Next Article in Journal
Characterization of Active Edible Films based on Citral Essential Oil, Alginate and Pectin
Next Article in Special Issue
The Synthesis Process and Thermal Stability of V2C MXene
Previous Article in Journal
Grain Refinement of a Powder Nickel-Base Superalloy Using Hot Deformation and Slow-Cooling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Chemical Stability of Ti3C2 MXene with Al in the Temperature Range 500–700 °C

Centre of Materials Science and Engineering, School of Mechanical and Electronic Control Engineering, Beijing Jiaotong University, Beijing 100044, China
*
Author to whom correspondence should be addressed.
Materials 2018, 11(10), 1979; https://doi.org/10.3390/ma11101979
Submission received: 21 September 2018 / Revised: 10 October 2018 / Accepted: 12 October 2018 / Published: 15 October 2018
(This article belongs to the Special Issue MAX Phases and MXenes: Synthesis and Applications)

Abstract

:
Ti3C2Tx MXene, a new 2D nanosheet material, is expected to be an attractive reinforcement of metal matrix composites because its surfaces are terminated with Ti and/or functional groups of –OH, –O, and –F which improve its wettability with metals. Thus, new Ti3C2Tx/Al composites with strong interfaces and novel properties are desired. To prepare such composites, the chemical stability of Ti3C2Tx with Al at high temperatures should be investigated. This work first reports on the chemical stability of Ti3C2Tx MXene with Al in the temperature range 500–700 °C. Ti3C2Tx is thermally stable with Al at temperatures below 700 °C, but it reacts with Al to form Al3Ti and TiC at temperatures above 700 °C. The chemical stability and microstructure of the Ti3C2Tx/Al samples were investigated by differential scanning calorimeter, X-ray diffraction analysis, scanning electron microscopy, and transmission electron microscopy.

Graphical Abstract

1. Introduction

MXenes, as a new family of two-dimensional (2D) materials, have attracted much attention because of their unusual combination of mechanical, physical, and chemical properties [1,2,3,4,5,6,7,8,9,10]. The MXenes are produced by the selective extraction of A-element from the layered MAX phases (where M is an early transition metal, A is an A group element, and X is C or N), such as Ti3AlC2, Ti2AlC, and Ti3SiC2 by either HF solutions [1,2,11,12,13], HCl/fluoride salt solutions [3,14,15,16,17,18], or HF/oxidant solutions [19]. After removal of the A layers from MAX, the top and bottom surfaces of the bare MXene monolayer or single sheet are terminated by the redistribution of M atoms with metallic bonds. Hence, the MXenes exhibit metallic conductivity. However, the exposed M surfaces of MXenes are always attached by functional groups such as hydroxyl (–OH), oxide (–O), or fluorine (–F) after etching in acid solutions. Hence, the general formula Mn+1XnTx (where Tx denotes the MXene terminal groups and x is number of terminal groups, n = 1, 2, or 3) is always used to represent the MXenes.
MXenes have similar structure and properties to graphene and other 2D nanosheets. The attractive 2D MXenes are now being used in a wide range of applications in energy storage, electromagnetic interference shielding, water purification, gas- and biosensors, lubrication, photo-, electro- and chemical catalysis, and as reinforcement for composites [3,5,6,7,12,20,21,22,23,24,25,26,27,28,29,30].
In the 2D MXene group, Ti3C2Tx is the first reported and the most explored material because it is easily synthesized from the Ti3AlC2 precursor using HF solution [1]. The 2D Ti3C2Tx single nanosheet is micrometers in width but less than a nanometer in thickness. Hence, the Ti3C2Tx MXene, with its attractive properties and high aspect ratios, has already been used as a reinforcement in composites. So far, the majority of the current work on composites containing Ti3C2Tx has focused on polymer matrix composites [30,31,32,33,34], but less attention has been devoted to the use of Ti3C2Tx to reinforce metal matrix composites.
In recent years, there has been intense research into Al matrix composites reinforced with 1D carbon nanotubes (CNTs) and 2D graphene to improve their mechanical properties [35,36,37,38]. However, there remains a challenge to overcome weak interfaces in the CNT/Al and graphene/Al composites, due to the poor wettability of CNTs and graphene with Al matrix.
Using Ti3C2Tx to reinforce the Al metal matrix composites possibly endows the Ti3C2Tx/Al composites with a strong interfacial strength because the surfaces of Ti3C2Tx nanosheets are terminated with Ti or functional groups such as –OH, –O, and –F. It has been reported that the surface modification of CNTs by –O and –OH functional groups enhanced the interfacial bonding between metals and CNTs [39,40]. The functionalized 2D nanotubes can be employed as reinforcing fillers [41,42]. In addition, the –F and –O surface functional groups on the Ti3C2Tx nanosheets can be eliminated at high temperatures [43,44]. The loss of surface functional groups possibly promotes the bonding of Ti surface atoms on bare Ti3C2 with Al. However, work on the Ti3C2Tx/Al composites has been much less focused so far. The ability to practically produce Ti3C2Tx/Al composites remains to be demonstrated. Al matrix composites reinforced with 1D CNTs and 2D graphene have always been prepared in the temperature range 500–640 °C. To practically prepare Ti3C2Tx/Al composites and to understand the chemical stability of Ti3C2Tx with Al, the temperature range 500–700 °C was selected in the present study.
The purpose of this work was to assess the chemical stability of Ti3C2Tx with Al in the temperature range from 500 °C to 700 °C by X-ray diffraction, differential scanning calorimetry, scanning electron microscopy, and transmission electron microscopy methods. The interfaces between Ti3C2Tx and Al were characterized.

