Cold Sintered Ceramic Nanocomposites of 2D MXene and Zinc Oxide
Cold Sintered Ceramic Nanocomposites of 2D MXene and Zinc Oxide
See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
Communication
ZnO–MXene Nanocomposites www.advmat.de
Adv. Mater. 2018, 30, 1801846 1801846 (1 of 6) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 32, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201801846 by Missouri University Of Science, Wiley Online Library on [25/10/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de
for both individual flakes as well as in the stacked films.[24,25] functionalization and without the need for high-energy, costly
MXenes have shown great promise for a variety of applica- processing conditions.
tions, including energy storage, electromagnetic interference The fabrication method of dense ZnO–Ti3C2Tx nanocompos-
shielding, sensors, water purifications, and medicine.[19] To ites is shown in Scheme 1b. To obtain a homogeneous disper-
date, MXenes have only been used as the reinforcement for sion, the prepared MXene solution was mixed and sonicated
polymer composites,[26–28] and no studies are available on with the ZnO, flash frozen, and then freeze-dried for 72 h.
MXene–metal or MXene–ceramic nanocomposites. Afterward, 17–20 wt% 1.5 M acetic acid was added into the
MXenes are ≈1 nm thick sheets of transition metal carbides, ZnO–Ti3C2Tx mixture and homogenized. Finally, the wetted
which are susceptible to oxidation in air. Furthermore, oxida- ZnO–Ti3C2Tx powders were pressed in a die under a pressure
tion has been shown to be accelerated at temperatures higher of 250 MPa at a temperature of 300 °C for 1 h. More details can
than 350 °C.[29,30] Therefore, traditional high temperature pro- be found in the Experimental Section.
cesses are not the best route for fabrication of MXene com- As shown in Figure 1a, the relative densities of the cold sin-
posites. However, this issue can be avoided through the CSP tered (1-y)ZnO–yTi3C2Tx composites (y = 0, 0.5, 1, 3, 5 wt%)
because of sintering temperatures ≤300 °C. Additionally, the are all higher than 90%,demonstrating that CSP is a promi
oxide-like surfaces of MXenes provide potential for compatible sing technique to sinter composite materials with MXenes.
integration in the oxide matrix of a ceramic, like ZnO. The X-ray diffraction (XRD) peaks of ZnO can be indexed to a
Among the cold sintered oxide ceramics, ZnO is an impor- wurtzite/zincite structure with the space group of P63mc, and
tant semiconducting material, which has the potential for a the MXene shows a broad (002) peak in the range of 8°–9° due
broad range of applications, such as varistors, thermoelectrics, to some multilayer flakes (Figure S1, Supporting Information),
optoelectronics, piezoelectric transducers, photocatalysts, and indicating that there is no reaction between ZnO and Ti3C2Tx
gas sensors.[31–33] ZnO is also a model system for the funda- under the cold sintering condition.
mental study of most sintering processes. Typically, ZnO is Figure 1b–e presents the SEM images of powders and
sintered to >90% of theoretical density at a temperature higher cold sintered samples of ZnO and 99ZnO–1Ti3C2Tx. In the
than 1000 °C with a holding time of several hours, using the case of cold sintered ZnO at 300 °C for 1 h, the grains grow
conventional thermal sintering.[34,35] By using the CSP, ZnO from nanometers (100–900 nm) to micrometers (1–4 µm)
has been shown to densify at temperatures ≤300 °C with (Figure 1b,d). In contrast, the grain growth of ZnO is sup-
a holding time of 1 h (total time of 2–3 h including heating, pressed in the cold sintered 99ZnO–1Ti3C2Tx composites
holding, and cooling).[14] The densification and grain growth (Figure 1c,e). The Ti3C2Tx located at the grain boundaries of
of the CSP sintered ZnO are controlled by the addition of an ZnO minimizes the coarsening of ZnO but does not prevent
acetic acid aqueous solution where the acetic acid dissolves the densification demonstrating that dense ZnO–Ti3C2Tx nano-
some of the exterior Zn ions from the surface of the powdered composites can be obtained by CSP. The cold sintered ZnO and
ceramic. This liquid phase coats each particle of the ceramic, 99ZnO–1Ti3C2Tx have highly compacted grain structures with a
assisting mass transport throughout each of the sintering steps, low porosity, which is in agreement with the measured density
including the rearrangement of particles, densification, and the data.
