
can now begin to intercalate, as attested by the development of a
new feature in the neutron-diffraction data centered at z 0.46 1
(cf. Fig. 6). In line with these considerations, the intensities of both
(0 0 1) and (0 0 2) reections decrease above Tt (see Fig. 4) in spite
of an increase in N (part of the diffraction signal is diverted to the
bilayer structure) while the higher microporosity induces an
appreciably higher 1-Pr content within the scattering volume (see
Fig. 2). Given unavoidable uncertainties in the homogeneity of
packing of a powder sample like our GO host, the observed levels of
1-Pr uptake (z wt. 48%) are entirely compatible with a uniform
distribution of the initial load (z wt. 44%). This result also indicates
that impregnation of GO by 1-Pr was quite uniform. Moreover, the
20% 1-Pr excess used in this process precludes the possibility of fullbilayer formation under our experimental conditions.

200

9.0
8.5
150

8.0
7.5
0:00

%g
Temperature
4:00

8:00
12:00
16:00
Elapsed time (h:m)

20:00

100
24:00

Fig. 5. Evolution of the average number of graphene layers, N, and spread in interlayer
distance, % g. Before the slow-heating stage, data have been binned in groups of three
consecutive runs. To better illustrate the inherent spread of these data, blue thin lines
show the corresponding raw data during the loading stage. N is represented in log
scale where data dispersion is approximately constant. As expected, % g shows larger
uncertainties when N is smaller. The second dotted line marks the crossing through
Tg2D on heating while the dashed-grey line denotes the transition temperature Tt (see
text). (A colour version of this gure can be viewed online.)

3.3. 1-Pr Evacuation


Fig. 7 provides a visual summary of the evacuation process. First,
we note the clear disappearance of bilayer signatures from the
neutron data. This happens after about one-and-a-half hours of
expansion in a large volume at 305 K. Once the bilayer has disappeared (blue diffractogram in Fig. 7), the main consequence of
pumping out while heating (stepwise increase up to 350 K) is a
strong drop in the signal intensity at low Qs. After 15 h of in vacuo
conditions, the low-Q signal remains constant, telling us that nonintercalated 1-Pr has been removed whilst preserving most of the
intercalated fraction. The inset on the right in Fig. 7 shows a
monotonic decrease in 1-Pr wt. %. For clarity, this gure shows
5 mA h data only at the beginning of the evacuation process and
during cooling to 100 K. In between, only one run out of ten is
displayed.

C. Cabrillo et al. / Carbon 100 (2016) 546e555

10:26:01, 100 K
d = 8.945
15:06:33, 201 K
d = 9.068
16:03:27, 221 K
d = 9.103
17:00:03, 240 K
d = 9.141
23:00:41, 300 K
d = 9.256

0.30

300

0.28
250

0.26
0.24

200

0.22

Intensity

0.20

Temperature

150

0.18
0.16
0:00

4:00

8:00
12:00
16:00
Elapsed time (h:m)

Bilayer
d =13.659

100
24:00

20:00

(002)

I(Q) (arbitrary units)

(001)

552

0
0.20

0.40

0.60

0.80

1.0

Q (-1)

1.2

1.4

Fig. 6. Temporal evolution of neutron-diffraction data during slow heating, as detailed in the text. Note the rapid appearance of a bilayer peak in these data. The inset shows peak
intensities as a function of time, where the coloured arrows indicate the location of the corresponding neutron data in the main gure. In these data, the intensity shows a clear
decrease above Tt. This transition temperature is marked by a vertical solid-gray line. (A colour version of this gure can be viewed online.)

Fig. 7. Evolution of the neutron-diffraction data during de-intercalation (note the Linear-Log scales). The inset on the left is a zoom-in of the main gure using a Linear-Linear
representation. The inset on the right corresponds to the evolution of 1-Pr wt. %. The origin of elapsed time coincides with the beginning of evacuation. Again the arrows mark
the runs corresponding to the shown I(Q)s. (A colour version of this gure can be viewed online.)