2. Materials and Methods

Ti3C2Tx MXene was prepared by acid etching of Ti3AlC2 powder. The fabrication of Ti3AlC2 has been described elsewhere [45]. Briefly, Ti (325-mesh, >99.2% purity), Al (particle size <5 μm, 99.5% purity), and C (graphite, <45 µm, >99.5% purity) powders with a molar ratio of Ti:Al:C = 3:1.1:2 were mixed for 10 h. The mixed mixture was cold-pressed to form compacts with a diameter of 50 mm and a height of approximately 5 mm. The compacts were pressurelessly sintered at 1450 °C for 1 h in an Ar atmosphere. The sintered samples were pulverized and then sifted with a 300-mesh sieve to make Ti3AlC2 powder. A total of 2.5 g of Ti3AlC2 powder was immersed in 60 mL of 40% HF solution in a polytetrafluoroethylene (PTFE) container. A heating magnetic stirrer was used to continuously stir the solution in the PTFE container at 50 °C for 0.5 h. The solution was centrifugally separated in a centrifuge (LD-4, Jinan Wohong Experimental instrument Co., Ltd., Jinan, China) with a rotation rate of 4000 rpm for 5 min and then washed with deionized water until a pH of about 7 was attained. The obtained sediment was vacuum dried at 80 °C for 24 h to make the desired Ti3C2Tx powder. The prepared Ti3C2Tx powder was sieved with a 200-mesh sieve.
Al and 10 wt % Ti3C2Tx powders were mixed in a polypropylene container with agate balls for 10 h in a rotary drum type ball-miller with a speed of 150 rpm. The mixture was cold-pressed in a stainless steel mold with 100 MPa to form pellets with a diameter of 20 mm and a height of about 5 mm. The pellets were put into a graphite crucible coated with boron nitride and then sintered in the absence of additional pressure conditions in the temperature range 500–700 °C for 1 h in Ar.
The phase composition of the samples before and after sintering was identified by X-ray diffraction (XRD) analysis using a D/Max 2200 PC diffractometer (Rigaku Co. Ltd., Tokyo, Japan) applying monochromatic Cu Kα radiation. The operating voltage and current were 40 kV and 20 mA, respectively. The microstructures of the sintered samples were characterized with a ZEISS EVO 18 scanning electron microscope (SEM) equipped with an energy-dispersive spectrometer system (EDS), and a JEM-2100F (JEOL Ltd., Tokyo, Japan) transmission electron microscope (TEM) with an operating voltage of 200 kV. TEM images were processed with the software RADIUS Desktop 2.0 (EMSIS GmbH, Muenster, Germany). Differential scanning calorimetry (DSC) analysis was used to measure the amount of energy absorbed or released by the Ti3C2Tx/Al mixture heated in a NETZSCH STA 449F3 thermal analyzer (Netzsch, Germany) from room temperature to 800 °C with a heating rate of 15 °C/min in flowing Ar.