grain growth.[14] To further study the detailed microstructures of cold sintered
As a potential candidate in thermoelectric energy conversion, (1-y)ZnO–yTi3C2Tx nanocomposites, high-resolution TEM and
ZnO shows advantages of abundance, low cost, nontoxicity, STEM EDS mapping were employed, as shown in Figure 1f–l
and thermal stability.[36,37] However, the low electrical conduc- and Figures S2, S3 (Supporting Information). Figure 1f pre-
tivity limits its application.[37] The dispersion of 2D MXene sents the microstructural overview of the cold sintered 99ZnO–
along grain boundaries of ZnO, as shown in Scheme 1a, can 1Ti3C2Tx nanocomposite. The Ti3C2Tx nanosheets are shown
improve the electrical conductivity as well as the mechanical to be distributed around ZnO grains, which is further con-
properties. In this study, we report the cold co-sintering of firmed in the images with a higher magnification (Figure 1g,h;
ZnO with Ti3C2Tx in air to form ZnO–Ti3C2Tx nanocomposites, Figure S2, Supporting Information), where several Ti3C2Tx
demonstrating the feasibility of CSP as an effective fabrication nanosheets with thicknesses of a few nanometers can be
method to develop functional ceramic–matrix nanocomposites. found in the grain boundaries. Another proof of the Ti3C2Tx
We believe this is the first report of a 2D material being densi- distribution is shown in the STEM EDS mapping (Figure 1i–l;
fied into a ceramic without altering its structure or chemical Figure S3, Supporting Information), where a homogeneous
Scheme 1. Schematic illustration showing: a) the grain boundary of (1-y)ZnO–yTi3C2Tx nanocomposites and b) the fabrication process via cold sintering.
Adv. Mater. 2018, 30, 1801846 1801846 (2 of 6) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 32, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201801846 by Missouri University Of Science, Wiley Online Library on [25/10/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de
Figure 1. The density and microstructure of ZnO–Ti3C2Tx nanocomposites cold sintered at 300 °C for 1 h. a) The densities of cold sintered ZnO–Ti3C2Tx
nanocomposites. SEM images of: b) ZnO and c) 99ZnO–1Ti3C2Tx raw powders, and the cross sections of cold sintered: d) ZnO and e) 99ZnO–1Ti3C2Tx
ceramics. f–h) TEM, i) HAADF-STEM images and j–l) energy dispersive spectroscopy (EDS) elemental mapping of cold sintered 99ZnO–1Ti3C2Tx
nanocomposite. The red circles in TEM and HAADF-STEM images show one example of the Ti3C2Tx region. In the HAADF image, the bright areas
belong to ZnO and the dark areas belong to Ti3C2Tx. EDS maps, where elemental Zn is shown in red and Ti is shown in cyan, show the presence of
Ti3C2Tx at the ZnO grain boundaries.
dispersion of Ti3C2Tx is observed in the 99ZnO–1Ti3C2Tx nano- Figure 2a–c plots the temperature-dependent electrical
composites cold sintered at 300 °C. The sintering of pure ZnO conductivities, Seebeck coefficients, and power factors of
leads to formation of micron-size grains even at 300 °C. How- (1-y)ZnO–yTi3C2Tx nanocomposites cold sintered at 300 °C for
ever, in the case of ZnO–Ti3C2Tx composites, the 2D Ti3C2Tx 1 h. ZnO ceramics prepared by conventional thermal sintering
nanosheets located at the grain boundaries of ZnO inhibit in air at 1550 °C for 2 h,[34] 1400 °C for 10 h,[35] and 950 °C[38]
the final grain growth of ZnO, and thus, ZnO–Ti3C2Tx nano for 3 h have a conductivity of 0.17 S cm−1 at 250 °C, ≈0.2 S cm−1
composites can be obtained by CSP at 300 °C for 1 h. at ≈120 °C, and 10−3 to 10−2 S cm−1 at room temperature,
Figure 2. The temperature-dependent: a) electrical conductivities, b) Seebeck coefficients, and c) power factors (PF) of ZnO–Ti3C2Tx nanocomposites
cold sintered at 300 °C for 1 h. 300: Heat treatment in inert atmosphere at 300 °C. 750: Heat treatment in inert atmosphere at 750 °C. All figures share
the same legend.
