You are on page 1of 7

Journal of The Electrochemical Society, 163 (5) A785-A791 (2016) A785

0013-4651/2016/163(5)/A785/7/$33.00 © The Electrochemical Society

Composites of TiO2 Nanoparticles Deposited on Ti3 C2 MXene


Nanosheets with Enhanced Electrochemical Performance
JianFeng Zhu,a,=,z Yi Tang,a,=,z ChenHui Yang,b Fen Wang,a and MinJuan Caoa
a Key Laboratory of Auxiliary Chemistry & Technology for Chemical Industry, Ministry of Education,
Shaanxi University of Science & Technology, Xi’an 710021, People’s Republic of China
b Electronic Materials Research Laboratory, School of Electronic and Information Engineering, and International
Center for Dielectric Research, Xi’an Jiaotong University, Xi’an 710049 Shaanxi, People’s Republic of China

MXene-based materials are promising electrode materials for electrochemical capacitors (ECs) due to their unique two-dimensional
layered structure, high surface area, remarkable chemical stability, and electrical conductivity. TiO2 nanoparticles decorated Ti3 C2
MXene were synthesized through a simple in situ hydrolysis and heat-treatment process and subsequently fabricated as an electrode
for ECs. The as-prepared Ti3 C2 , TiO2 , and TiO2 -Ti3 C2 were characterized by X-ray diffraction, scanning electron microscopy,
transmission electron microscopy, and X-ray photoelectron spectroscopy. The results indicated that TiO2 nanoparticles with a
diameter of less than 30 nm were decorated onto the Ti3 C2 MXene nanosheets. The resulting composites exhibited significantly
higher specific capacitance of 143 F g−1 at 5 mV s−1 , which was 1.5 times that of pure Ti3 C2 (93 F g−1 ). Moreover, TiO2 -Ti3 C2
showed excellent cycling stability, retaining ∼92% of its initial capacitance after 6000 cycles. These results suggest that TiO2 -Ti3 C2
nanocomposite has the potential as an electrode material for high-performance energy storage devices.
© 2016 The Electrochemical Society. [DOI: 10.1149/2.0981605jes] All rights reserved.

Manuscript submitted September 30, 2015; revised manuscript received January 26, 2016. Published February 19, 2016.

Electrochemical capacitors (ECs), also called ultracapacitors or nanotubes/nanowires,31–35 and nanosheets36–39 between the graphene
electrochemical double layer capacitors,1 have attracted a tremendous layers has been shown to prevent the restacking of the latter, lead-
amount of attention as energy-storage devices due to their high power ing to greatly improved performance in supercapacitors, Li-ion, and
density, fast charge–discharge ability, excellent reversibility, and long Li–S batteries.12 Notably, one of the most effective ways for doing
cycling life.2–5 Due to these advantages, supercapacitors make up this is employing various transition metal oxides to insert among the
many markets ranging from electronics to transportation and station- nanosheets.5,40–42
ary applications.6 There are two types of ECs, which differ by the Among the transition metal oxides, titanium dioxide (TiO2 ) is con-
charge storage mechanism: the first, known as electrical double-layer sidered as one of the most promising candidates to be applied in Li-
capacitors (EDLCs), in which the capacity is due to the electrosorption ion capacitor43 and supercapacitor,44 due to its low cost, nontoxicity,
of ions on porous carbon electrodes, have limited energy density.6,7 eco-friendliness, abundant availability high surface area, and ease of
Typical EDLC materials have a high surface area, such as activated preparation in defined nanoscale dimensions.40,45 Hence, nano-scale
carbon, carbon nanotubes, and graphene-based active materials.3,5,8 TiO2 with different morphologies is a practical candidate for low-cost
The second, known as pseudocapacitors, in which the capacity is electrode material in supercapacitor applications. For example, Rakhi
due to redox reactions, provide higher energy densities but usually et al.14 stated that the Ti2 C MXene nanosheets annealed in air con-
suffer from shorter cyclic lifetimes.4,5,9 The typical pseudocapacitor sisted of nanosheets and numerous TiO2 nanocrystals on thin graphitic
materials have transition metal oxides and conductive polymers.2,4,10 nanosheets, which were similar to the ones reported by Naguib et al.46
To overcome these obstacles, research has been carried out on the However, the oxidized Ti2 C MXene nanosheets used in ECs had poor
development of new materials,11 hybrid structure,12,13 and surface capacitive performance (specific capacitance value less than 5 F/g at
modification,14 for EC eletrodes. 5 mV/s, and rate performance 35%). Naguib et al.46 showed a similar
Recently, MXenes (of the formula Mn+1 Xn Tx , where M is a transi- one-step synthesis by heating 2D Ti3 C2 in air at 1150◦ C for 30 s. The
tion metal, X is C and/or N, and Tx denotes -OH, -F, and =O surface resulting TiO2 nanocrystals were enmeshed in thin sheets of disor-
groups), are a novel family of two-dimensional (2D) metal carbides, dered graphitic carbon structures which could handle extremely high
which can be produced by the selective etching of the A-group (gener- cycling rates when tested as anodes in lithium-ion batteries (LIBs).
ally group IIIA and IVA elements) layers from the MAX phases.15–20 But the method was limited by the uncontrolled synthesis condition.
MXenes have already demonstrated their potential as promising elec- It should be mentioned that the hybrid TiO2 -C phase was differ-
trode materials for Li-ion batteries,21–24 supercapacitors,6,7,12,13 and ent from the TiO2 -Ti3 C2 phase. Recently, Gao et al.47 reported that
sensors,19,25,26 because of their high electrical conductivity, large sur- TiO2 /Ti3 C2 nanocomposites could be fabricated by a hydrothermal
face area, layered structure, remarkable chemical stability, and en- process in which titanium sulfate (TiSO4 ) and Ti3 C2 were mixed and
vironmentally friendly characteristics.7,12,18,27,28 Particularly, Ti3 C2 is autoclaved. The resulted hybrid structure showed good photocatalytic
one of the most widely studied and most promising members of this performance to methyl orange (MO) under ultraviolet light illumi-
family,7,15,17,19,25,26 which is frequently applied in ECs. Ti3 C2 MXene nation. Our previous work revealed that TiO2 /Ti3 C2 nanocompos-
is one of very few materials which exhibits “true” pseudocapacitive be- ite could be synthesized by hydrolysis of tetrabutyl titanate (TBOT)
havior. It presents a continuous change in the titanium oxidation state and an microwave hydrothermal system.25 The prepared TiO2 /Ti3 C2
during charge/discharge, producing rectangular-shaped CVs. Such be- nanocomposite exhibited excellent electrochemical performance for
havior can be attributed to the 2D nature of Ti3 C2 MXene: spontaneous direct electrochemical biosensor.
ion intercalation naturally provides access to electrochemically active In this work, multilayered TiO2 -Ti3 C2 nanocomposites were syn-
transition metal oxide surfaces and the conductive carbide layer en- thesized by a simple in situ hydrolysis of TBOT followed by a heat-
sures rapid charge transfer.11 treatment process in vacuum, and subsequently used as a novel elec-
To further enhance electrochemical performances of the 2D layerd trode material for supercapacitor. The schematic in Fig. 1 shows nu-
Ti3 C2 electrode material, one straightforward strategy which has been merous TiO2 nanoparticles were deposited on the multilayered Ti3 C2
extensively investigated is the introduction of interlayer spacers.12 For nanosheets substrate, which made it possible for the nanocomposite to
the typical 2D graphene, the incorporation of nanoparticles (NPs),29,30 offer a number of advantages when it was used as an electrode material
for supercapacitor energy storage. The synergetic effects of Ti3 C2 and
TiO2 endowed TiO2 -Ti3 C2 nanocomposites with excellent properties
=
These authors contributed equally to this work. and improved functionalities. Ti3 C2 and TiO2 nanoparticles offered
z
E-mail: zhujf@sust.edu.cn; tangyi150@163.com excellent connection between them and facilitated electron exchange.