3. Results and Discussion

The DSC curve of the Ti3C2Tx/Al mixture, together with that of pure Al for comparison, is presented in Figure 1. A broad endothermic peak appearing at around 110 °C on the DSC curve for Ti3C2Tx/Al is ascribed to the evaporation of water absorbed in the interlayers of Ti3C2Tx. The thermal gravity (TG) curve correspondingly exhibits a downward trend on mass loss. A sharp endothermic peak at 664 °C corresponds to the melting of Al (660 °C). Comparing the two DSC curves, it can be found that there is an exothermic peak at 717 °C on the DSC curve of Ti3C2Tx/Al, which can be ascribed to the occurrence of a reaction between Ti3C2Tx and Al. In the temperature range from 150 °C to 700 °C, the TG curve reveals a gradual mass loss, possibly due to the removal of surface groups such as –OH and –O. Zhou et al. [43] reported that there was a broad and weak exothermic peak in the DSC curve for the pure Ti3C2Tx in the temperature range 200–800 °C in Ar due to the loss of surface groups. Shah et al. [46] reported that the bonding energies for –O, –OH, and –F groups in the Ti3C2Tx MXene are 530.3 eV, 531.3 eV, and 685.1 eV, respectively. The bonding energies suggest that the removal sequence of the surface groups is –O > –OH > –F. Li and coworkers [44] reported that the disappearance of –OH and –O groups occurred at approximately 500 K and the removal of –F group was required at a high temperature of 1173 K in Ar. Sang and coworkers also confirmed that the functional group of –O was removed after annealing of Ti3C2Tx MXene at 500 °C [47].
To identify the reaction products and to characterize the microstructure, the samples sintered at temperatures from 500 °C to 700 °C were examined using XRD, SEM, and TEM.
Figure 2 shows XRD patterns of 10 wt % Ti3C2Tx/Al samples before and after sintering at various temperatures. Before sintering, only Ti3C2Tx MXene and Al are detected, see Figure 2a. The broad peaks indicate the nanosheet structure of Ti3C2Tx. After sintering at 650 °C, some sharp diffraction peaks but with lower intensities appeared, in addition to the peaks corresponding to Al. These peaks are in good agreement with the simulated peaks for pure Ti3C2 and Ti3C2F2, see Figure 2b. It should be noted that the (002) peak at 8.7°, (004) peak at 17.98°, and (006) peak at 27.38° belonging to the Ti3C2Tx MXene disappeared in the XRD patterns of the sintered samples, see Figure 2b,c. The XRD results suggest that water molecules associated with the interlayer functional groups such as –O and –OH were gradually removed from 100 °C to 650 °C. The loss of water and functional groups causes a decrease in the basal space distance, or, alternatively, the newly formed surfaces possibly bond to form a new structure, which potentially induces the disappearance of basal peaks. The disappearance of basal plane peaks is similar to those for some layered compounds such as montmorillonite [48] and tetratitanate [49]. However, the diffractogram of the sample sintered at 700 °C shows the appearance of new phases of Al3Ti and TiC, indicating the occurrence of a reaction between Ti3C2Tx and Al at 700 °C. This result is in good agreement with the DSC analysis. Some weak peaks belonging to unreacted Ti3C2 and Ti3C2F2 can still be found. However, several peaks appearing in Figure 2c were not confirmed.
Figure 3 presents the morphologies of the raw Al and Ti3C2Tx powders and microstructures of the sintered samples. The raw Al particles are spherical, and each Ti3C2Tx particle has stacked multilayers, as shown in Figure 3a. After sintering at 650 °C, the polished surface shows that the Ti3C2Tx particles mainly distribute at the Al grain boundaries, see Figure 3b. No other new phases were detected in the sample sintered at 650 °C. The fracture surface clearly shows the stacked multilayers of Ti3C2Tx particles, see Figure 3c.
Figure 4a presents a TEM image of a multilayer Ti3C2Tx flake. The EDS line analysis, see Figure 4b, reveals the Al signal in the Ti3C2Tx flake, indicating the diffusion of Al atoms into the multilayers of Ti3C2Tx. The interlayer spacing areas of Ti3C2Tx provide active sites for Al nucleation and crystal growth. It should be noted that the F element was detected in the Ti3C2Tx multilayers, but O was mainly distributed in the areas between the Ti3C2Tx flake and Al grains. The presence of F in the multilayers suggests that the surface-layer termination of MXene is mostly F, i.e., Ti3C2F2. The –O group in Ti3C2Tx may react with Al to form Al2O3.
Figure 5 presents the TEM micrographs of a Ti3C2Tx/Al sample sintered at 650 °C. These Ti3C2Tx flakes are from several nanometers to tens of nanometers in thickness, indicating the thinner flake is composed of at least two Ti3C2Tx layers, see Figure 5a. The interspaces in the Ti3C2Tx flakes are filled with Al. A high-resolution TEM (HRTEM) image shows that the interlayer spacing of Ti3C2Tx is about 0.855 nm, see Figure 5b. This value is smaller than the value of 1.17 to 1.28 nm for the initial Ti3C2Tx flakes due to the removal of functional groups after heat treatment at 650 °C. The HRTEM image of the Ti3C2/Al interface reveals that the lattices of the Ti3C2 and Al regions are in direct contact, as shown in Figure 5c. The interface is clean and continuous. Neither precipitates nor amorphous regions are observed at the interface. The above observation suggests that the Ti3C2/Al interface is chemically and structurally stable at 650 °C.