The results of our line-shape analysis are shown in Fig. 8,


showing the time evolution of N and g. These results show that
during the initial stages at 305 K, the de-intercalation process went
well beyond the second layer. During the subsequent drying
following the disappearance of bilayers, annealing induces a slow
recovery of the monolayer structure clearly evidenced in the evolution of N. Once the pumping is stopped and cooling initiated, N
undergoes a second transient where, after a rapid initial increase, it
relaxes to values similar to those just before the thermal cycle.
Meanwhile, g displays a mild evolution along the whole process.
Such a behaviour points to a partial re-intercalation, now from the
interstitial vapour phase, induced by the sudden loss of pumping
speed after sample isolation from the vacuum system, followed by a
relatively rapid cooling of the sample. These observations indicate a
stable 1-Pr/GO nanostructure that can withstanding quite harsh

drying conditions in vacuo. In other words, once the previous


thermal cycle has led to easily discernible molecular rearrangements within the GO galleries, the thermal stability of the resulting
intercalated compound undergoes a remarkable improvement.
Fig. 9 compares the nal intercalated structure with pristine GO.
The relevant parameters characterizing both structures are also
shown. From the above experiments, we have been able to ascertain that the nal composite system corresponds to a fully intercalated and dry specimen. Since the neutron data also enables us
to quantify the amount of 1-Pr irradiated by the beam, we can then
calculate the pillar density in this nanostructure. Given a
measured 20 wt. % of 1-Pr in the dry sample and a GO stoichiometry
of C8 H1.05 O2.5, we obtain a pillar density of 0.11 molecules per
graphene-like carbon hexagon, i.e., one 1-Pr molecule for every
nine hexagons.

C. Cabrillo et al. / Carbon 100 (2016) 546e555

100

350

80
300

60

250
K

40

200

20

150

N
Temperature

350
300

9.0

250

8.5

200

9.5

8.0
7.5
0:00

150

%g
Temperature
5:50

11:40
17:30
23:20
Elapsed time (h:m)

29:10

100
35:00

Fig. 8. Evolution of N and g during de-intercalation. (A colour version of this gure can
be viewed online.)

Fig. 9. Final dry intercalated structure compared to pristine GO. (A colour version of
this gure can be viewed online.)

3.4. Discussion
On chemical grounds, one expects that the intercalation of 1-Pr
in GO is a process primarily driven by HB interactions. Within the
interlayer spacing, such interactions are primarily mediated by
hydroxyl (-OH) and epoxy (-O-) groups in GO, whereas at the plane
edges one could expect that interactions with carboxylic-acid
groups (-COOH) dominate [11]. Therefore, the different nature of
molecular arrangements within the GO galleries and at the
entrance edges of the GO grains is not purely topological. The
analysis performed in terms of Eq. (3) during the loading process
reveals a crossover to an increasing trend in the number of intercalated layers (N) at Ta z 300 K. This result is suggestive of some
structural transition plausibly related to the breaking of weak
chemical bonds. Interactions between eCOOH groups at the edges
of adjacent GO layers could well lead to a kind of terminating