Adv. Mater. 2018, 30, 1801846 1801846 (3 of 6) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 32, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201801846 by Missouri University Of Science, Wiley Online Library on [25/10/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de
Figure 3. The mechanical properties of ZnO–Ti3C2Tx nanocomposites cold sintered at 300 °C for 1 h: a) hardness; b) elastic modulus.
respectively. When sintered in vacuum at 950 °C under a pres- a metallic conductivity of 312 S cm−1, a Seebeck coefficient
sure of 100 MPa, ZnO shows a conductivity of 4–5 S cm−1 at of −120 µV K−1, and a power-factor of ≈4.5 × 10−4 W mK−2 at
100 °C.[39] In the case of sintering in Ar atmosphere at 900 °C 750 °C, which is promising for the thermoelectric applications,
under a pressure of 75 MPa, the reported conductivity of ZnO outperforming pure ZnO[43] (as seen in Figure S5, Supporting
reached 21.6 S cm−1 at room temperature; this is probably due Information) or MXenes[44] studied to date.
to oxygen vacancies or Zn interstitial.[37] As shown in Figure 2a Figure 3 presents the room temperature mechanical prop-
and Figure S4 (Supporting Information), our cold sintered pure erties of ZnO–Ti3C2Tx nanocomposites cold sintered at
ZnO shows an electrical conductivity of ≈0.08 S cm−1 at 100 °C 300 °C for 1 h. Polycrystalline ZnO ceramic prepared by con-
without further heat treatment and 2–5 S cm−1 with heat treat- ventional thermal sintering at 1300 °C shows a hardness of
ment, which is within the range reported in the literature. The 1.5–1.83 GPa.[45] Microwave sintering and hot pressing of ZnO
increase of electrical conductivity with heat treatment in inert produce hardness values of 0.49–1.72[46] and 2 GPa,[47] respec-
atmosphere may result from the change of the concentration of tively. With hot pressing in vacuum at 1200 °C, ZnO ceramic
dominant native donors in ZnO.[40–42] (7 wt% impurity, ZnAl2O4) has a higher hardness of ≈3.4 GPa.[48]
In the case of ZnO–Ti3C2Tx nanocomposites, most of The hardness of ZnO ceramic fabricated by CSP at 300 °C falls
Ti3C2Tx nanosheets are distributed around the grain bounda- within the range of other ZnO ceramics, but the hardness of
ries of ZnO, providing an efficient pathway for the electron ZnO–Ti3C2Tx nanocomposites increases dramatically as the
transport. Increasing the amount of 2D Ti3C2Tx improved the amount of Ti3C2Tx is increased (Figure 3a). The hardness can
electrical conductivity of the ZnO–Ti3C2Tx nanocomposites by be increased by 40–50% with 0.5 wt% Ti3C2Tx, and more than
1–2 orders of magnitude as plotted in Figure 2a and Figure S4 doubled with 5 wt% of Ti3C2Tx, which results from the micro-
(Supporting Information), confirming that the overall electrical structure of ZnO–Ti3C2Tx nanocomposites and the outstanding
conductivity is improved with the addition of Ti3C2Tx. The mechanical properties of Ti3C2Tx.[49] The elastic modulus of
95ZnO–5Ti3C2Tx nanocomposites cold sintered in air without cold sintered ZnO also falls in the range in the literature[33,50,51]
further heat treatment shows a conductivity of 24 S cm−1 at and was increased with the addition of Ti3C2Tx, as shown in
100 °C (ZnO: 0.08 S cm−1) whereas 99ZnO–1Ti3C2Tx heat- Figure 3b.