Downloaded on 2016-02-23 to IP 113.200.58.77 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
A786 Journal of The Electrochemical Society, 163 (5) A785-A791 (2016)

Figure 1. Schematic showing the procedure used herein to prepare TiO2 -Ti3 C2 nanocomposite electrodes.

Furthermore, exposed TiO2 on the surface of the Ti3 C2 nanolayers times and deionized water three times. Then the precipitate was dried
was beneficial for improving the contact with electrolyte and inser- at 40◦ C. Afterward, the dried product was heat treated by 2◦ C/min at
tion/extraction of cations, which resulted in higher capacitance and 500◦ C for 4 h in vacuum (less than 4.0 × 10−2 Pa). The nanocomposite
presented new avenues for research in energy conversion and storage. (TiO2 -Ti3 C2 ) was then collected. Figure S1 schematically illustrates
Herein, compared to TiO2 and Ti3 C2 , the synthesized TiO2 -Ti3 C2 the procedure used to prepare TiO2 -Ti3 C2 nanocomposite. Figures S3
nanocomposite with advantageous synergistic nano-sized effects ex- (c)-(d) show the SEM image of TiO2 -Ti3 C2 . Meanwhile, individual
hibited excellent properties in a supercapacitor. The higher specific TiO2 and Ti3 C2 for control were synthesized separately following the
surface area of TiO2 -Ti3 C2 and greater ion diffusion process between same procedure. Figure S4 is an SEM image of TiO2 .
TiO2 -Ti3 C2 and the electrolyte could enhance specific capacitance,
the rate capability, and exceptional durability. These results indicated Materials characterization.—X-ray diffraction (XRD) patterns
that TiO2 -Ti3 C2 nanocomposite has potential as an electrode material were recorded on a Rigaku D/max 2200pc diffractometer using Cu
for high-performance ECs. Kα radiation of wavelength λ = 0.15418 nm at 40 kV and 40 mA.
Raman spectra were collected through a Renishaw 2000 model con-
Experimental focal microscopy Raman spectrometer with a CCD detector and a
holographic notch filter (Renishaw Ltd., Gloucestershire, U.K.) at
PTFE (60 wt%) was purchased from Sigma. Unless otherwise ambient conditions, using the radiation of 514.5 nm from an air-
noted, all other chemicals (analytical grade) were purchased from cooled argon ion laser to excite the SERS. Field-emission scanning
Sinopharm Chemical Reagent Co., Ltd., China. In all experiments electron microscopy (FE-SEM) and energy-dispersive X-ray analy-
deionized water was used. sis (EDAX) were performed using a Hitachi S-4800 &Hiroba en-
ergy dispersive X-ray electron miscroscope. Transmission electron
Synthesis of Ti3 C2 MXene.—Ti3 C2 was successfully prepared microscopy (TEM; FEI company Tecnai G220 S-twin, 200 kV) was
by etching Al from Ti3 AlC2 in HF at room temperature.17,19 First, used to observe the morphology of the samples. X-ray photoelectron
3.0 g as-prepared Ti3 AlC2 powders were immersed in 60 mL 40% spectroscopy (XPS) measurements were performed on a VG Thermo
HF solution under magnetic stirring at room temperature for 24 h. ESCALAB 250 spectrometer with an exciting source of Al Kα. Ni-
Then, the resulting MXene suspension was washed six times using trogen sorption isotherms were measured at 77 K on an automatic N2
deionized water and centrifuged to separate the powder at 4000 rpm adsorption/desorption instrument (ASAP 2460, Micromeritics Instru-
until the pH value of the liquid reached ∼6. After decantation, the ment Corporation, Norcross (Atlanta), Georgia, USA) and the samples
resulting powder was washed three times with absolute ethanol and were outgassed in vacuum at 200◦ C for >3 h before the test. The spe-
centrifuged to separate the powder, and left to dry in air at RT for cific surface areas of the materials were calculated by the Brunauere
3 days. Finally, the powder was dried in the vacuum oven (<0.09 Emmette Teller (BET) method. The pore size distributions were de-
MPa) at 60◦ C for 48 h. Figure S3(b) show the SEM image of Ti3 C2 . rived from the adsorption branch using the Barrette JoynereHalenda
(BJH) model. The total pore volume was determined from the amount
Synthesis of TiO2 -Ti3 C2 nanocomposite.—In a typical prepara- of N2 uptake at P/P0 = 0.99.
tion of TiO2 -Ti3 C2 nanocomposite, 150 mg of Ti3 C2 was dispersed in
200 mL ethanol under magnetic stirring at room temperature for 1 h to Fabrication of electrodes.—To investigate the electrochemical be-
obtain a homogenous suspension of Ti3 C2 . The suspension was mixed havior of TiO2 -Ti3 C2 , a three-electrode experimental cell was assem-
0.5 mL of 0.4 mM KCl and then stirred for 20 min. Next, 1.0 mL of bled. A mixture of 80 wt% the as-perpared TiO2 -Ti3 C2 nanocom-
tetrabutyltitanate (TBOT) was added to the mixture and stirred for 6 h posite, 15 wt% acetylene black, and 5 wt% polytetrafluoroethylene
in a dry atmosphere. At the end of the reaction, a gray precipitate was (PTFE) binder was fabricated using ethanol as a solvent and mixed
collected by centrifugation and rinsed sequentially with ethanol six in an agate mortar. At this stage, the resultant slurries exhibited