4. Conclusions

The chemical stability of Ti3C2 with Al was investigated in the temperature range 500–700 °C in Ar for 1 h. Ti3C2Tx is thermally stable with Al at 650 °C, but it reacts with Al to form Al3Ti and TiC at 700 °C. For the Ti3C2Tx/Al sample sintered at 650 °C, the Ti3C2Tx/Al interface is clean and continuous, without precipitates and amorphous regions. The loss of water molecules and the removal of functional groups of –OH and –O induces a decrease in the interlayer spacing of Ti3C2Tx. The present work demonstrates the possibility to prepare Ti3C2Tx MXene reinforced Al or other metal matrix composites under certain processing conditions.

Author Contributions

Conceptualization, S.L. and Y.Z.; methodology, S.L.; formal analysis, J.Z. and S.H.; investigation, J.Z. and S.H.; writing—original draft preparation, J.Z.; writing—review and editing, S.L.; supervision, S.L.; project administration, S.L.; funding acquisition, S.L.

Funding

This work was funded by the National Natural Science Foundation of China, grant number 51772020, Beijing Natural Science Foundation, grant number 2182058, and Beijing Government Funds for the Constructive Project of Central Universities.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon, M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. Two-dimensional nanocrystals produced by exfoliation of Ti3AlC2. Adv. Mater. 2011, 23, 4248–4253. [Google Scholar] [CrossRef] [PubMed]
  2. Mashtalir, O.; Naguib, M.; Mochalin, V.N.; Dall’Agnese, Y.; Heon, M.; Barsoum, M.W.; Gogotsi, Y. Intercalation and delamination of layered carbides and carbonitrides. Nat. Commun. 2013, 4, 1716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Ghidiu, M.; Lukatskaya, M.R.; Zhao, M.Q.; Gogotsi, Y.; Barsoum, M.W. Conductive two-dimensional titanium carbide ‘Clay’ with high volumetric capacitance. Nature 2014, 516, 78–81. [Google Scholar] [CrossRef] [PubMed]
  4. Naguib, M.; Mochalin, V.N.; Barsoum, M.W.; Gogotsi, Y. MXenes: A new family of two-dimensional materials. Adv. Mater. 2014, 26, 992–1005. [Google Scholar] [CrossRef] [PubMed]
  5. Sun, S.; Liao, C.; Hafez, A.M.; Zhu, H.; Wu, S. Two-dimensional MXenes for energy storage. Chem. Eng. J. 2018, 338, 27–45. [Google Scholar] [CrossRef]
  6. Sun, Y.; Chen, D.; Liang, Z. Two-dimensional MXenes for energy storage and conversion applications. Mater. Today Energy 2017, 5, 22–36. [Google Scholar] [CrossRef]
  7. Ran, J.; Gao, G.; Li, F.T.; Ma, T.Y.; Du, A.; Qiao, S.Z. Ti3C2 MXene co-catalyst on metal sulfide photo-absorbers for enhanced visible-light photocatalytic hydrogen production. Nat. Commun. 2017, 8, 13907. [Google Scholar] [CrossRef] [PubMed]
  8. Ng, V.M.H.; Huang, H.; Zhou, K.; Lee, P.S.; Que, W.; Xu, J.Z.; Kong, L.B. Recent progress in layered transition metal carbides and/or nitrides (MXenes) and their composites: Synthesis and applications. J. Mater. Chem. A 2017, 5, 3039–3068. [Google Scholar]