treated at 750 °C shows an increased conductivity of 544 S cm−1 In conclusion, this work introduces the CSP to densify
at 100 °C (ZnO: 5 S cm−1). This may result from the interface ZnO nanocomposites with Ti3C2Tx. The 2D Ti3C2Tx was evenly
and interdiffusion changing the defects and band bending with dispersed at the grain boundaries of the ZnO and minimized
heat treatment. the coarsening of ZnO under the CSP conditions. With the
Considering the temperature dependence, ZnO and addition of up to 5 wt% Ti3C2Tx, the electrical conductivity was
99.5ZnO–0.5Ti3C2Tx have a semiconducting behavior while increased by 1–2 orders of magnitude, and the overall power-
the ZnO–Ti3C2Tx nanocomposites with larger amounts of factor improved dramatically. The Seebeck coefficients of all
Ti3C2Tx have a metallic behavior, which results from the the ZnO–Ti3C2Tx nanocomposites were negative, indicating
metallic properties of Ti3C2Tx (Figure 2). The Seebeck coef- n-type charge carrier. The 99ZnO–1Ti3C2Tx nanocomposite
ficients of all the ZnO–Ti3C2Tx nanocomposites are negative, showed a metallic conductivity of 312 S cm−1, a Seebeck coef-
demonstrating that electrons are dominating charge carriers ficient of −120 µV K−1 and a power-factor of ≈4.5 × 10−4 W mK−2
(n-type). It has also been seen that the absolute value of See- at 750 °C. The hardness and elastic modulus of ZnO–Ti3C2Tx
beck coefficient of the ZnO–Ti3C2Tx nanocomposites slightly nanocomposites were enhanced dramatically, increasing by
decreases with increasing the amount of Ti3C2Tx; however, the 40–50% with only 0.5 wt% Ti3C2Tx and more than doubling
overall power-factor is improved dramatically. The 99.5ZnO– with 5 wt% of Ti3C2Tx. This investigation has demonstrated the
0.5Ti3C2Tx nanocomposite without further annealing has con- successful integration of a 2D material into an oxide ceramic,
ductivity, Seebeck coefficient, and power-factor of 16 S cm−1, suggesting that CSP is a promising new processing route
−270 µV K−1, and 1.2 × 10−4 W mK−2 at 300 °C, respectively. for electrically functional nanoceramics utilizing the innate
The 99ZnO−1Ti3C2Tx nanocomposite annealed at 750 °C shows properties of 2D materials.
Adv. Mater. 2018, 30, 1801846 1801846 (4 of 6) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 32, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201801846 by Missouri University Of Science, Wiley Online Library on [25/10/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de
Adv. Mater. 2018, 30, 1801846 1801846 (5 of 6) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim
15214095, 2018, 32, Downloaded from https://onlinelibrary.wiley.com/doi/10.1002/adma.201801846 by Missouri University Of Science, Wiley Online Library on [25/10/2022]. See the Terms and Conditions (https://onlinelibrary.wiley.com/terms-and-conditions) on Wiley Online Library for rules of use; OA articles are governed by the applicable Creative Commons License
www.advancedsciencenews.com www.advmat.de
[21] J. B. Pang, A. Bachmatiuk, Y. Yin, B. Trzebicka, L. Zhao, L. Fu, [36] W. H. Nam, Y. S. Lim, S. M. Choi, W. S. Seo, J. Y. Lee, J. Mater.
R. G. Mendes, T. Gemming, Z. F. Liu, M. H. Rummeli, Adv. Energy. Chem. 2012, 22, 14633.
Mater. 2018, 8, 1702093. [37] D. S. Chen, Y. Zhao, Y. N. Chen, B. A. Wang, H. Y. Chen, J. Zhou,
[22] Q. Hao, J. B. Pang, Y. Zhang, J. W. Wang, L. B. Ma, O. G. Schmidt, Z. Q. Liang, ACS Appl. Mater. Inter. 2015, 7, 3224.
Adv. Opt. Mater. 2018, 6, 1700984. [38] P. Jood, R. J. Mehta, Y. L. Zhang, G. Peleckis, X. L. Wang,
[23] M. Naguib, M. Kurtoglu, V. Presser, J. Lu, J. J. Niu, M. Heon, R. W. Siegel, T. Borca-Tasciuc, S. X. Dou, G. Ramanath, Nano Lett.
L. Hultman, Y. Gogotsi, M. W. Barsoum, Adv. Mater. 2011, 23, 4248. 2011, 11, 4337.
[24] A. Lipatov, M. Alhabeb, M. R. Lukatskaya, A. Boson, Y. Gogotsi, [39] K. F. Cai, E. Muller, C. Drasar, A. Mrotzek, Mat. Sci. Eng. B-Solid
A. Sinitskii, Adv. Electron. Mater. 2016, 2, 1600255. 2003, 104, 45.
[25] C. F. Zhang, B. Anasori, A. Seral-Ascaso, S. H. Park, N. McEvoy, [40] D. C. Look, D. C. Reynolds, J. R. Sizelove, R. L. Jones, C. W. Litton,
A. Shmeliov, G. S. Duesberg, J. N. Coleman, Y. Gogotsi, V. Nicolosi, G. Cantwell, W. C. Harsch, Solid State Commun. 1998, 105, 399.