Downloaded on 2016-02-23 to IP 113.200.58.77 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 163 (5) A785-A791 (2016) A787

Figure 2. SEM image of (a) typical as-fabricated Ti3 C2 ; (b)-(c) TiO2 -Ti3 C2 nanocomposite; TEM image of (d) typical as-fabricated Ti3 C2 ; (d)-(f) TiO2 -Ti3 C2
nanocomposite.

clay-like properties and could be directly processed into freestanding more, the introduction of TiO2 nanoparticles into the inner surface
films by rolling. Then the freestanding film (1 × 2 cm) was subse- of the Ti3 C2 nanolayers could effectively impede the stacking of in-
quently pressed onto nickel foam under a pressure of 20 MPa for dividual Ti3 C2 nanosheets and enlarge the distance between Ti3 C2
1 min. The prepared electrode was placed in a vacuum drying oven at nanosheets.13 Note that deposition of numerous TiO2 nanoparticles
80◦ C for 24 h. The as-prepared electrodes were then obtained. on Ti3 C2 nanolayers have larger surface area compared to pure Ti3 C2 .
For comparison, TiO2 and Ti3 C2 were used for the fabrication of Figure 2d shows a TEM image indicating that the Ti3 C2 had a per-
electrode with similar procedures as described above. fect layered structure. Meanwhile, numerous small TiO2 nanoparti-
cles were found to evenly attach on the Ti3 C2 layers, as shown in
Electrochemical setup.—All electrochemical experiments were Figs. 2e–2f. TEM observations further demonstrated that TiO2
performed with a CHI 660E electrochemical workstation (CH In- nanoparticles were deposited on Ti3 C2 nanosheets surfaces and TiO2 -
struments, Shanghai, China). A conventional three-electrode system, Ti3 C2 nanocomposites were synthesized. Hence, the incorporation
which consisted of a platinum (1 × 2 × 0.1 cm) as the counter of TiO2 nanoparticles between the Ti3 C2 nanolayers not only in-
electrode, an Ag/AgCl/3 M KCl as the reference electrode and the creased surface area of TiO2 -Ti3 C2 and enlarged interlayer space
as-prepared TiO2 , Ti3 C2 , and TiO2 -Ti3 C2 as the working electrode, between the Ti3 C2 flakes for cation intercalation, but also pro-
was used in all electrochemical experiments. All experiments were vided additional diffusion paths for electrolyte ions.12 Therefore, the
conducted in 6 M KOH as the electrolyte solution. unique multilayered TiO2 -Ti3 C2 was expected to exhibit improved
supercapacitive performances, when employed as supercapacitor
Electrochemical measurements.—Cyclic voltammetry, electro- electrodes.
chemical impedance spectroscopy, galvanostatic charge–discharge, Energy-dispersive X-ray analysis (EDAX) was employed to study
and galvanostatic cycling were performed. Cyclic voltammetry of the the distribution of the elements Ti, Al, C, O, and F in as-prepared
working electrode were measured at different scan rates of 5, 10, 20, Ti3 C2 and TiO2 -Ti3 C2 samples (Fig. S5). The relative intensity of
50, 100, and 200 mV s−1 from −1 V to −0.35 V versus Ag/AgCl. Elec- the F element peak of TiO2 -Ti3 C2 nanocomposite became less in-
trochemical impedance spectroscopy was performed at open-circuit tense than that of the spectrum of Ti3 C2 . Note that surface chemistry
potential, with a 5 mV amplitude, and frequencies that ranged from had a significant effect on the capacitive response: reduction of F-
10 mHz to 100 kHz. Galvanostatic charge–discharge was performed terminations, relative to those of O or OH, resulted in a significant
at different current densities of 0.5, 1, 1.5, 2, and 3 A g−1 between the increase in specific capacitance.14,49 Meanwhile, after deposition, the
potential limits of −1 V to −0.35 V versus Ag/AgCl. Galvanostatic increase relative intensity of C element showed the formation of a
cycling was performed at 1 A g−1 from −1 V to −0.35 V versus small amount of graphite carbon after heat-treatment. In high vacuum
Ag/AgCl. (less than 4.0 × 10−2 Pa), this was possible within oxidation of a
little Ti-OH/Ti = O on the Ti3 C2 surface, which could lead to the
formation of a little graphite carbon. The presence of a graphite car-
Results and Discussion bon indicated that the conductive carbide layer ensured rapid charge
Characterization of TiO2 -Ti3 C2 .—The morphology of as- transfer,11 which improved supercapacitive performance.
fabricated Ti3 C2 (a) and TiO2 -Ti3 C2 nanocomposite (b)–(c) were stud- The crystal structure and orientation of the as-prepared samples
ied by using SEM, as shown in Fig. 2. In Fig. 2a, the as-synthesized were studied using XRD. The diffractogram of pure TiO2 was ob-
Ti3 C2 sample possessed a layered morphology resembling exfoliated served in Fig. 3a (curve a), the peak characteristics at 25.28◦ , 37.80◦ ,
graphite,48 where the nanolayers were clearly separated from each 48.05◦ , 53.89◦ , and 55.06◦ on the 2θ scale corresponded to the (101),
other compared to the unreacted powders (Fig. S3(a)). As Figs. 2b–2c (004), (200), (105), and (211) planes of the anatase phase. The ob-
shows, after deposition, numerous TiO2 nanoparticles (less than 30 served data were in good agreement with the standard anatase phase
nm in size) could be easily loaded on the Ti3 C2 nanosheets. What’s of TiO2 (JCPDS card number 21-1272). The diffractogram of the