  9. Ling, Z.; Ren, C.E.; Zhao, M.Q.; Yang, J.; Giammarco, J.M.; Qiu, J.; Barsoum, M.W.; Gogotsi, Y. Flexible and conductive MXene films and nanocomposites with high capacitance. Proc. Natl. Acad. Sci. USA 2014, 111, 16676–16681. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  10. Miranda, A.; Halim, J.; Barsoum, M.W.; Lorke, A. Electronic properties of freestanding Ti3C2Tx MXene monolayers. Appl. Phys. Lett. 2016, 108, 033102. [Google Scholar] [CrossRef]
  11. Anasori, B.; Xie, Y.; Beidaghi, M.; Lu, J.; Hosler, B.C.; Hultman, L.; Kent, P.R.; Gogotsi, Y.; Barsoum, M.W. Two-dimensional ordered, double transition metals carbides (MXenes). ACS Nano 2015, 9, 9507–9516. [Google Scholar] [CrossRef] [PubMed]
  12. Naguib, M.; Halim, J.; Lu, J.; Cook, K.M.; Hultman, L.; Gogotsi, Y.; Barsoum, M.W. New two-dimensional niobium and vanadium carbides as promising materials for Li-Ion batteries. J. Am. Chem. Soc. 2013, 135, 15966–15969. [Google Scholar] [CrossRef] [PubMed]
  13. Ghidiu, M.; Naguib, M.; Shi, C.; Mashtalir, O.; Pan, L.M.; Zhang, B.; Yang, J.; Gogotsi, Y.; Billinge, S.J.; Barsoum, M.W. Synthesis and characterization of two-dimensional Nb4C3 (MXene). Chem. Commun. 2014, 50, 9517–9520. [Google Scholar] [CrossRef] [PubMed]
  14. Halim, J.; Kota, S.; Lukatskaya, M.R.; Naguib, M.; Zhao, M.Q.; Moon, E.J.; Pitock, J.; Nanda, J.; May, S.J.; Gogotsi, Y.; et al. Synthesis and characterization of 2D molybdenum carbide (MXene). Adv. Funct. Mater. 2016, 26, 3118–3127. [Google Scholar] [CrossRef]
  15. Liu, F.; Zhou, A.; Chen, J.; Jia, J.; Zhou, W.; Wang, L.; Hu, Q. Preparation of Ti3C2 and Ti2C MXenes by fluoride salts etching and methane adsorptive properties. Appl. Surf. Sci. 2017, 416, 781–789. [Google Scholar] [CrossRef]
  16. Zhang, T.; Pan, L.; Tang, H.; Du, F.; Guo, Y.; Qiu, T.; Yang, J. Synthesis of two-dimensional Ti3C2Tx MXene using HCl+LiF etchant: Enhanced exfoliation and delamination. J. Alloy. Compd. 2017, 695, 818–826. [Google Scholar] [CrossRef]
  17. Naguib, M.; Unocic, R.R.; Armstrong, B.L.; Nanda, J. Large-scale delamination of multi-layers transition metal carbides and carbonitrides ‘MXenes’. Dalton Trans. 2015, 44, 9353–9358. [Google Scholar] [CrossRef] [PubMed]
  18. Halim, J.; Lukatskaya, M.R.; Cook, K.M.; Lu, J.; Smith, C.R.; May, S.J.; Hultman, L.; Gogotsi, Y.; Eklund, P.; Barsoum, M.W. Transparent conductive two-dimensional titanium carbide epitaxial thin films. Chem. Mater. 2014, 26, 2374–2381. [Google Scholar] [CrossRef] [PubMed]
  19. Alhabeb, M.; Maleski, K.; Mathis, T.S.; Sarycheva, A.; Hatter, C.B.; Uzun, S.; Levitt, A.; Gogotsi, Y. Selective etching of silicon from Ti3SiC2 (MAX) produces 2D titanium carbide (MXene). Angew. Chem. 2018. [Google Scholar] [CrossRef]
  20. Han, M.; Yin, X.; Wu, H.; Hou, Z.; Song, C.; Li, X.; Zhang, L.; Cheng, L. Ti3C2 MXenes with modified surface for high-performance electromagnetic absorption and shielding in the X-Band. ACS Appl. Mater. Interfaces 2016, 8, 21011–21019. [Google Scholar] [CrossRef] [PubMed]
  21. Shahzad, F.; Alhabeb, M.; Hatter, C.B.; Anasori, B.; Hong, S.M.; Koo, C.M.; Gogotsi, Y. Electromagnetic interference shielding with 2D transition metal carbides (MXenes). Science 2016, 353, 1137–1140. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Zhang, Q.; Teng, J.; Zou, G.; Peng, Q.; Du, Q.; Jiao, T.; Xiang, J. Efficient phosphate sequestration for water purification by unique sandwich-like MXene/Magnetic iron oxide nanocomposites. Nanoscale 2016, 8, 7085–7093. [Google Scholar] [CrossRef] [PubMed]
  23. Chen, J.; Chen, K.; Tong, D.; Huang, Y.; Zhang, J.; Xue, J.; Huang, Q.; Chen, T. CO2 and temperature dual responsive “Smart” MXene phases. Chem. Commun. 2015, 51, 314–317. [Google Scholar] [CrossRef] [PubMed]
  24. Xiao, B.; Li, Y.C.; Yu, X.F.; Cheng, J. MXenes: Reusable materials for NH3 sensor or capturer by controlling the charge injection. Sens. Actuators B 2016, 235, 103–109. [Google Scholar] [CrossRef]
  25. Xu, B.; Zhu, M.; Zhang, W.; Zhen, X.; Pei, Z.; Xue, Q.; Zhi, C.; Shi, P. Ultrathin MXene-micropattern-based field-effect transistor for probing neural activity. Adv. Mater. 2016, 28, 3333–3339. [Google Scholar] [CrossRef] [PubMed]
  26. Yang, J.; Chen, B.; Song, H.; Tang, H.; Li, C. Synthesis, characterization, and tribological properties of two-dimensional Ti3C2. Cryst. Res. Technol. 2014, 49, 926–932. [Google Scholar] [CrossRef]
  27. Zhang, X.; Xue, M.; Yang, X.; Wang, Z.; Luo, G.; Huang, Z.; Sui, X.; Li, C. Preparation and tribological properties of Ti3C2(OH)2 nanosheets as additives in base oil. RSC Adv. 2015, 5, 2762–2767. [Google Scholar] [CrossRef]
  28. Ma, T.Y.; Cao, J.L.; Jaroniec, M.; Qiao, S.Z. Interacting carbon nitride and titanium carbide nanosheets for high-performance oxygen evolution. Angew. Chem. Int. Ed. 2015, 55, 1138–1142. [Google Scholar] [CrossRef] [PubMed]
  29. Peng, C.; Yang, X.; Li, Y.; Yu, H.; Wang, H.; Peng, F. Hybrids of two-dimensional Ti3C2 and TiO2 exposing {001} facets toward enhanced photocatalytic activity. ACS Appl. Mater. Interfaces 2016, 8, 6051–6060. [Google Scholar] [CrossRef] [PubMed]
  30. Boota, M.; Anasori, B.; Voigt, C.; Zhao, M.Q.; Barsoum, M.W.; Gogotsi, Y. Pseudocapacitive electrodes produced by oxidant–Free polymerization of pyrrole between the layers of 2D titanium carbide (MXene). Adv. Mater. 2016, 28, 1517–1522. [Google Scholar] [CrossRef] [PubMed]
  31. Ye, Q.; Chen, T. Exploring the potential of exfoliated ternary ultrathin Ti4AlN3 nanosheets for fabricating hybrid patterned polymer brushes. RSC Adv. 2015, 5, 70339–70344. [Google Scholar] [CrossRef]
  32. Zhang, H.; Wang, L.; Chen, Q.; Li, P.; Zhou, A.; Cao, X.; Hu, Q. Preparation, Mechanical and antifriction performance of MXene/polymer composites. Mater. Des. 2016, 92, 682–689. [Google Scholar] [CrossRef]
  33. Liu, Y.H.; Zhang, J.K.; Zhang, X.; Li, Y.F.; Wang, J.T. Ti3C2Tx filler effect on the proton conduction property of polymer electrolyte membrane. ACS Appl. Mater. Interfaces 2016, 8, 20352–20363. [Google Scholar] [CrossRef] [PubMed]
  34. Cao, Y.; Deng, Q.; Liu, Z.; Shen, D.; Wang, T.; Huang, Q.; Du, S.; Jiang, N.; Lin, C.T.; Yu, J. Enhanced thermal properties of poly (vinylidenefluoride) composites with ultrathin nanosheets of MXene. RSC Adv. 2017, 7, 20494–20501. [Google Scholar] [CrossRef]