553

mechanism below Ta, a mechanism which would not be operative


for adjacent GO layers already expanded following the intercalation
of 1-Pr.
As expected for a less-expanded GO host, the interlayer distance
for maximal uptake observed in this work (9.24 ) remains below
those seen in other 1-Pr-intercalated GOs (z 9.50 , see
Refs. [14,30]). Our value is in fact comparable to those corresponding to ethanol intercalation in B-GO [30]. The expansion
induced by 1-Pr here is, however, substantially larger than those
reported previously (z 3.6 , as opposed to z 2.5 reported in
Refs. [14] and z 2.9 in Ref. [30]).
Somewhat counterintuitively, 1-Pr intercalation greatly improves the quality of the layer stacking, as attested by a strong
narrowing of the (0 0 1) reection and the emergence of a welldened (0 0 2) peak which is almost absent in pristine GO. The
observed (and rather substantial) temperature sensitivity of this
effect above Tg2D indicates that not only the 1-Pr molecules but the
also eOH and -O- groups are somewhat mobile. In this respect,
recent calculations [45] show that the GO structure is kinematically
constrained, that is, large energy barriers effectively prevent access
to thermodynamically stable structures during synthesis. In
particular, theoretical calculations indicate that certain eOH chain
structures are energetically very stable [46] but kinetically unreachable [45]. Additional mobility of the GO oxygen groups facilitated by the disruption caused by intercalated 1-Pr molecules
seems a plausible mechanism in a such metastable scenario.
Furthermore, the possibility of an ordering transition as the one
observed at Tt z 180 K after adequate annealing, emerges quite
naturally from the above considerations. According to the observations, intercalate reorganization leads to a thermally more stable
structure involving the participation of GO oxygen groups and 1-Pr
monolayer pillars. Such pillaring process allows for the reversible
intercalation of a second 1-Pr layer and, more importantly, it is
stable enough to support quite harsh drying conditions in vacuo.
The above phenomenology is in stark contrast with what have
been observed for methanol and ethanol using Brodie's and
Hummers' GO. In the later case, continuous GO expansion is followed by a sudden contraction upon cooling [20]. In the former
case, bilayer formation is observed at precise temperatures on
cooling [28]. Such an extremely rich intercalation phenomenology
lends further support to the kinematically constrained picture of
the various GO structures.
The nal dry structure is characterised by a pillar density of
one 1-Pr molecule every nine graphene-like carbon hexagons,
and an increase in interlayer distance by z3.3 . This increase is not
compatible with an upright arrangement of 1-Pr molecules (bphase). Since a fully extended horizontal propanol molecule can
spread across three GO hexagons, these gures indicate a loose
packing of pillars with some useable empty space between them.
In our case, the empty space between galleries is similar to the
pillaring obtained from the maximal uptake of methanol either
from the vapour [31] or by liquid immersion at room temperature
[27]. In this latter case, the packing is approximately of one methanol molecule for every three host hexagons.
The rearrangement observed after the transition at Tt provides
enough room for bilayer formation. From the point of view of potential applications of these novel materials, we note that the
possibility of stable pillaring arrangements is not restricted to the
case found in the present work. Other proportions of 1-Pr and/or
degrees of GO oxidation could result in similar types of dry stabilized congurations, ultimately leading to nanostructures with a
rather diverse and tunable degree of porosity or tortuosity.
The present results thus call for additional experiments where
the microscopic dynamics of the intercalate may be monitored by
means of, for example, high-resolution quasielastic neutron

554

C. Cabrillo et al. / Carbon 100 (2016) 546e555

scattering or dielectric spectroscopy. These studies will certainly


serve to corroborate the interpretation of the experimental results
we have so far obtained using neutron-diffraction techniques.

[9]
[10]

4. Conclusions and outlook

[11]

Using state-of-the-art real-time pulsed neutron diffraction we


have observed an unexpectedly rich phenomenology associated
with the intercalation of liquid 1-Pr in Brodie's GO. At Ta z 300 K, 1Pr intercalation leads to the formation of a monolayer structure
with a much-improved stacking order than that of the pristine GO
host and an interlayer distance of d 9:24 at maximal uptake.
The intercalated 1-Pr molecules remain mobile down to
Tg2D z 130 K. Heating slowly from below Tg2D, we observe an
ordering transition to a thermally stable structure at Tt z 180 K.
The stabilized intercalated material allows for the reversible
intercalation of a second 1-Pr layer. This new nanostructure is
stable enough to undergo a harsh drying process (heating in vacuo)
without signicant structural changes. The nal dry nanostructure is characterised by a far-improved stacking order relative
to pristine GO with an interlayer distance of d 8:95 and a pillar
density of one 1-Pr molecule every nine carbon hexagons.
From a methodological viewpoint, we have also provided a
relatively straightforward protocol to analyse stacking disorder in
nanostructured materials. Using the position and width of the rst
two Bragg reections, this procedure has provided important
physical insights into molecular intercalation at the nanoscale.

[12]

Acknowledgements

[24]

We thank J.A.R. Bones, B.C. Eltham, C.M. Goodway, and M.G.


Kibble from the ISIS Experimental Operations Division for their
expert assistance and superb technical support during the course of
the neutron experiments. This work was supported in part by Grant
n General de
No. MAT2012-33633 from the Spanish Direccio
 n Cientca y Te
cnica. The authors also gratefully
Investigacio
acknowledge support of Basque Government (IT-654-13), and the
UK Science & Technology Facilities Council for partial nancial
support and access to beamtime at the ISIS Facility.

[25]

[13]

[14]
[15]

[16]

[17]
[18]
[19]
[20]

[21]
[22]
[23]

[26]
[27]

[28]

[29]

Appendix A. Supplementary data


[30]

Supplementary data related to this article can be found at http://


dx.doi.org/10.1016/j.carbon.2016.01.062.

[31]

References

[32]

[1] R.K. Joshi, P. Carbone, F.C. Wang, V.G. Kravets, Y. Su, I.V. Grigorieva, et al.,
Precise and ultrafast molecular sieving through graphene oxide membranes,
Science 343 (6172) (2014) 752e754.
[2] R.R. Nair, H.A. Wu, P.N. Jayaram, I.V. Grigorieva, A.K. Geim, Unimpeded
permeation of water through helium-leaktight graphene-based membranes,
Science 335 (6067) (2012) 442e444.
 , D. Cazorla[3] Y. Matsuo, S. Ueda, K. Konishi, J.P. Marco-Lozar, D. Lozano-Castello
s, Pillared carbons consisting of silsesquioxane bridged graphene layers
Amoro
for hydrogen storage materials, Int. J. Hydrogen Energy 37 (14) (2012)
10702e10708.
[4] G. Srinivas, J.W. Burress, J. Ford, T. Yildirim, Porous graphene oxide frameworks: Synthesis and gas sorption properties, J. Mater Chem. 21 (2011)
11323e11329.
[5] M. Seredych, T.J. Bandosz, Removal of ammonia by graphite oxide via its
intercalation and reactive adsorption, Carbon 45 (10) (2007) 2130e2132.
[6] F. Barroso-Bujans, S. Cerveny, A. Alegria, J. Colmenero, Chain length effects on
the dynamics of poly(ethylene oxide) conned in graphite oxide: a broadband
dielectric spectroscopy study, Macromolecules 46 (19) (2013) 7932e7939.
[7] F. Barroso-Bujans, F. Fernandez-Alonso, S. Cerveny, S. Arrese-Igor, A. Alegria,
J. Colmenero, Two-dimensional subnanometer connement of ethylene glycol
and poly(ethylene oxide) by neutron spectroscopy: Molecular size effects,
Macromolecules 45 (7) (2012a) 3137e3144.
[8] F. Barroso-Bujans, F. Fernandez-Alonso, S. Cerveny, S.F. Parker, A. Alegra,

[33]

[34]

[35]
[36]
[37]

[38]
[39]
[40]
[41]