Downloaded on 2016-02-23 to IP 113.200.58.77 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
A788 Journal of The Electrochemical Society, 163 (5) A785-A791 (2016)

Figure 3. Characterization of TiO2 -Ti3 C2 nanocomposite. (a) XRD plots of TiO2 -Ti3 C2 , Ti3 C2 and TiO2 . (b) Raman spectroscopy of TiO2 -Ti3 C2 , Ti3 C2 and
TiO2 .

HF-etched Ti3 C2 nanosheets (curve b) in Fig. 3a, in its vacuum-dried fore, the results suggested that TiO2 -Ti3 C2 nanocomposite had been
multilayered state, showed the (00l) peaks, such as the (002), (004), successfully synthesized.
and (0010), broadened, lost intensity, and shifted to lower angles Nitrogen adsorption–desorption isotherms as shown in Fig. 4a
compared to their location before treatment (Fig. S2). Because the ex- were measured to evaluate the BET surface area and the pore size
periments were conducted in HF aqueous solution environment, -OH, distribution of Ti3 C2 and TiO2 -Ti3 C2 samples. The N2 adsorption–
-F and/or =O were the probable ligands. The EDAX pattern of HF desorption isotherms at –195.8◦ C showed an increase in the specific
treated sample (Fig. S5) further confirmed the removal of Al and the surface area (SSA) after deposition of TiO2 . The SSA estimated using
presence of OH, F and O groups, indicating the possible surface ter- the BET equation51 for the deposited sample was found to be 31.5679
mination in the exfoliated nanosheets with F, OH, or O groups. And ± 0.6176 m2 g−1 , a factor of 4 greater than that of the as-synthesized
the experimental results provided strong evidence of the formation Ti3 C2 (8.1792 m2 g−1 ). This increase in SSA could be explained by the
of Ti3 C2 (OH)x Fy Oz ,17,28 labeled Ti3 C2 for short. Curve c exhibited formation of numerous nanometer-sized TiO2 particles, and enlarging
XRD pattern of the as-prepared TiO2 -Ti3 C2 hybrid nanostructure, all interlayer space between the Ti3 C2 flakes after heat-treatment.
the diffraction peaks were good agreement with that of anatase TiO2 The obvious hysteresis loop between adsorption and desorption
(curve a) and Ti3 C2 (curve b), which confirmed the co-existence of branches could be observed at relative pressure scope of 0.45–1.0 for
anatase TiO2 and Ti3 C2 in the nanocomposites. two samples in Fig. 4a, which demonstrated the existence of meso-
Figure 3b (curve b) shows Raman spectroscopy of the as-prepared pores. The almost vertical tails at the relative pressure near to 1.0
Ti3 C2 ; all the diffraction peaks were consistent with that the reported pointed to the presence of macropores. And, the ratio of micropores
data of Ti3 C2 .46 Raman spectroscopy of TiO2 -Ti3 C2 nanocomposites was also low for two samples. Pore size distributions of two sam-
(Fig. 3b curve c) showed a strong peak at 144 cm−1 , together with ples calculated from the nitrogen desorption branches were observed
three other peaks at 394, 513, and 635 cm−1 . These peaks could be as- in Fig. 4b, which displayed very close pore size distribution with a
signed to the following anatase vibrational modes (curve a).50 The two peak centering at ca. 15.0 nm and more (nearly 2 times) mesopores
broad peaks between 1000 and 1800 cm−1 were characteristic for the of TiO2 -Ti3 C than that of Ti3 C2 . This was in agreement with the
D- and G- modes of graphitic carbon. After heat-treatment, D- and G- N2 adsorption–desorption isotherms. Therefore, the incorporation of
band intensity increased in the Raman spectra of the sample, which TiO2 nanoparticles into the Ti3 C2 nanolayers provided larger SSA
might show the formation of a little graphite carbon. The presence and more mesopores for cation intercalation. The unique multilay-
of graphite carbon indicated that the conductive carbide layer en- ered TiO2 -Ti3 C2 could enhance supercapacitive performance when
sured rapid charge transfer,11 which improved supercapacitive perfor- employed as supercapacitor electrodes.
mances. After heat-treatment, the as-prepared nanocomposite mainly X-ray photoelectron spectroscopy (XPS), as shown in Fig. 5, was
consisted of Ti3 C2 and TiO2 from TBOT (Fig. 3a (curve c)). There- acquired to determine the nature of chemical bonding in the Ti 2p

Figure 4. (a) Nitrogen adsorption–desorption isotherms and (b) pore size distributions calculated from N2 desorption isothermals for Ti3 C2 and TiO2 -Ti3 C2
samples. The closed and open symbols refer to adsorption and desorption branches, respectively.

Downloaded on 2016-02-23 to IP 113.200.58.77 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 163 (5) A785-A791 (2016) A789

Figure 5. High-resolution XPS in the Ti 2p region for Ti3 C2 Tx , (a) before TiO2 deposition, (b) after TiO2 deposition. The fits are color-coded; the key is shown
on this figures. An increase in the Ti 2p3/2 component corresponding to TiO2 (colored magenta) can be observed at 459.2 eV.

region for Ti3 C2 , (a) before TiO2 deposition, (b) after deposition. interface and more ion diffusion process between TiO2 -Ti3 C2 and the
High-resolution XPS spectra in the Ti 2p region of the powder revealed electrolyte.
that peaks could be deconvoluted into components corresponding to Specific capacitances (Cs) of the TiO2 -Ti3 C2 electrode were cal-
Ti bound to C, Ti(II), Ti(III) and Ti(IV) peaks (Fig. 5(a)). Ti-C 2p3/2, culated by Eq. 1 at various scan rates. A capacitance of 143 F g−1
Ti(II) 2p3/2, Ti(III) 2p3/2, and Ti(IV) 2p3/2 peaks were detected at was achieved for the TiO2 -Ti3 C2 electrode at a scan rate of 5 mV s−1 ,
binding energies of 455.3, 456.1, 457.2, and 458.8 eV, respectively. which was almost 1.5 times as great as that of Ti3 C2 electrode at the
It could be seen that intensity of Ti-C peak was the strongest, Ti(IV) same scan rate. Even at 200 mV s−1 , a capacitance of 117 F g−1 was
was the weakest, together with Ti-C signals arising from Ti atoms in measured, 82% capacitance retention as the scan rate increases from
the interior of the Ti3 C2 MXene nanolayers,28,52,53 which indicated 5 to 200 mV s−1 .
that this spectrum was typical of Ti3 C2 . After TiO2 deposition, Ti(IV) The Cs values of an electrode can be calculated from the CV curve
peak corresponding to TiO2 became quite clear and the strongest using18
(Fig. 5b). The relative intensities of Ti-C, Ti(II) and Ti(III) peaks 
were less intense than that of the spectrum of Ti3 C2 . Concomitant Cs = I dV / (msV ) [1]
with the increase in the TiO2 signal, there was a decrease in the Ti3 C2
signal. The observation might show the formation of a small amount where Cs is the specific capacitance of the electrode (F g−1 ), I is the
of graphite carbon and agree with EDAX and Raman spectroscopy response current under the integerated area of the CV curves (A), m
results of TiO2 -Ti3 C2 nanocomposite. The slight shift to lower binding is the mass of the electrode material (g), s is the scan rate (mV s−1 ),
energy for the Ti-C component and the slight shift to higher binding and V is the potential window (V).
energy for the TiO2 component indicated that the original stages of Figure 6c shows the variation in specific capacitance as a function
the synthesis of these phases as the local effect of removing oxygen of scan rate. These results indicate that specific capacitances decreased
from amorphous TiO2 into anatase TiO2 could cause a shift in electron with scan rate increasing. The decreasing trend in capacitance indi-
density. All of these observations indicated that TiO2 was deposited cated that parts of the electrode surface were inaccessible at high
on the surface of multilayered Ti3 C2 . The introduction of TiO2 on the charging–discharging rates. Furthermore, TiO2 -Ti3 C2 electrode con-
Ti3 C2 surface could lead to an increase in the accessible surface area, sistently demonstrated that specific capacitances were much higher
which could explain supercapacitive performances of TiO2 -Ti3 C2 was than those of Ti3 C2 and TiO2 electrodes at the same rate (some sta-
better than that of single Ti3 C2 phase. tistical experiment data of Ti3 C2 and TiO2 electrodes were placed in
Figs. S7 and S8). The maximum Cs of Ti3 C2 and TiO2 electrodes were
Electrochemical performance of the as-prepared Ti3 C2 , TiO2 , 93 F g−1 and 6.4 F g−1 at 5 mV s−1 , respectively. The Cs values of
and TiO2 -Ti3 C2 electrode.—Figure 6a showed the cyclic voltammet- Ti3 C2 and TiO2 electrodes were less than that of TiO2 -Ti3 C2 electrode.
ric (CV) curves of the TiO2 -Ti3 C2 , Ti3 C2 and TiO2 electrode at a scan Electrochemical impedance spectroscopy (EIS) was also per-
rate of 20 mV s−1 . The three CV curves approximated symmetrical formed to investigate the internal resistance and capacity of the elec-
rectangular shapes. Ti3 C2 exhibited “true” pseudocapacitive behavior: trode material. Figure 6d revealed the Nyquist plots of Ti3 C2 and
it presented a continuous change in the titanium oxidation state during TiO2 -Ti3 C2 electrodes. The Nyquist plots for both types of electrodes
charge/discharge, producing rectangular-shaped CVs. And the pseu- consisted of a semicircle arc in the high-frequency region and a straight
docapacitive behavior of the nanocrystalline TiO2 particles14 was poor. line in the low-frequency region. In general, the semicircle arc in the
In comparison to the Ti3 C2 and TiO2 , the TiO2 -Ti3 C2 showed substan- high-frequency region was related to electronic resistance, and the
tially greater area and more symetrical shape, suggesting its higher vertical line in the low-frequency region indicated pure capacitive
electrochemical activity. The improvement of TiO2 -Ti3 C2 could be behavior.54 Equivalent series resistance (ESR) was determined by the
attributed to introduction of TiO2 for larger SSA (about 4 times) and interception of the high-frequency arc with the real axis.14 From the
more mesopores, and the formation of a little graphite carbon for inset(a) of Fig. 6d, ESR of TiO2 -Ti3 C2 and Ti3 C2 electrodes were
better electronic conductivity,11 synergistically. found to be 0.46 and 0.52 , respectively. The lower value indicates
Figure 6b and Fig. S6 displayed the CV curves of TiO2 -Ti3 C2 consistent interfacial contact between active materials and substrate.
electrode at scan rates from 5 to 200 mVs−1 . These results showed that The semicircle was associated with electrode surface properties and
the areas defined by the CV curves increased with scan rate increasing corresponded to the charge transfer resistance (Rct).14 Such small re-
and exhibited quasi-rectangular shape without obvious distortion even sistance of TiO2 -Ti3 C2 electrode indicated larger electroactive surface
at scan rates of up to 200 mV s−1 , indicating excellent capacitive area and lower charge transfer resistance than Ti3 C2 electrode. The
behavior and high rate capability. The excellent capacitance behavior inset (b) of Fig. 6d shows the Nyquist plot of the TiO2 electrode.
of TiO2 -Ti3 C2 composites was mainly provided by higher surface area Note that the TiO2 Nyquist plot did not contain a semicircle region.

Downloaded on 2016-02-23 to IP 113.200.58.77 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
A790 Journal of The Electrochemical Society, 163 (5) A785-A791 (2016)

Figure 6. (a) Cyclic voltammetric (CV)


curves of TiO2 -Ti3 C2 ,Ti3 C2 and TiO2 elec-
trode at a scan rate of 20 mV s−1 ; (b) CV
curves for TiO2 -Ti3 C2 electrode in the po-
tential window of −1 V to −0.35 V at vari-
ous scan rates (5–200 mV s−1 ); (c) The plots
of the Cs of TiO2 , Ti3 C2 , and TiO2 -Ti3 C2
nanocomposite electrodes as a function of
scan rate; (d) Nyquist plots of Ti3 C2 , TiO2 -
Ti3 C2 composites and TiO2 electrodes. All
tests were conducted in 6 M KOH electrolyte
solution.

This was probably due to the low faradaic resistances of the electrode
and the high electrical conductivity between TiO2 nanoparticles and
the current collector.54 These results demonstrate that the introduction
of TiO2 enhanced the electrochemical performance of the TiO2 -Ti3 C2
hybrid material, indicating a synergistic effect between TiO2 nanopar-
ticles and Ti3 C2 nanosheets.
Figure 7a shows the typical galvanostatic charge–discharge (GCD)
curves of the TiO2 -Ti3 C2 electrode for current densities of 0.5, 1, 1.5,
2, and 3 A g−1 . It can be seen that all of the curves were linear
and exhibited typical triangular shape, which indicated that the good
electrochemical capacitive behavior was achieved by the TiO2 -Ti3 C2
electrode. Based on the charge–discharge curve, the Cs of the electrode
can be calculated using18
Cs = I t/ (mV ) [2]
where I is the discharge current, t is the discharge time, m is the
mass of the electro-active material, and V is the potential window
(V). The calculated Cs values of the TiO2 -Ti3 C2 electrode were 39,
35, 33, 33, and 33 F g−1 at 0.5, 1, 1.5, 2, and 3 A g−1 , respectively.
These results are mainly consistent with the order indicated by the CV
curves.
Electrochemical stability was one of the most important factors
determining the usefulness of supercapacitors in commercial appli-
cations. Figure 7b shows the typical galvanostatic cycling curve of
TiO2 -Ti3 C2 electrode at current density of 1 A g−1 . The results sug-
gest that the TiO2 -Ti3 C2 electrode showed little capacitance losses
even after 6000 cycles (92% initial capacitance retention after 6000
cycles), confirming good cycle stability.
The performances of various electrode materials are listed in
Table I. It was clear that the performance of TiO2 -Ti3 C2 electrode in
this work was much higher than similar structure TiO2 -Ti2 C “paper”
in Ref. 14 and TiO2 /GO composites in Refs. 54 and 5. Meanwhile,
the performance of TiO2 -Ti3 C2 electrode was slightly lower than that
of d-Ti3 C2 /CNT in Ref. 12, the enhanced performance was mainly at-
tributed to higher SSA of d-Ti3 C2 and good conductivity provided by
introduction of CNT and absence of binder. In addition, electrochemi-
cal performance of d-Ti3 C2 “paper” in Ref. 18 was slightly higher than Figure 7. (a) The typical galvanostatic charge–discharge (GCD) curves of
that of the TiO2 -Ti3 C2 electrode, the improved performance was not the TiO2 -Ti3 C2 electrode for current densities of 0.5, 1, 1.5, 2 and 3 A g−1 ;
only owing to higher SSA of delaminated flakes and lack of binder, (b) the galvanostatic cycling curve of TiO2 -Ti3 C2 electrode (6000 GCD cy-
but also ascribed to smaller intercalating cation of H+ provided by cles), the inset of (b), the galvanostatic cycling data collected at 1 A g−1 .

Downloaded on 2016-02-23 to IP 113.200.58.77 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).
Journal of The Electrochemical Society, 163 (5) A785-A791 (2016) A791

Table I. Comparison of the performances of the various electrochemical supercapacitors.