  35. Kuzumaki, T.; Miyazawa, K.; Ichinose, H.; Ito, K. Processing of carbon nanotube reinforced aluminum composite. J. Mater. Res. 1998, 13, 2445–2449. [Google Scholar] [CrossRef]
  36. He, C.; Zhao, N.; Shi, C.; Du, X.; Li, J.; Li, H.; Cui, Q. An approach to obtaining homogeneously dispersed carbon nanotubes in Al powders for preparing reinforced Al-Matrix composites. Adv. Mater. 2007, 19, 1128–1132. [Google Scholar] [CrossRef]
  37. Tjong, S.C. Recent progress in the development and properties of novel metal matrix nanocomposites reinforced with carbon nanotubes and graphene nanosheets. Mater. Sci. Eng. R 2013, 74, 281–350. [Google Scholar] [CrossRef]
  38. Li, Z.; Guo, Q.; Li, Z.; Fan, G.; Xiong, D.B.; Su, Y.; Zhang, J.; Zhang, D. Enhanced mechanical properties of graphene (reduced graphene oxide)/aluminum composites with a bioinspired nanolaminated structure. Nano Lett. 2015, 15, 8077–8083. [Google Scholar] [CrossRef] [PubMed]
  39. Park, M.; Kim, B.H.; Kim, S.; Han, D.S.; Kim, G.; Lee, K.R. Improved binding between copper and carbon nanotubes in a composite using oxygen-containing functional groups. Carbon 2011, 49, 811–818. [Google Scholar] [CrossRef]
  40. Lee, H.J.; Choi, W.S.; Nguyen, T.; Lee, Y.B.; Lee, H. An easy method for direct metal coordination reaction on unoxidized single-walled carbon nanotubes. Carbon 2011, 49, 5150–5157. [Google Scholar] [CrossRef]
  41. Cavallaro, G.; Danilushkina, A.A.; Evtugyn, V.G.; Lazzara, G.; Milioto, S.; Parisi, F.; Rozhima, E.V.; Fakhrullin, R.F. Halloysite nanotubes: Controlled access and release by smart gates. Nanomaterials 2017, 7, 199. [Google Scholar] [CrossRef] [PubMed]
  42. Lazzara, G.; Cavallaro, G.; Panchal, A.; Fakhrullin, R.; Stavitskaya, A.; Vinokurov, V.; Lvov, Y. An assembly of organic-inorganic composites using halloysite clay nanotubes. Curr. Opin. Colloid Interface Sci. 2018, 35, 42–50. [Google Scholar] [CrossRef]
  43. Li, Z.; Wang, L.; Sun, D.; Zhang, Y.; Liu, B.; Hu, Q.; Zhou, A. Synthesis and thermal stability of two-dimensional carbide MXene Ti3C2. Mater. Sci. Eng. B 2015, 191, 33–40. [Google Scholar] [CrossRef]
  44. Li, J.; Du, Y.; Huo, C.; Wang, S.; Cui, C. Thermal stability of two-dimensional Ti2C nanosheets. Ceram. Int. 2015, 41, 2631–2635. [Google Scholar] [CrossRef]
  45. Hu, S.; Li, S.B.; Li, H.; Zhou, Y. Large-scale growth of TiB2 hexagonal sheets. J. Alloy. Compd. 2017, 690, 930–935. [Google Scholar] [CrossRef]
  46. Shah, S.A.; Habib, T.; Gao, H.; Gao, P.; Sun, W.; Green, M.J.; Radovic, M. Template–Free 3D titanium carbide (Ti3C2Tx) MXene particles crumpled by capillary forces. Chem. Commun. 2017, 53, 400–403. [Google Scholar] [CrossRef] [PubMed]
  47. Sang, X.; Xie, Y.; Yilmaz, D.E.; Lotfi, R.; Alhabeb, M.; Ostadhossein, A.; Anasori, B.; Sun, W.; Li, X.; Xiao, K.; et al. In situ atomistic insight into the growth mechanisms of single layer 2D transition metal carbides. Nat. Comm. 2018, 9, 2266:1–2266:9. [Google Scholar] [CrossRef] [PubMed]