J. Colmenero, Polymers under extreme two-dimensional connement:


poly(ethylene oxide) in graphite oxide, Soft Matter 7 (2011a) 7173e7176.
S. Cerveny, F. Barroso-Bujans, A. Alegra, J. Colmenero, Dynamics of water
intercalated in graphite oxide, J. Phys. Chem. C 114 (6) (2010) 2604e2612.
A. Buchsteiner, A. Lerf, J. Pieper, Water dynamics in graphite oxide investigated with neutron scattering, J. Phys. Chem. B 110 (45) (2006) 22328e22338.
A. Lerf, H. He, M. Forster, J. Klinowski, Structure of graphite oxide revisited,
J. Phys. Chem. B 102 (23) (1998) 4477e4482.
Y. Matsuo, T. Tabata, T. Fukunaga, T. Fukutsuka, Y. Sugie, Preparation and
characterization of silylated graphite oxide, Carbon 43 (14) (2005)
2875e2882.
S. Stankovich, R.D. Piner, S.T. Nguyen, R.S. Ruoff, Synthesis and exfoliation of
isocyanate-treated graphene oxide nanoplatelets, Carbon 44 (15) (2006)
3342e3347.
J. Cano Ruiz, D.M.C. MacEwan, Interlamellar sorption complexes of graphitic
acid with organic substances, Nature 176 (1955) 1222e1223.
F. Barroso-Bujans, S. Cerveny, R. Verdejo, J. del Val, J. Alberdi, A. Alegra, et al.,
Permanent adsorption of organic solvents in graphite oxide and its effect on
the thermal exfoliation, Carbon 48 (4) (2010a) 1079e1087.
F. Barroso-Bujans, J.L.G. Fierro, A. Alegria, J. Colmenero, Revisiting the effects
of organic solvents on the thermal reduction of graphite oxide, Thermochim.
Acta 526 (1e2) (2011b) 65e71.
W. Hummers, R. Offeman, Preparation of graphitic oxide, J. Am. Chem. Soc. 80
(1958), 1339e1339.
B. Brodie, Sur le poids atomique du graphite, Ann. Chim. Phys. 59 (1860)
466e472.
C. Petit, M. Seredych, T.J. Bandosz, Revisiting the chemistry of graphite oxides
and its effect on ammonia adsorption, J. Mater Chem. 19 (2009) 9176e9185.
S. You, B. Sundqvist, A.V. Talyzin, Enormous lattice expansion of hummers
graphite oxide in alcohols at low temperatures, ACS Nano 7 (2) (2013a)
1395e1399.
n, Graphene oxide
J.I. Paredes, S. Villar-Rodil, A. Martnez-Alonso, J.M.D. Tasco
dispersions in organic solvents, Langmuir 24 (19) (2008) 10560e10564.
, A.V. Talyzin, Effect of synthesis method on solS. You, S.M. Luzan, T. Szabo
vation and exfoliation of graphite oxide, Carbon 52 (2013b) 171e180.
n, J. Cano Ruiz, D.M.C. MacEwan, Beta-type interlamellar sorption
F. Arago
complexes, Nature 183 (1959) 740e741.
A.R. Garca, J. Cano-Ruiz, D.M.C. MacEwan, Temperature variation of b-type
interlamellar sorption complexes of graphitic acid with alcohols, Nature 203
(1964) 1063e1064.
n, D.M.C. MacEwan, Sorption of organic molecules by graphitic acidF. Arago
and methylated graphitic acid: A preliminary study, Kolloid Zeitschrift Z. fur
Polym. 203 (1965) 36e42.
n, Phase transitions in interlamellar lms, Nature 184
D. MacEwan, F. Arago
(1959) 1859.
S. You, J. Yu, B. Sundqvist, L. Belyaeva, N.V. Avramenko, M.V. Korobov, et al.,
Selective intercalation of graphite oxide by methanol in water/methanol
mixtures, J. Phys. Chem. C 117 (4) (2013c) 1963e1968.
S. You, S. Luzan, J. Yu, B. Sundqvist, A.V. Talyzin, Phase transitions in graphite
oxide solvates at temperatures near ambient, J. Phys. Chem. Lett. 3 (7) (2012)
812e817.
A.V. Talyzin, A. Klechikov, M. Korobov, A.T. Rebrikova, N.V. Avramenko,
M.F. Gholami, et al., Delamination of graphite oxide in a liquid upon cooling,
Nanoscale 7 (29) (2015) 12625e12630.
 , I. De
k
A.V. Talyzin, B. Sundqvist, T. Szabo
any, V. Dmitriev, Pressure-induced
insertion of liquid alcohols into graphite oxide structure, J. Am. Chem. Soc. 131
(51) (2009) 18445e18449.
F. Barroso-Bujans, S. Cerveny, A. Alegra, J. Colmenero, Sorption and desorption behavior of water and organic solvents from graphite oxide, Carbon 48
(11) (2010b) 3277e3286.
F. Barroso-Bujans, F. Fernandez-Alonso, J.A. Pomposo, E. Enciso, J.L.G. Fierro,
J. Colmenero, Tunable uptake of poly(ethylene oxide) by graphite-oxide-based
materials, Carbon 50 (14) (2012b) 5232e5241.
D.T. Bowron, A.K. Soper, K. Jones, S. Ansell, S. Birch, J. Norris, et al., Nimrod:
The near and intermediate range order diffractometer of the isis second target
station, Rev. Sci. Instrum. 81 (033905) (2010) 1e9.
A.K. Soper, Gudrunn and Gudrunx: Programs for Correcting Raw Neutron and
X-ray Diffraction Data to Differential Scattering Cross Section, RAL Technical
Reports: RAL-TR-2011-013, Rutherford Appleton Laboratory, 2011. URL:
https://epubs.stfc.ac.uk/work/56240.
A.K. Soper, Inelasticity corrections for time-of-ight and xed wavelength
neutron diffraction experiments, Mol. Phys. 107 (16) (2009) 1667e1684.
P. Thompson, D. Cox, J. Hastings, Rietveld renement of debye-scherrer synchrotron x-ray data from al2o3, J. Appl. Crystallogr. 20 (2) (1987) 79e83.
J.C. Dore, M. Sliwinski, A. Burian, W. Howells, D. Cazorla, Structural studies of
activated carbons by pulsed neutron diffraction, J. Phys. Condens Matter 11
(47) (1999) 9189.
A. Szczygielska, A. Burian, J.C. Dore, Paracrystalline structure of activated
carbons, J. Phys. Condens Matter 13 (2001) 5545e5561.
A. Guinier, X-Ray diffraction, in: Crystals, Imperfect Crystals, and Amorphous
Bodies, 1 ed., Dover Publications, Inc, New York, 1994, p. 231 chap. 7.
A.M. Hindeleh, R. Hosemann, Paracrystals representing the physical state of
matter, J. Phys. C Solid State 21 (1988) 4155e4170.
W. Vogel, R. Hosemann, Evaluation of paracrystalline distortions from line
broadening, Acta Cryst. A 26 (1970) 272e277.

C. Cabrillo et al. / Carbon 100 (2016) 546e555


[42] J. Johnson, C. Benmore, S. Stankovich, R. Ruoff, A neutron diffraction study of
nano-crystalline graphite oxide, Carbon 47 (9) (2009) 2239e2243.
, O. Berkesi, P. Forgo
, K. Josepovits, Y. Sanakis, D. Petridis, et al.,
[43] T. Szabo
Evolution of surface functional groups in a series of progressively oxidized
graphite oxides, Chem. Mater 18 (11) (2006) 2740e2749.
[44] M. Ramos, C. Talon, R. Jimenez-Rioboo, S. Vieira, Low-temperature specic

555

heat of structural and orientational glasses of simple alcohols, J. Phys. Condens. Mat. 15 (2003) S1007eS1018.
[45] N. Lu, D. Yin, Z. Li, J. Yang, Structure of graphene oxide: Thermodynamics
versus kinetics, J. Phys. Chem. C 115 (24) (2011) 11991e11995.
[46] J.A. Yan, L. Xian, M. Chou, Structural and electronic properties of oxidized
graphene, Phys. Rev. Lett. 103 (8) (2009), 086802(1e4).

Das könnte Ihnen auch gefallen