Electrode material Binder mass Electrolyte Capacitance (Cs/scan rate) Stability (retention/cycle number) Ref.
TiO2 -Ti3 C2 5 wt.% KOH 143 F g−1 /5
mV s−1 96%/3000 In this work
TiO2 -Ti2 C 5 wt.% KOH ∼5 F g−1 /5 mV s−1 86%/6000 14
d-Ti3 C2 a 0 H2 SO4 150 F g−1 /5 mV s−1 NRb /10000 18
d-Ti3 C2 /CNT 0 MgSO4 150 F g−1 /2 mV s−1 NR/10000 12
TiO2 nanorod/GO NR Na2 SO4 100 F g−1 /5 mV s−1 80%/3000 54
rGO/TiO2 nanorod/rGO NR Na2 SO4 114.5 F g−1 /5 mV s−1 85%/3000 5

a d- Ti C -delaminated Ti3 C2 .12,18


3 2
b NR-not reported.12,18

1 M H2 SO4 electrolyte solution. Therefore, the electrochemical 18. M. Ghidiu, M. R. Lukatskaya, M.-Q. Zhao, Y. Gogotsi, and M. W. Barsoum, Nature,
performances of TiO2 -Ti3 C2 nanocomposites could be highly en- 516, 78 (2014).
19. F. Wang, C. Yang, C. Duan, D. Xiao, Y. Tang, and J. Zhu, J. Electrochem. Soc., 162,
hanced by depositing TiO2 nanoparticles on the Ti3 C2 nanosheets. B16 (2015).
20. J. Come, M. Naguib, P. Rozier, M. W. Barsoum, Y. Gogotsi, P. L. Taberna,
M. Morcrette, and P. Simon, J. Electrochem. Soc., 159, A1368 (2012).
Conclusions 21. D. Sun, M. Wang, Z. Li, G. Fan, L.-Z. Fan, and A. Zhou, Electrochem. Commun., 47,
80 (2014).
TiO2 -Ti3 C2 nanocomposite was synthesized through a simple syn- 22. M. Naguib, J. Come, B. Dyatkin, V. Presser, P.-L. Taberna, P. Simon, M. W. Barsoum,
thetic route and subsequently fabricated as an electrode material in and Y. Gogotsi, Electrochem. Commun., 16, 61 (2012).
electrochemical supercapacitors. The results indicat that the TiO2 - 23. Q. Tang, Z. Zhou, and P. W. Shen, J. Am. Chem. Soc., 134, 16909 (2012).
Ti3 C2 electrode exhibited high specific capacitance (143 F g−1 at 24. Y. Xie, M. Naguib, V. N. Mochalin, M. W. Barsoum, Y. Gogotsi, X. Yu, K.-W. Nam,
5 mV s−1 ), superior rate capability (82% of its initial capacitance re- X.-Q. Yang, A. I. Kolesnikov, and P. R. C. Kent, J. Am. Chem. Soc., 136, 6385 (2014).
25. F. Wang, C. H. Yang, M. Duan, Y. Tang, and J. F. Zhu, Biosens. Bioelectron., 74,
tention as the scan rate increased from 5 to 200 mV s−1 ), and excellent 1022 (2015).
cycling stability (92% initial capacitance retention after 6000 cycles) 26. H. Liu, C. Duan, C. Yang, W. Shen, F. Wang, and Z. Zhu, Sens. Actuators B: Chem.,
in 6 M KOH aqueous electrolyte. The incorporation of TiO2 nanoparti- 218, 60 (2015).
27. Q. Peng, J. Guo, Q. Zhang, J. Xiang, B. Liu, A. Zhou, R. Liu, and Y. Tian, J. Am.
cles into Ti3 C2 layers could significantly enhance the electrochemical Chem. Soc., 136, 4113 (2014).
performance by providing larger SSA, impeding the stacking of Ti3 C2 28. O. Mashtalir, M. Naguib, V. N. Mochalin, Y. Dall’Agnese, M. Heon, M. W. Barsoum,
and enlarging the distance between Ti3 C2 nanosheets for cation in- and Y. Gogotsi, Nat. Commun., 516, 78 (2013).
tercalation. More importantly, numerous TiO2 nanoparticles provided 29. Q. Qu, S. Yang, and X. Feng, Adv. Mater., 23, 5574 (2011).
30. S. Liu, J. Xie, C. Fang, G. Cao, T. Zhu, and X. Zhao, J. Mater. Chem., 22, 19738
additional diffusion paths for electrolyte ions. These results suggest (2012).
that TiO2 -Ti3 C2 electrode was a highly suitable, promising electrode 31. R. Lv, E. Cruz-Silva, and M. Terrones, ACS Nano, 8, 4061 (2014).
material for new generation high performance supercapacitors. 32. J. Zhu, H. Zhang, and N. A. Kotov, ACS Nano, 7, 4818 (2013).
33. C. Tang, Q. Zhang, M.-Q. Zhao, J.-Q. Huang, X.-B. Cheng, G.-L. Tian, H.-J. Peng,
and F. Wei, Adv. Mater., 26, 6100 (2014).
Acknowledgments 34. Y. Wu, T. Zhang, F. Zhang, Y. Wang, Y. Ma, Y. Huang, Y. Liu, and Y. Chen, Nano
Energy, 1, 820 (2012).
This work was supported by the National Natural Science Foun- 35. M.-Q. Zhao, Q. Zhang, J.-Q. Huang, G.-L. Tian, J.-Q. Nie, H.-J. Peng, and F. Wei,
dation of China (51572158), and the Graduate Innovation Fund of Nat. Commun., 5, 3410 (2014).
Shaanxi University of Science and Technology. 36. S. Yang, X. Feng, and K. Müllen, Adv. Mater., 23, 3575 (2011).
37. G. Zhou, S. Pei, L. Li, D.-W. Wang, S. Wang, K. Huang, L.-C. Yin, F. Li, and
H.-M. Cheng, Adv. Mater., 26, 625 (2014).
References 38. R. Chen, T. Zhao, J. Lu, F. Wu, L. Li, J. Chen, G. Tan, Y. Ye, and K. Amine, Nano
Lett., 13, 4642 (2013).
1. Q. Cheng, J. Tang, J. Ma, H. Zhang, N. Shinya, and L.-C. Qin, Carbon, 49, 2917 39. Y. Cao, X. Li, I. A. Aksay, J. Lemmon, Z. Nie, Z. Yang, and J. Liu, Phys Chem Chem
(2011). Phys, 13, 7660 (2011).
2. P. Simon and Y. Gogotsi, Nature Materials, 7, 845 (2008). 40. H. Kim, M.-Y. Cho, M.-H. Kim, K.-Y. Park, H. Gwon, Y. Lee, K. C. Roh, and
3. L. L. Zhang and X. S. Zhao, Chem. Soc. Rev., 38, 2520 (2009). K. Kang, Adv. Energy. Mater., 3, 1500 (2013).
4. Q. Zhang, E. Uchaker, S. L. Candelaria, and G. Cao, Chem. Soc. Rev., 42, 3127 41. R. Liu, W. Guo, B. Sun, J. Pang, M. Pei, and G. Zhou, Electrochimica Acta, 156, 274
(2013). (2015).
5. A. Ramadoss, G.-S. Kim, and S. J. Kim, CrystEngComm, 15, 10222 (2013). 42. A. Ramadoss and S. J. Kim, Carbon, 63, 434 (2013).
6. X. Wang, S. Kajiyama, H. Iinuma, E. Hosono, S. Oro, I. Moriguchi, M. Okubo, and 43. H. Wang, C. Guan, X. Wang, and H. J. Fan, Small, 11, 1470 (2015).
A. Yamada, Nat. Commun., 6, 6544 (2015). 44. M. Sawczak, M. Sobaszek, K. Siuzdak, J. Ryl, R. Bogdanowicz, K. Darowicki,
7. M. R. Lukatskaya, O. Mashtalir, C. E. Ren, Y. Dall’Agnese, P. Rozier, P. L. Taberna, M. Gazda, and A. Cenian, J. Electrochem. Soc., 162, A2085 (2015).
M. Naguib, P. Simon, M. W. Barsoum, and Y. Gogotsi, Science, 341, 1502 45. L. T. Yan, Y. Xu, M. Zhou, G. Chen, S. G. Deng, S. Smirnov, H. M. Luo, and
(2013). G. F. Zou, Electrochim. Acta, 169, 73 (2015).
8. G. Lota, K. Fic, and E. Frackowiak, Energy Environ. Sci., 4, 1592 (2011). 46. M. Naguib, O. Mashtalir, M. R. Lukatskaya, B. Dyatkin, C. Zhang, V. Presser,
9. W. Wei, X. Cui, W. Chen, and D. G. Ivey, Chem. Soc. Rev., 40, 1697 (2011). Y. Gogotsi, and M. W. Barsoum, Chem Commun, 50, 7420 (2014).
10. G. Wang, L. Zhang, and J. Zhang, Chem. Soc. Rev., 41, 797 (2012). 47. Y. Gao, L. Wang, A. Zhou, Z. Li, J. Chen, H. Bala, Q. Hu, and X. Cao, Mater. Lett.,
11. M. R. Lukatskaya, S.-M. Bak, X. Yu, X.-Q. Yang, M. W. Barsoum, and Y. Gogotsi, 150, 62 (2015).
Adv. Energy. Mater., 5 (2015). 48. X. Li, X. Wang, L. Zhang, S. Lee, and H. Dai, Science, 319, 1229 (2008).
12. M. Zhao, C. E. Ren, Z. Ling, M. R. Lukatskaya, C. Zhang, K. L. Van Aken, 49. Y. Dall’Agnese, M. R. Lukatskaya, K. M. Cook, P.-L. Taberna, Y. Gogotsi, and
M. W. Barsoum, and Y. Gogotsi, Adv. Mater., 27, 339 (2015). P. Simon, Electrochem. Commun., 48, 118 (2014).
13. P. Yan, R. Zhang, J. Jia, C. Wu, A. Zhou, J. Xu, and X. Zhang, J. Power Sources, 284, 50. V. Swamy, A. Kuznetsov, L. S. Dubrovinsky, R. A. Caruso, D. G. Shchukin, and
38 (2015). B. C. Muddle, Phys. Rev. B, 71, 184302 (2005).
14. R. B. Rakhi, B. Ahmed, M. N. Hedhili, D. H. Anjum, and H. N. Alshareef, Chem. 51. S. Brunauer, P. H. Emmett, and E. Teller, J. Am. Chem. Soc., 60, 309 (1938).
Mater., 27, 5314 (2015). 52. O. Wilhelmsson, J. P. Palmquist, E. Lewin, J. Emmerlich, P. Eklund, P. O. Å. Persson,
15. M. Naguib, V. N. Mochalin, M. W. Barsoum, and Y. Gogotsi, Adv. Mater., 26, 992 H. Högberg, S. Li, R. Ahuja, O. Eriksson, L. Hultman, and U. Jansson, J. Cryst.
(2014). Growth, 291, 290 (2006).
16. M. Naguib, O. Mashtalir, J. Carle, V. Presser, J. Lu, L. Hultman, Y. Gogotsi, and 53. S. Myhra, J. A. A. Crossley, and M. W. Barsoum, J. Phys. Chem. Solids, 62, 811
M. W. Barsoum, ACS Nano, 6, 1322 (2012). (2001).
17. M. Naguib, M. Kurtoglu, V. Presser, J. Lu, J. Niu, M. Heon, L. Hultman, Y. Gogotsi, 54. R. R. Liu, W. J. Guo, B. Sun, J. L. Pang, M. S. Pei, and G. W. Zhou, Electrochim.
and M. W. Barsoum, Adv. Mater., 23, 4248 (2011). Acta, 156, 274 (2015).

Downloaded on 2016-02-23 to IP 113.200.58.77 address. Redistribution subject to ECS terms of use (see ecsdl.org/site/terms_use) unless CC License in place (see abstract).

You might also like