  48. Yang, J.; Ding, J. Preparation and characterization of alumina-pillared ultrafine layered tetratitanate. Mater. Lett. 2004, 58, 3872–3875. [Google Scholar] [CrossRef]
  49. Kooli, F.; Liu, Y.; Alshahattet, S.F.; Messali, M.; Bergaya, F. Reaction of acid activated montmorillonites with hexadecyl trimethylammonium bromide solution. Appl. Clay Sci. 2009, 43, 357–363. [Google Scholar] [CrossRef]
Figure 1. DSC analysis for 10 wt % Ti3C2/Al mixture in the temperature range 50–800 °C, together with Al for comparison.
Figure 1. DSC analysis for 10 wt % Ti3C2/Al mixture in the temperature range 50–800 °C, together with Al for comparison.
Materials 11 01979 g001
Figure 2. XRD patterns of the Ti3C2/Al mixture before (a), and after heat treatment at (b) 650 °C and (c) 700 °C in Ar for 1 h.
Figure 2. XRD patterns of the Ti3C2/Al mixture before (a), and after heat treatment at (b) 650 °C and (c) 700 °C in Ar for 1 h.
Materials 11 01979 g002
Figure 3. SEM micrographs of the initial powders (a), and polished surface (b) and fracture surface (c) of the samples sintered at 650 °C in Ar for 1 h. The left- and right-hand side micrographs in (a) show the morphologies of Al and Ti3C2Tx powders, respectively.
Figure 3. SEM micrographs of the initial powders (a), and polished surface (b) and fracture surface (c) of the samples sintered at 650 °C in Ar for 1 h. The left- and right-hand side micrographs in (a) show the morphologies of Al and Ti3C2Tx powders, respectively.
Materials 11 01979 g003aMaterials 11 01979 g003b
Figure 4. TEM image (a) and EDS results of the Ti3C2Tx/Al sample sintered at 650 °C (b). The EDS scan analysis along the line in (a) showing the elemental distribution of Al, Ti, C, O, and F.
Figure 4. TEM image (a) and EDS results of the Ti3C2Tx/Al sample sintered at 650 °C (b). The EDS scan analysis along the line in (a) showing the elemental distribution of Al, Ti, C, O, and F.
Materials 11 01979 g004
Figure 5. TEM images of the sample sintered at 650 °C. (a) A TEM micrograph showing stacked multilayers of Ti3C2Tx, (b) a high-resolution TEM (HRTEM) micrograph of Ti3C2Tx, and (c) a HRTEM micrograph of Ti3C2Tx/Al interface.
Figure 5. TEM images of the sample sintered at 650 °C. (a) A TEM micrograph showing stacked multilayers of Ti3C2Tx, (b) a high-resolution TEM (HRTEM) micrograph of Ti3C2Tx, and (c) a HRTEM micrograph of Ti3C2Tx/Al interface.
Materials 11 01979 g005

Share and Cite

MDPI and ACS Style

Zhang, J.; Li, S.; Hu, S.; Zhou, Y. Chemical Stability of Ti3C2 MXene with Al in the Temperature Range 500–700 °C. Materials 2018, 11, 1979. https://doi.org/10.3390/ma11101979

AMA Style

Zhang J, Li S, Hu S, Zhou Y. Chemical Stability of Ti3C2 MXene with Al in the Temperature Range 500–700 °C. Materials. 2018; 11(10):1979. https://doi.org/10.3390/ma11101979

Chicago/Turabian Style

Zhang, Jing, Shibo Li, Shujun Hu, and Yang Zhou. 2018. "Chemical Stability of Ti3C2 MXene with Al in the Temperature Range 500–700 °C" Materials 11, no. 10: 1979. https://doi.org/10.3390/ma11101979

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop