Sie sind auf Seite 1von 10

Journal of Industrial and Engineering Chemistry 83 (2020) 449–458

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


journal homepage: www.elsevier.com/locate/jiec

Preparation of synthetic graphite from waste PET plastic


Seunghyun Koa , Yeon Ju Kwona,b , Jea Uk Leea,c , Young-Pyo Jeona,c,*
a
Carbon Industry Frontier Research Center, Korea Research Institute of Chemical Technology (KRICT), 141, Gajeong-ro, Yuseong-gu, Daejeon, 34114, Republic of
Korea
b
School of Chemical Engineering, Sungkyunkwan University (SKKU), 2066, Seobu-ro, Jangan-gu, Suwon, 16419, Republic of Korea
c
Advanced Materials and Chemical Engineering, University of Science and Technology (UST), 217, Gajeong-ro, Yuseong-gu, Daejeon, 34113, Republic of Korea

A R T I C L E I N F O A B S T R A C T

Article history: The development of new technologies for converting waste plastics into value-added products is
Received 26 June 2019 attracting widespread attention because of the global plastic waste crisis. In this study, we present a
Received in revised form 3 December 2019 novel route for the upcycling of waste plastics, wherein synthetic graphite is prepared from a waste
Accepted 16 December 2019
polyethylene-terephthalate (PET) bottle through a facile and scalable method. PET was successfully
Available online 23 December 2019
converted to graphite via a synthetic method of pyrolysis at 900  C followed by boron-assisted catalytic
graphitization at 2400  C. This technique overcame the intrinsic non-graphitizable property of PET and
Keywords:
yielded graphite showing high crystallinity with the maximum crystallite size of 20.9 nm in Lc and the
Plastic waste
Polyethylene terephthalate
d(002) spacing of 3.373 Å. In particular, it showed a much higher degree of graphitization (80.6%) than that
Graphitization (68.9%) derived from a well-known AR mesophase pitch (Mitsubishi). In addition, via a microwave-
Graphite assisted liquid-phase exfoliation, the PET-derived graphite was successfully exfoliated as graphene sheets
Graphene with the average lateral size of 410 nm. We expect that our work can guide the innovative upcycling of
waste plastics to invaluable synthetic graphite, which has many potential applications such as anode
materials in secondary batteries and fillers for carbon composites, and may serve as an alternative source
for graphene production.
© 2019 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
reserved.

Introduction Upcycling is any technique that transforms waste materials into


more valuable products [4]. Recently, many studies on upcycling of
Rapidly increasing plastic waste is becoming a major threat to waste PET into various valuable carbon materials, including
living organisms on earth. As of 2015, a cumulative total of 6.3 activated carbons (ACs), carbon spheres, carbon sheets, and
billion metric tons (MT) of plastic waste had been generated, but nanostructured carbons, have been reported. The main constituent
only 9% recycled [1]. Of the rest of the plastic waste, 79% has been of PET is carbon with 62.5 wt%; it is free of inorganic components
landfilled or discarded in the ocean while 12% has been incinerated [5], making it a good source for the preparation of carbon
[1,2], causing serious environmental contamination. Polyethylene materials. Many attempts for the conversion of PET into ACs have
terephthalate (PET) is one of the most commonly used plastics, been reported [6–9]. Mendoza-Carrasco et al. prepared ACs
mainly applied as containers for various liquids and foods because through physical and chemical activation methods [6]. Bratek
of its affordability, transparency, and good mechanical properties. et al. converted PET into AC through carbonization followed by CO2
The estimated amount of PET consumption has exceeded 24 activation; the AC showed hydrogen storage capability and
million MT per year and continues to increase [3]. This contributes applicability as a supercapacitor electrode [7]. The application of
a significant volume of plastic waste. Hence, new innovative PET-derived AC as a sensing platform was also suggested [8]. The
technologies for the post-utilization of PET and other plastic preparation of porous carbons other than ACs from PET, such as
wastes are in high demand to maximize their utility and avoid porous carbon sheets [10] and mesoporous carbons [11–13], has
landfilling or incineration. also attracted significant attention. Regarding nanocarbons, the
conversion of PET into graphene-like materials [5,14], nanotubes
[15,16], and other nanostructured carbons has been attempted.
Unfortunately, previous studies have been limited to the
* Corresponding author at: Carbon Industry Frontier Research Center, Korea
conversion of PET into carbon materials with disordered (amor-
Research Institute of Chemical Technology (KRICT), 141, Gajeong-ro, Yuseong-gu,
Daejeon, 34114, Republic of Korea. phous or turbostratic) structures or nanostructured carbons, rather
E-mail address: ypjeon@krict.re.kr (Y.-P. Jeon). than the highly ordered graphite structure. Graphite is of great

https://doi.org/10.1016/j.jiec.2019.12.018
1226-086X/© 2019 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.
450 S. Ko et al. / Journal of Industrial and Engineering Chemistry 83 (2020) 449–458

research interest and widely used industrially because of the Experimental


properties arising from its unique crystal structure. In the graphite
structure, sp2-hybridized carbon atoms are covalently bonded and Preparation of synthetic graphite from waste PET
arranged hexagonally in a planar condensed ring system that forms
two-dimensional (2D) layers. The layers are stacked parallel to A PET water bottle (Jejusamdasoo, Jejuprovince Development
each other and held together with loose van der Waals forces. Co., Republic of Korea) was used as a starting material. It was cut
Because of this structural feature, graphite is a good conductor of into pieces measuring 10 mm  10 mm and subsequently
heat and electricity and shows a high regular stiffness and carbonized and graphitized.
strength. It sustains these properties at temperatures reaching For carbonization, 10 g of the PET flakes were placed in a tube-
~3600  C with almost no thermal expansion. These special type furnace and heated to 900  C at a ramping rate of 5  C/min and
properties permit many uses in mechanical, nuclear, aerospace, held for 1 h. To create an inert atmosphere, N2 gas was flowed
and chemical applications. In particular, because of its high energy- through the furnace throughout heat treatment. The carbonaceous
storage capacity and stable cycle life, graphite is the most material obtained through this process showed the carbon yield of
commonly used anode material for rechargeable lithium-ion 19.0% and was named wPET-C9.
batteries [17] since the first commercialization of such batteries For catalytic graphitization, wPET-C9 was ground and sieved
in the early 1990s [18]. Research on PET conversion to graphitic into powder of <250 mm in particle size. Then, the wPET-C9
structures is limited because of the non-graphitizable nature of powder was mixed with amorphous boron powder (Sigma-
PET. The non-graphitizability suppresses the structural evolution Aldrich, 1 mm in average particle size) with various mixing ratios
to graphite during graphitization processes, which are usually of 0, 1, 2, 3, 4, and 5 wt%. The mixtures were graphitized in a box-
performed at temperatures above 1800  C under inert atmos- type furnace (ThermVac Engineering, Republic of Korea) equipped
pheres [19]. Through the conventional graphitization process, with graphite heating elements. Heat treatment was performed at
graphite cannot be obtained from PET. Therefore, the graphitiza- 2400  C for 1 h under flowing helium gas. The ramping rate was
tion of PET remains a significant challenge. 15  C/min. These graphitization processes produced six samples of
In this study, we attempted to convert PET waste to graphite. A the series wPET-Bx-G24 (0–5% boron), where x indicates the
PET bottle was used as the starting material and successfully amount of boron added. All the samples exhibited similar
converted to graphite with a well-developed crystal structure. The graphitization yields of 86.0  0.6% without distinct variations
synthetic method involved the pyrolysis of PET under an inert according to the boron content.
atmosphere and subsequent catalytic graphitization in the For direct graphitization, 1 g of the PET flakes was placed in the
presence of boron (Fig. 1). Moreover, we performed liquid-phase same box-type furnace described above under the same conditions
exfoliation on the resultant graphite and successfully obtained but without any additives. The thus-prepared product was named
graphene sheets. The intrinsic non-graphitizability of the PET was wPET-DG24 and showed the carbon yield of 17.1%.
overcome by the boron-induced catalytic graphitization. The
addition of boron to carbon fibers improves the mechanical and Preparation of graphene from PET-derived graphite
thermal properties of the carbon fibers, as has been reported
previously [20,21]. This is because of the catalytic effect of boron on Graphene was prepared from wPET-B5-G24 via microwave-
graphitization [19,22,23]. Boron-induced catalytic graphitization assisted liquid phase exfoliation. For expansion of graphite, 1 g of
has been used to prepare high-performance carbon materials from wPET-B5-G24 was placed in a 100 mL flask, and 40 mL of sulfuric acid
various carbon sources such as phenolic resin [24,25], polyimide (95%, Samchun Chemicals) and 12 mL of fuming nitric acid
[26], glucose [27], coke [28], and bituminous coal [29,30]. (95%, Samchun Chemicals) were separately added. After stirring
However, the catalytic effect of boron has not previously been for 24 h at ambient temperature, the mixture was poured slowly into
used for the graphitization of PET or other plastic wastes. To the deionized (DI) water and vacuum-filtered, yielding a solid sample.
best of our knowledge, we report the first use of boron for the This solid was washed using DI water several times and dried at 60  C
graphitization of plastic waste and the first successful conversion for 24 h. The acid-treated graphite was sealed in a glass vial and
of PET to graphite. purged with Ar for 2 h. The vial was then heated in a microwave oven

Fig. 1. Schematic for preparation of graphite and graphene from PET plastic.
S. Ko et al. / Journal of Industrial and Engineering Chemistry 83 (2020) 449–458 451

(MD-203CC, 60 Hz, LG Electronics) for 6 s. The expanded graphite residual matter shows progressive carbonization. The TGA curve
flakes were then suspended in N-methyl-2-pyrrolidone (NMP) shows that the decomposition produces 17.3 wt% residue, and the
(99.5%, Samchun Chemicals) with the concentration of 3.33 mg/mL. subsequent carbonization produces finally 13.4 wt% carbon. Other
The resulting suspension was then exfoliated in an ultrasonicator plastics comprising only carbon and hydrogen cannot undergo
(JAC-3010, U1TECH, Republic of Korea) for 3 h. The obtained carbonization; instead, they decompose completely and vaporize
dispersion was centrifuged for 5 min at 5000 rpm. The final without producing any carbonaceous matter. As examples, the
graphene powder was collected by vacuum filtration. TGA/DSC results of polystyrene (PS) and low-density polyethylene
(LDPE) are depicted in Fig. S1 in the Supplementary material. They
Material characterization showed complete decomposition under heating to 800  C and
produced almost 0 wt% residue. However, the presence of oxygen
X-ray diffraction (XRD) was conducted using a Rigaku D/Max-3C in the PET structure allows PET to withstand decomposition to a
diffractometer equipped with a rotating anode and a Cu Kα radiation certain extent and experience carbonization, because the oxygen
source (l = 1.5418 Å); the samples were scanned in the 2u range atoms in PET are present as ester groups serving as chemical
10–90 with a step size of 0.02 . Thermal gravimetric analysis (TGA) linkages.
and differential scanning calorimetry (DSC) were performed using a
SDT Q600 (TA Instruments) under N2 flow with a heating rate of Structural change of the waste PET upon carbonization and
5  C/min. Fourier transform infrared spectroscopy (FTIR) was graphitization
conducted with Bruker ALPHA-T. X-ray photoelectron spectroscopy
(XPS) data were recorded using an AXIS NOVA spectrometer (Kratos In order to identify the crystal structure developed after
Analytical Ltd., United Kingdom) with a monochromatic Al Kα carbonization, wPET-C9 (produced by heat-treatment of waste PET
source at a power of 150 W (15 KeV, 12 mA). Scanning electron at 900  C for 1 h) was analyzed using XRD. Fig. 3(a) presents the
microscopy (SEM) images were obtained using a CX-200 (COXEM). obtained XRD pattern of wPET-C9. It shows three broad peaks at
Atomic force microscopy (AFM) characterization was performed 22.7, 43.6 , and 79.2 , corresponding to the (002), (100), and (110)
using a Bruker Nanoscope in the tapping mode with a scan size of crystal planes of the graphite structure, respectively. All peaks are
2 mm and total scan time of 10 h. TEM images were obtained using low in intensity with large peak widths, consistent with typical
an F-200s model (FEI Talos). diffraction patterns of amorphous carbons [32,33]. This indicates
that heat-treatment of PET at 900  C produces carbon with
Results and discussions amorphous structure.
To investigate its graphitizability, PET was directly heat-treated
Crystallinity and thermal behavior of waste PET at 2400  C for 1 h. Fig. 3b shows an XRD pattern of the resulting
product, denoted wPET-DG24. No distinctive differences are
Firstly, the crystallinity of the waste PET used in this study was observed compared to the pattern of wPET-C9, except for slightly
examined by XRD. PET has a triclinic crystal structure; when higher intensities of the (002) and (100) peaks and the appearance
analyzed by XRD, it shows three characteristic peaks at 18 , 23.4 , of a very low-intensity (004) peak. This showed that graphite
and 25.8 in its XRD pattern, corresponding to the (010), (110), structure is not obtained from PET even through heat-treatment at
and (100) crystal planes, respectively [31]. Fig. 2(a) shows the temperatures beyond 1800  C. Normally, graphitizable carbons
XRD pattern of the waste PET sample used in this study. One large progressively undergo development of graphite structures at
peak appears at 25.6 , corresponding to the (100) plane. The temperatures above 1800  C. Mesophase pitch is a well-known
deconvolution of this peak, as shown in the inset, reveals two graphitizable carbon source capable of graphitization at such
peaks at 18.1 and 22.9 with low intensities. These peaks temperatures. Using mesophase pitch (AR pitch, Mitsubishi Gas
correspond to the (010) and (110) planes, respectively. This Chemical), synthetic graphite was prepared via heat-treatment at
demonstrates that the PET used in this study has a triclinic 2400  C for 1 h and its XRD pattern was depicted in Fig. 3(c). It is
structure with low crystallinity. seen that its diffraction pattern is quite different from that of wPET-
Fig. 2(b) shows the thermal behavior and stability of the waste DG24. It had a pattern with a high signal-to-noise, and is consistent
PET as measured by DSC/TGA. The DSC curve shows the heat flow with that of natural graphite (Kanto Chemical Co. Inc.) depicted in
upon heating at temperatures reaching 800  C. The endothermic Fig. 3(d), suggesting synthetic graphite had graphite structure.
peak at 250  C indicates the melting of PET. A subsequent However, in the case of wPET-DG24, the characteristic features in
endothermic peak shows that PET decomposition begins at diffraction pattern of graphite structure were not observed, and
360  C, with the peak maximum at 417  C. After decomposition, thereby confirming that it still remained to be amorphous

Fig. 2. (a) XRD pattern of PET; inset shows deconvolution of (100) peak. (b) TGA and DSC curves of PET.
452 S. Ko et al. / Journal of Industrial and Engineering Chemistry 83 (2020) 449–458

Fig. 3. XRD patterns of carbon products obtained after heat-treatment of PET at (a) 900  C (wPET-C9) and (b) 2400  C (wPET-DG24). XRD patterns of (c) synthetic graphite and
(d) natural graphite. The synthetic graphite is prepared by heat-treatment at 2400  C of mesophase pitch, a well-known graphitizable carbon precursor.

structure. This indicates graphite cannot be obtained from PET via graphitization can be expressed as the increase in size of the
conventional graphitization process. individual carbon layers and the increase in number of the carbon
The non-graphitizable property of PET likely arises from the layers. The consequences of graphitization are evident from XRD
presence of oxygen atoms in the PET structure. PET contains 33 wt analysis [39–41]. The most common feature is the (002) diffraction
% oxygen in its polymer structure. Heat treatment in the presence peak appearing at~26 , which is attributed to the stacking of carbon
of oxygen induces the oxidation of organic materials at temper- layers; the diffraction intensity strengthens as the number of layers
atures above 180  C [34]. This causes crosslinking within their increases. The growth of individual carbon layers along the in-
molecular structures between oxygen-containing groups, which plane direction is apparent through a change in the (110) peak at
fixes the molecular structures [35,36] and prevents the liquid- ~77. This peak is broader at the beginning and becomes sharper as
phase carbonization necessary for graphitization [37]. Many graphitization progresses. In addition, two peaks of (100) and (101)
organic species solidify and undergo carbonization without appear as a single peak at ~43 when the crystallinity is low; the
liquid-state carbonization when they are heated beyond a certain peaks gradually separate as the crystallinity increases. Other peaks
temperature; the subsequent dehydrogenation and condensation including (004) and (112) also appear as evidence of graphitization.
reactions yield polycyclic aromatic hydrocarbons (PAHs) and After the catalytic graphitization, these changes are distinc-
consequently carbonaceous materials. Some organic species tively observed in the XRD patterns of the PET-derived products.
undergo liquid-phase carbonization before solid-phase carboni- Fig. 4 shows the XRD patterns of products obtained after boron-
zation. During the liquid-phase carbonization, reactions similar to assisted catalytic graphitization (wPET-B0-G24 to wPET-B5-G24).
those occurring in solid-phase carbonization yield PAHs, but the The (002) reflection peak is dramatically increased in intensity and
liquid phase permits the molecular rearrangement of PAHs, decreased in width as the boron addition amount increases from
thereby producing highly oriented graphite-like structures. Such 0 wt% (wPET-B0-G24) to 5 wt% (wPET-B5-G24), suggesting
graphite-like structures are advantageous for crystal growth increases in the number of stacked graphene layers. In addition,
without significant PAH migration or rotation, thus enabling the insets of Fig. 4 show the gradual separation of the peak near 43
graphitization. Meanwhile, because PET skips liquid-phase into two different peaks of (100) and (101), as well as the
carbonization and directly begin solid-phase carbonization due appearance of a new peak at ~83.3 corresponding to the (112)
to the crosslinking by oxygen containing groups, highly oriented plane. The (110) peak at around 77.2 becomes very sharp and
structures are not formed and disordered structures are main- narrow, suggesting the growth of individual graphene layers along
tained even after graphitization treatment. the in-plane direction as well. All these changes in the XRD
patterns and the corresponding crystal growths indicate the
Graphitization of waste PET evolution of the graphite structure. Comparison with the diffrac-
tion patterns of synthetic graphite and natural graphite, depicted
Graphite has a structure of two-dimensional carbon planes at the bottom of the patterns (like those presented in Fig. 3),
composed of hexagonal rings; each plane is one graphene layer, confirms that the graphitization of wPET-B5-G24 has progressed
and graphite is an orderly stack of graphene layers along the well, with a similar crystallinity to both synthetic graphite and
direction perpendicular to the graphene plane [38]. Thus, natural graphite.
S. Ko et al. / Journal of Industrial and Engineering Chemistry 83 (2020) 449–458 453

Based on the model given by Maire and Mering [42,43], the


degree of graphitization is calculated using the following equation:
0:3440  dð002Þ
Degree of graphitization ð%Þ ¼  100 ð1Þ
0:3440   0:3354 
where 0.3440 and 0.3354 are the d-spacings of the fully non-
graphitized carbon and the ideal graphite crystal, respectively. This
formula defines the fully non-graphitized carbon as turbostratic
carbon with the d(002) spacing of 0.3440 nm; therefore, carbons
with much poorer crystallinities than turbostratic carbon can show
negative values in the degrees of graphitization. The calculated
degrees of graphitization of wPET-DG24, wPET-B5-G24, synthetic
graphite, and natural graphite are presented in Fig. 5(b). The
sample obtained by direct graphitization from PET (wPET-DG24)
exhibits the degree of graphitization of 42%. This is far less than
that of the fully non-graphitized structure defined by the formula,
indicating that the evolution of the graphite structure does not
progress at all during the direct graphitization without both
carbonization process and boron catalyst, yielding a very poor
crystallinity. Meanwhile, wPET-B5-G24 obtained after the boron-
Fig. 4. XRD patterns of PET-derived products obtained after heat treatment at
2400  C according to boron addition amount. Insets: magnified images of 40–60
induced catalytic graphitization shows the significantly higher
and 75–90 regions. For easy comparison, XRD patterns of synthetic graphite and value of 80.6%, indicative of a high-quality graphite structure.
natural graphite are added at the bottom of the patterns. Notably, this value is comparable to that of natural graphite
(89.5%), and more significantly, it is much higher than that of
synthetic graphite (68.9%) derived from the well-known graphi-
tizable carbon source of mesophase pitch.
To quantify the crystallinity, the crystallite sizes of Lc and La Fig. 6 shows TEM images of wPET-C9, wPET-DG24, and wPET-
were calculated using the Scherrer equation. Lc is the measure of B5-G24. In Fig. 6(a), wPET-C9 shows crystallites of a few
crystallite size along the c-axis in graphite structure (perpendicu- nanometers in lateral size forming clusters of several stacked
lar to the graphene layers), and La is along the a-axis (in-plane layers; the clusters lack regular orientation, indicating an
direction). Fig. 5(a) shows the resulting values plotted against the amorphous carbon structure. Similarly, wPET-DG24 (Fig. 6b)
amount of boron addition. When graphitized without boron shows no ordered structure; it remains amorphous even after
(wPET-B0-G24), the crystallite sizes are 2.1 nm and 4.3 nm for Lc graphitization heat treatment. However, wPET-B5-G24 clearly
and La, respectively. However, as the amount of boron increases, shows a graphite structure with parallel stacks of graphene layers
the values for Lc and La are rapidly increased, consequently and a higher degree of orientation compared to those of the former
reaching 20.9 nm and 31.1 nm, respectively, in wPET-B5-G24. These two samples.
values are comparable to those of synthetic graphite and natural
graphite. The calculated Lc values of synthetic graphite and natural Mechanism of PET graphitization
graphite are 25.6 nm and 21.0 nm, respectively; the La values are
44.1 nm and 45.3 nm, respectively. This shows that the crystal Non-graphitizable carbons, after heat-treatment at graphitiza-
growths of wPET-B5-G24 along both the c- and a-axes progress to tion temperatures, generally exhibit complicated (002) diffraction
an extent comparable to those of natural graphite and synthetic peaks, which can be separated into three components of A
graphite. (amorphous), T (turbostratic), and G (graphitic) [44]. Turbostratic

Fig. 5. (a) Crystallite sizes of Lc and La calculated based on XRD patterns of PET-derived carbons as a function of boron content. (b) Degree of graphitization of wPET-DG24,
wPET-B5-G24, synthetic graphite, and natural graphite.
454 S. Ko et al. / Journal of Industrial and Engineering Chemistry 83 (2020) 449–458

Fig. 6. TEM images of (a) wPET-C9, (b) wPET-DG24, and (c) wPET-B5-G24.

carbon is intermediate between amorphous and graphitic carbon The A- and g-components form a large portion of the wPET-B0-G24
[45], with irregularly stacked layers and interlayer (002) spacings (002) peak, while the T-component has a small contribution and no
of approximately 0.344 nm. The (002) spacing of ideal graphite is G-component is observed. This implies that wPET-B0-G24 mainly
0.335 nm; increases in structural disorder cause increases in this comprises amorphous carbon with long aliphatic chains and no
value. We hypothesized that the PET-derived carbons would graphite. Upon the addition of boron, noticeable changes are found.
exhibit similar tendencies with these three different structures, Fig. 7(b) shows the deconvolution result for wPET-B1-G24, wherein
and attempted deconvolution of the (002) peaks in their XRD the intensities of the A- and g-components are significantly
patterns. However, we could not obtain good peak fitting results decreased compared to those of wPET-B0-G24, while the intensity
using only these three components, probably due to the presence of T-component is increased and a G-component appears at 26.35 as
of residual alkyl side chains. Therefore, we introduced another a very sharp peak. This tendency is confirmed with the samples
component to the (002) peak deconvolution, which was the produced with higher boron contents. As the boron content
g-component corresponding to the stacking of alkyl side chains. increases, the intensities of the T- and G-components are rapidly
Such a g-component is frequently used in the structural analysis of increased, accompanying continuous decreases in the intensities of
asphaltene molecules that are common in heavy aromatic oils the g-and A-components.
[46–48]. Asphaltene molecules comprise PAHs and long alkyl side To observe these changes in more detail, the relative intensities
chains; the presence of these alkyl chains yields one peak at of each component are plotted as a function of boron content
approximately 20 in XRD patterns, offering important structural (Fig. 8a), where the relative intensities are obtained by dividing
information related to the structure of alkyl chains. With the each intensity by the sum of all components. For the A- and
g-component in the (002) peak deconvolution of the XRD patterns g-components, their intensities are steadily decreased as the
of PET-derived carbons, we obtained peak fittings with good boron content increases. They show relative intensities of 63% and
coherences. The presence of g-components in PET-derived carbons 22% for 0 wt% boron; these are finally decreased down to 5% and 2%
implies that they contain residual alkyl side chains as well as PAHs, for 5 wt% boron, respectively. This accompanies the continuous
even after graphitization. The presence of these alkyl chains was and rapid increase in the intensity of the G-component, which is
confirmed by FTIR spectral analysis of wPET-B0-G24 (Fig. S2, initially 0% for 0 wt% boron and finally reaches 65% for 5 wt% boron.
Supplementary material). The spectra showed groups of peaks at However, the T-component shows a different tendency; it shows a
3000–2800 cm1, attributed to sp3 aliphatic C H bond stretching steady increase in intensity to 49% for 3 wt% boron and then a
involving both CH2  and CH3 [49]. In particular, the peak at decrease to 29% for 5 wt% boron.
2924 cm1 assigned to  CH2  was more intense than the others, Fig. 8(b) shows the (002) interlayer spacing values of the four
indicating the presence of long alkyl chains in the structure of components plotted against the boron content. As mentioned
wPET-B0-G24. above, the (002) spacing reflects the degree of graphitization in
The deconvolution results are presented in Fig. 7. Firstly, the carbon materials. Spacing values for the g-component are
deconvolution of the (002) peak for wPET-B0-G24 (Fig. 7a) shows somewhat larger than those of the other three components, with
g-, A-, and T-components at 21.10 , 25.29 , and 25.90 , respectively. a decrease from 4.207 to 4.047 Å for 0–5 wt% boron, respectively.
S. Ko et al. / Journal of Industrial and Engineering Chemistry 83 (2020) 449–458 455

Fig. 7. Deconvolutions of (002) peaks on XRD patterns of (a) wPET-B0-G24, (b) wPET-B1-G24, (c) wPET-B2-G24, (d) wPET-B3-G24, (e) wPET-B4-G24, and (f) wPET-B5-G24.

Fig. 8. (a) Relative intensities of g (aliphatic), A (amorphous), T (turbostratic), and G (graphitic) structural components as a function of boron content. (b) (002) interlayer
spacing of the structural components as a function of boron content.

Similarly, the spacing values for the A- and T-components show results yielded further insight into the mechanism of structural
decreasing tendencies with values from 3.519 Å and 3.437 Å (0 wt% change from a disordered structure to graphite.
boron) to 3.490 Å and 3.416 Å (5 wt% boron), respectively. These It is likely that the boron-assisted development of graphite
indicate gradual improvements in the crystallinities of the g-, A-, structure in the current study was governed by two mechanisms.
and T-components. Meanwhile, the spacing of the G-component The first is due to doped boron. Boron atoms readily diffuse into
remains nearly constant for all compositions with the average disordered carbon structures through crystal defects and substi-
value of 3.373 Å. This indicates no increase in crystallinity, but only tute for carbon atoms at lattice positions [50,51]. This doped boron
an increase in the relative prevalence of the G-component. These removes crystal defects and relaxes lattice distortion by diffusion
456 S. Ko et al. / Journal of Industrial and Engineering Chemistry 83 (2020) 449–458

into the carbon lattice [52], which possibly enables the rearrange- appears near 282.0 eV and is increased in intensity as the boron
ment of adjacent layers and the formation of new sp2 bonds, content is increased, which is ascribed to the formation of the B4C
consequently inducing gradual improvements in the crystallinity phase. Although boron readily forms a solid solution with carbon
of disordered carbon. The increased prevalence of the turbostratic occupying substitutional sites, it also readily forms the B4C phase.
structure represented by the T-component in Fig. 8(a) was B4C is highly effective in promoting the graphitization of
attributed to the doped boron. Boron diffused into the amorphous disordered carbons [24,58]. When a disordered carbon is heat-
structure during the graphitization of the PET-derived carbons and treated at graphitizable temperatures in the presence of B4C, the
facilitated the elimination of defects involving sp3 carbons, thus carbon is dissolved and enter into the B4C phase and reprecipitated
causing the conversion of amorphous structures with aliphatic as graphite instead of disordered carbon [24]. This is because
chains to turbostratic structures. graphite is thermodynamically more stable than disordered
However, the formation of graphite structure probably occurred carbon. In this mechanism, the disordered structure collapses
differently. It would be very difficult to form graphite only through entirely to allow the formation of a new and more stable graphite
the diffusion of doped boron. Graphitizable carbon or soft carbon structure, regardless of the graphitizability of the precursor
that can be graphitized at high temperatures comprises crystallites carbon. In the same way, in the present study, the appearance
that are small in size but well-oriented in a certain direction, of the B4C phase during the graphitization of the PET-derived
exhibiting a structure similar to that of graphite. Thus, when carbons was observed, which could have caused the transition
graphitizable carbon is heat-treated at high temperatures above from disordered carbon to graphite.
2000  C, connections between the crystallites form with long-
range ordering through dehydrogenative condensation reactions, Exfoliation of PET-derived graphite
thus yielding the graphite structure. In other words, even before
graphitization, soft carbon already possesses crystallites that are In order to examine whether the PET-derived graphite could be
well-aligned in a manner similar to that found in graphite, and used as a graphene source, wPET-B5-G24 was exfoliated into
graphitization treatment only serves to connect them. Meanwhile, graphene sheets through a microwave-assisted liquid-phase
with non-graphitizable carbon or hard carbon, the orientation is exfoliation method. Before exfoliation, it was expanded by
not well-aligned but random; thus, regardless of the applied heat, chemical treatment with sulfuric acid and fuming nitric acid
long-range ordering cannot be created and graphite cannot form. under concomitant microwave radiation. Fig. 10(a) and (b) show
For the graphitization of such hard carbon, both the alignment of SEM images of the pristine and expanded forms of wPET-B5-G24,
crystallites within the disordered structure and their connection respectively. The acid treatment followed by microwave irradia-
are required. However, boron doped in crystal lattice does not tion clearly expands the wPET-B5-G24, which exhibits a pleated
cause the alignment of all crystallites. Instead, it only removes morphology with spacing between separated sheets. The expand-
partial crystal defects such as sp3 carbon atoms and lattice ed wPET-B5-G24 was then immersed in NMP to induce the
vacancies. Therefore, another mechanism must occur in the intercalation of solvent molecules in the separated sheets and
conversion of non-graphitizable carbon to graphite. subsequently exfoliated by ultrasonication treatment, which, in
XPS results provide support for this second mechanism. The turn, yielded PET-derived graphene sheets. A TEM image of thus
high-resolution XPS spectra of the C 1s regions for the boron- prepared graphene sheets is shown in Fig. 10(c, d). It exhibits a flat
doped samples are shown in Fig. 9. A distinct peak shift of sp2 and transparent morphology; some are partly wrinkled and curled
carbon towards lower binding energies is observed with increasing due to the thinness and flexibility of the resulting graphene sheets.
boron content. For pristine carbon without boron, the peak located In terms of lateral size, the average size was 410 nm with the
at 284.3 eV is ascribed to sp2 carbon [53–55], but it is moved to maximum of about 1.2 mm. The topographic profile measured by
284.0 eV with 1% boron. As the boron content increases, the sp2 AFM is shown in Fig. 10(e), which demonstrates the layer thickness
carbon peak is further shifted towards lower energies, finally of approximately 4.8 nm sheets, corresponding to few-layered
reaching 283.8 eV for 5 wt% boron. This peak shift is due to graphene. All these microscope data clearly show that the wPET-
substitutional doping of boron in the carbon lattice [55–57]. This derived graphite is capable of expansion by acid and exfoliation
supports the first mechanism of the structural development by the into few-layered graphene sheets, like common graphite. In
diffusion of doped boron. In addition to this peak, a new peak addition, we examined applicability of the resulting graphene
sheets by fabricating into a graphene paper and measuring its
sheet resistance. To fabricate a graphene paper, we applied the
vacuum filtration method that is a well-known method widely
used when fabricating a graphene paper [59]. Following this
method, graphene samples with bad quality normally produce
uneven papers showing many wrinkles or cracks on its surface.
However, in the case of wPET-derived graphene sheets sample, it
showed very good quality, producing a paper with very even
surface without being wrinkled or cracked; its photo image is
shown in Fig. S2 (Supplementary material). As to the measurement
of sheet resistance, it showed an average value of 41.1 V/& as
shown in Fig. 10(f). For the comparison, we also applied the very
same procedure to natural graphite-derived graphene sheets.
Notably, the natural graphite-derived one showed almost the same
value of 41.8 V/&, indicating that wPET-derived graphite sheets
sample has the same performance with that of natural graphite-
derived one in terms of electrical property. Thus, it was
demonstrated that the wPET-derived graphene has a high potential
in applications like electrical devices showing good processability
Fig. 9. High-resolution C 1s XPS spectra of wPET-B0-G24, wPET-B1-G24, wPET- and good electrical property compatible to that of natural graphite-
B2-G24, wPET-B3-G24, wPET-B4-G24, and wPET-B5-G24. derived one.
S. Ko et al. / Journal of Industrial and Engineering Chemistry 83 (2020) 449–458 457

Fig. 10. SEM images of (a) pristine and (b) expanded wPET-B5-G24. (c, d) TEM images and (e) AFM image of graphene sheets derived from wPET-B5-G24. (f) Sheet resistance of
wPET-B5-G24 graphene sheets and natural graphite-derived graphene sheets.

Conclusion proper treatments. Related studies are currently in progress in our


group. We expect that our work present here can guide the
In the current study, we introduced a new route for the innovative upcycling of waste plastics.
upcycling of PET waste, wherein synthetic graphite was success-
fully prepared through a facile and scalable method. The synthetic Conflict of interest
method included the pyrolysis and carbonization of PET under an
inert atmosphere and subsequent graphitization in the presence of The authors declare that they have no conflict of interest.
boron as a catalyst. This method overcame the non-graphitizability
of PET and yielded graphite with high crystallinity comparable to Acknowledgments
that of natural graphite. Furthermore, the thus-prepared PET-
derived graphite was effectively expanded and exfoliated into This work was supported by the Korea Evaluation institute of
graphene sheets. The PET-derived graphite presented here possibly Industrial Technology through Development of Core Industrial
has many potential applications for various purposes, such as Technology (10082582, Development of petroleum-based high
anode materials in secondary batteries and fillers for carbon quality mesophase pitch and high yield mesophase pitch for
composites materials, and may serve as an alternative source for premium carbon materials) and Synthetic Graphite Development
graphene production. While the current study only investigated Project (20006642, Development of high quality precursors for
the utilization of PET, we assume that the conversion of other premium grade synthetic graphite) funded by Ministry of Trade,
heavily used plastics such as PE, PP, and PS is also plausible with the Industry and Energy (MOTIE, Korea).
458 S. Ko et al. / Journal of Industrial and Engineering Chemistry 83 (2020) 449–458

Appendix A. Supplementary data [28] T. Liu, R. Luo, S.-H. Yoon, I. Mochida, J. Power Sources 195 (2010) 1714.
[29] B. Xing, C. Zhang, Y. Cao, G. Huang, Q. Liu, C. Zhang, Z. Chen, G. Yi, L. Chen, J. Yu,
Fuel Process. Technol. 172 (2018) 162.
Supplementary material related to this article can be found, in [30] E. Rodríguez, I. Cameán, R. García, A.B. García, Electrochim. Acta 56
the online version, at doi:https://doi.org/10.1016/j.jiec.2019.12.018. (2011) 5090.
[31] S. Mallakpour, M. Javadpour, Ultrason. Sonochem. 40 (2018) 611.
[32] S. Ko, C.W. Lee, J.S. Im, J. Ind. Eng. Chem. 36 (2016) 125.
References [33] P. Lu, Y. Sun, H. Xiang, X. Liang, Y. Yu, Adv. Energy Mater. 8 (2018)1702434.
[34] L.M. Manocha, M. Patel, S.M. Manocha, C. Vix-Guterl, P. Ehrburger, Carbon 39
[1] R. Geyer, J.R. Jambeck, K.L. Law, Sci. Adv. 3 (2017)e1700782. (2001) 663.
[2] J.R. Jambeck, R. Geyer, C. Wilcox, T.R. Siegler, M. Perryman, A. Andrady, R. [35] S. Ko, J.-E. Choi, C.W. Lee, Y.-P. Jeon, J. Ind. Eng. Chem. 54 (2017) 252.
Narayan, K.L. Law, Science 347 (2015) 768. [36] T. Senda, Y. Yamada, M. Morimoto, N. Nono, T. Sogabe, S. Kubo, S. Sato, Carbon
[3] J. Ren, X. Dyosiba, N.M. Musyoka, H.W. Langmi, B.C. North, M. Mathe, M.S. 142 (2019) 311.
Onyango, Int. J. Hydrogen Energy 41 (2016) 18141. [37] D. Liu, B. Lou, G. Chang, Y. Zhang, R. Yu, Z. Li, C. Wu, M. Li, Q. Chen, Fuel 231
[4] V.G. Pol, Environ. Sci. Technol. 44 (2010) 4753. (2018) 495.
[5] A.R. Kamali, J. Yang, Q. Sun, Appl. Surf. Sci. 476 (2019) 539. [38] S. Ko, J.-E. Choi, H. Yim, J. Miyawaki, S.-H. Yoon, Y.-P. Jeon, Carbon Lett. 29
[6] R. Mendoza-Carrasco, E.M. Cuerda-Correa, M.F. Alexandre-Franco, C. Fernán- (2019) 307.
dez-González, V. Gómez-Serrano, J. Environ. Manag. 181 (2016) 522. [39] S.-H. Lee, D.-S. Kang, S.-M. Lee, J.-S. Roh, Carbon Lett. 26 (2018) 11.
[7] W. Bratek, A. Swia  ˛tkowski, M. Pakuła, S. Biniak, M. Bystrzejewski, R. [40] H.-S. Kil, K. Oh, Y.-J. Kim, S. Ko, Y.P. Jeon, H.-I. Joh, Y.-K. Kim, S. Lee, J. Anal. Appl.
Szmigielski, J. Anal. Appl. Pyrolysis 100 (2013) 192. Pyrolysis 136 (2018) 153.
[8] S. Ayyalusamy, S. Mishra, V. Suryanarayanan, Sci. Rep. 8 (2018) 13151. [41] D.-W. Kim, H.-S. Kil, J. Kim, I. Mochida, K. Nakabayashi, C.K. Rhee, J. Miyawaki,
[9] E. Lorenc-Grabowska, M.A. Diez, G. Gryglewicz, J. Colloid Interface Sci. 469 S.-H. Yoon, Carbon 121 (2017) 301.
(2016) 205. [42] J. Maire, J. Mering, Chem. Phys. Carbon 6 (1970) 125.
[10] J. Ma, J. Liu, J. Song, T. Tang, RSC Adv. 8 (2018) 2469. [43] S. Faraji, O. Yildiz, C. Rost, K. Stano, N. Farahbakhsh, Y. Zhu, P.D. Bradford,
[11] H.H. Mohamed, A.A. Alsanea, N.A. Alomair, S. Akhtar, D.W. Bahnemann, Carbon 111 (2017) 411.
Environ. Sci. Pollut. Res. 26 (2019) 12288. [44] C.L. Burket, R. Rajagopalan, H.C. Foley, Carbon 46 (2008) 501.
[12] J. Yang, Y.X. Jin, X.P. Yu, Q.F. Yue, J. Sol-Gel Sci. Technol. 83 (2017) 413. [45] B. Manoj, A. Kunjomana, Int. J. Electrochem. Sci. 7 (2012) 3127.
[13] B. Kaur, J. Singh, R.K. Gupta, H. Bhunia, J. Environ. Manag. 242 (2019) 68. [46] N.T. Nguyen, K.H. Kang, C.W. Lee, G.T. Kim, S. Park, Y.-K. Park, Fuel 235
[14] N.A. El Essawy, S.M. Ali, H.A. Farag, A.H. Konsowa, M. Elnouby, H.A. Hamad, (2019) 677.
Ecotoxicol. Environ. Saf. 145 (2017) 57. [47] H.P. Halim, J.S. Im, C.W. Lee, Carbon Lett. 14 (2013) 152.
[15] A. Joseph Berkmans, M. Jagannatham, S. Priyanka, P. Haridoss, Waste Manag. [48] F.S. AlHumaidan, A. Hauser, M.S. Rana, H.M.S. Lababidi, M. Behbehani, Fuel 150
34 (2014) 2139. (2015) 558.
[16] R. Shamsi, G. Mir Mohamad Sadeghi, G.H. Asghari, Polym. Compos. 39 (2018) E754. [49] F. Meng, J. Yu, A. Tahmasebi, Y. Han, H. Zhao, J. Lucas, T. Wall, Energy Fuels 28
[17] N.R. Kim, H.J. An, Y.S. Yun, H.-J. Jin, Carbon Lett. 22 (2017) 110. (2014) 275.
[18] E.C. Evarts, Nature 526 (2015) S93. [50] I. Mochida, S.H. Yoon, Y. Korai, Chem. Rec. 2 (2002) 81.
[19] B.E. Barton, M.J. Behr, J.T. Patton, E.J. Hukkanen, B.G. Landes, W. Wang, N. [51] S. Ha, G.B. Choi, S. Hong, D.W. Kim, Y.A. Kim, Carbon Lett. 27 (2018) 1.
Horstman, J.E. Rix, D. Keane, S. Weigand, M. Spalding, C. Derstine, Small 13 [52] L.E. Jones, P.A. Thrower, Carbon 29 (1991) 251.
(2017)1701926. [53] J. Zemek, J. Houdkova, P. Jiricek, M. Jelinek, Carbon 134 (2018) 71.
[20] G.A. Cooper, R.M. Mayer, J. Mater. Sci. 6 (1971) 60. [54] S. Kaciulis, Surf. Interface Anal. 44 (2012) 1155.
[21] N. Yusof, A.F. Ismail, J. Anal. Appl. Pyrolysis 93 (2012) 1. [55] Z.-H. Sheng, H.-L. Gao, W.-J. Bao, F.-B. Wang, X.-H. Xia, J. Mater. Chem. 22
[22] H. Wang, H. Xiao, Y. Lu, J. Jiang, J. Mater. Sci. 51 (2016) 10690. (2012) 390.
[23] J.Y. Howe, L.E. Jones, Carbon 42 (2004) 461. [56] Z.-X. Wang, X.-H. Yu, F. Li, F.-Y. Kong, W.-X. Lv, W. Wang, J. Mater. Chem. B 6
[24] A. Oya, R. Yamashita, S. Otani, Fuel 58 (1979) 495. (2018) 1771.
[25] A.P. Luz, C.G. Renda, A.A. Lucas, R. Bertholdo, C.G. Aneziris, V.C. Pandolfelli, [57] R.K. Das, S. Mohapatra, J. Mater. Chem. B 5 (2017) 2190.
Ceram. Int. 43 (2017) 8171. [58] S. Fan, L. He, C. Yang, H. Li, S. Wang, Y. Lin, B. Zou, J. Deng, L. Zhang, L. Cheng,
[26] D.H. Zhong, H. Sano, Y. Uchiyama, K. Kobayashi, Carbon 38 (2000) 1199. Mater. Sci. Eng.: A 706 (2017) 201.
[27] F. Sun, H.B. Wu, X. Liu, F. Liu, R. Han, Z. Qu, X. Pi, L. Wang, J. Gao, Y. Lu, Mater. [59] Y.J. Kwon, Y. Kwon, H.S. Park, J.U. Lee, Adv. Mater. Interfaces 6 (2019)1900095.
Today Energy 9 (2018) 428.

Das könnte Ihnen auch gefallen