Adv. Mater. 2017, 29, 1702678. [41] D. C. Look, J. W. Hemsky, J. R. Sizelove, Phys. Rev. Lett. 1999, 82,
[26] Z. Ling, C. E. Ren, M. Q. Zhao, J. Yang, J. M. Giammarco, J. S. Qiu, 2552.
M. W. Barsoum, Y. Gogotsi, P. Natl. Acad. Sci. USA 2014, 111, [42] C. G. Van de Walle, Phys. Rev. Lett. 2000, 85, 1012.
16676. [43] B. Zhu, D. Li, T. Zhang, Y. Luo, R. Donelson, T. Zhang, Y. Zheng,
[27] H. Zhang, L. B. Wang, Q. Chen, P. Li, A. G. Zhou, X. X. Cao, C. Du, L. Wei, H. Hng, Ceram. Int. 2018, 44, 6461.
Q. K. Hu, Mater. Design 2016, 92, 682. [44] H. Kim, B. Anasori, Y. Gogotsi, H. N. Alshareef, Chem. Mat. 2017,
[28] F. Shahzad, M. Alhabeb, C. B. Hatter, B. Anasori, S. M. Hong, 29, 6472.
C. M. Koo, Y. Gogotsi, Science 2016, 353, 1137. [45] T. K. Roy, Mater. Sci. Eng. A-Struct. Mater. Prop. Microstruct. Process.
[29] C. F. J. Zhang, S. Pinilla, N. McEyoy, C. P. Cullen, B. Anasori, 2015, 640, 267.
E. Long, S. H. Park, A. Seral-Ascaso, A. Shmeliov, D. Krishnan, [46] A. K. Mukhopadhyay, M. R. Chaudhuri, A. Seal, S. K. Dalui,
C. Morant, X. H. Liu, G. S. Duesberg, Y. Gogotsi, V. Nicolosi, Chem. M. Banerjee, K. K. Phani, Bull. Mat. Sci. 2001, 24, 125.
Mat. 2017, 29, 4848. [47] D. B. Marshall, T. Noma, A. G. Evans, J. Am. Ceram. Soc. 1982, 65,
[30] Z. Y. Li, L. B. Wang, D. D. Sun, Y. D. Zhang, B. Z. Liu, Q. K. Hu, C175.
A. G. Zhou, Mater. Sci. Eng. B-Adv. 2015, 191, 33. [48] H. Ruf, A. G. Evans, J. Am. Ceram. Soc. 1983, 66, 328.
[31] D. C. Look, Mat. Sci. Eng. B-Solid 2001, 80, 383. [49] V. N. Borysiuk, V. N. Mochalin, Y. Gogotsi, Comp. Mater. Sci. 2018,
[32] L. Vayssieres, Adv. Mater. 2003, 15, 464. 143, 418.
[33] U. Ozgur, Y. I. Alivov, C. Liu, A. Teke, M. A. Reshchikov, S. Dogan, [50] S. O. Kucheyev, J. E. Bradby, J. S. Williams, C. Jagadish, M. V. Swain,
V. Avrutin, S. J. Cho, H. Morkoc, J. Appl. Phys. 2005, 98, 041301. Appl. Phys. Lett. 2002, 80, 956.
[34] H. Ohta, W. S. Seo, K. Koumoto, J. Am. Ceram. Soc. 1996, 79, [51] S. Basu, M. W. Barsoum, J. Mater. Res. 2007, 22, 2470.
2193. [52] M. Alhabeb, K. Maleski, B. Anasori, P. Lelyukh, L. Clark, S. Sin,
[35] T. Tsubota, M. Ohtaki, K. Eguchi, H. Arai, J. Mater. Chem. 1997, 7, 85. Y. Gogotsi, Chem. Mat. 2017, 29, 7633.
Adv. Mater. 2018, 30, 1801846 1801846 (6 of 6) © 2018 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim