Surface Functional Groups and Interlayer Water Determine the

Apr 2, 2018 - Shenyang National Laboratory for Materials Science, Institute of Metal Research, Chinese Academy of Sciences, Shenyang 110016 , China ...
2 downloads 8 Views 2MB Size
Subscriber access provided by UNIV OF DURHAM

Surface Functional Groups and Interlayer Water Determine the Electrochemical Capacitance of Ti3C2Tx MXene Minmin Hu, Tao Hu, Zhaojin Li, Yi Yang, Renfei Cheng, Jinxing Yang, Cong Cui, and Xiaohui Wang ACS Nano, Just Accepted Manuscript • DOI: 10.1021/acsnano.8b00676 • Publication Date (Web): 02 Apr 2018 Downloaded from http://pubs.acs.org on April 2, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Surface Functional Groups and Interlayer Water Determine the Electrochemical Capacitance of Ti3C2Tx MXene ‡ †‡ Minmin Hu,†,§ Tao Hu,†, Zhaojin Li, , Yi Yang,# Renfei Cheng,†,§ Jinxing Yang,†,§ Cong Cui†,§

and Xiaohui Wang*,† †

Shenyang National Laboratory for Materials Science, Institute of Metal Research, Chinese

Academy of Sciences, Shenyang 110016, China §

School of Materials Science and Engineering, University of Science and Technology of

China, Shenyang 110016, China ‡

#

University of Chinese Academy of Sciences, Beijing 100049, China

Suzhou Niumag Analytical Instrument Corporation, Suzhou 215163, China

KEYWORDS: MXene, two-dimensional materials, functional groups, interlayer water, supercapacitor

ABSTRACT: MXenes, an emerging class of conductive two-dimensional materials, have been regarded as promising candidates in the field of electrochemical energy storage. The electrochemical performance of their representative, Ti3C2Tx, where T stands for the surface termination group of F, O, or OH, strongly relies on termination mediated surface functionalization, but an in-depth understanding of the relationship between them remains 1 ACS Paragon Plus Environment

ACS Nano 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

unresolved. Here, we studied comprehensively the structural feature and electrochemical performance of two kinds of Ti3C2Tx MXenes obtained by etching Ti3AlC2 precursor in aqueous HF solution with low concentration (6 mol/L) and high concentration of (15 mol/L), respectively. A significantly higher capacitance was recognized in low concentration HF-etched MXene (Ti3C2Tx–6M) electrode. In situ Raman spectroscopy and X-ray photoelectron spectroscopy demonstrate that Ti3C2Tx–6M has more component of –O functional group. In combination with X-ray diffraction analysis, low field 1H nuclear magnetic resonance spectroscopy in terms of relaxation time unambiguously underlines that Ti3C2Tx–6M is capable of accommodating more high-mobility H2O molecules between the Ti3C2Tx interlayers, enabling more hydrogen ions more readily accessible to the active sites of Ti3C2Tx–6M. The two main key factors, i.e., high content of –O functional groups that are involved bonding/debonding-induced pseudocapacitance and more high-mobility water intercalated between the MXene interlayers simultaneously account for the superior capacitance of Ti3C2Tx–6M electrode. This study provides a guideline for the rational design and construction of high-capacitance MXene and MXene-based hybrid electrodes in aqueous electrolytes.

The growth of energy demands and more attention to global environmental issues call for urgent exploration for clean energies and advanced energy storage systems.1,2 Electrochemical capacitors, known as supercapacitors, are the ideal candidate among the efficient energy storage methodologies, owing to the high power density, excellent rate performance and long cycle life, and so on.3,4 Pseudocapacitors that can provide higher energy densities than electrical double-layer capacitors making use of reversible and fast 2 ACS Paragon Plus Environment

Page 2 of 33

Page 3 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

surface redox reactions have attracted much attention.5–7 The most studied materials for pseudocapacitors are RuO2,8 MnO2,9 Nb2O5,10 NiO11 and so on,12 but the high electrode resistance resulting from the limited electronic conductivity of most pseudocapacitive oxides always leads to the lower rate performance.13 Therefore, it’s urgent to explore promising active materials with excellent electrochemical performance. Recently, MXenes, an emerging two dimensional (2D) materials group of early transition metal carbides and/or carbonitrides, have proved to be a pseudocapacitive electrode material with high capacitance.14–23 MXenes can be prepared by selectively removing the A element from MAX phases, a family of layered solids with a chemical formula of Mn + 1AXn, where M is a transition metal, A is an A-group element, and X is C and/or N.24,25 After etching of the A layers, the MX layers left in the etchants are spontaneously terminated with O, F or OH groups, giving a general formula Mn

+ 1XnTx,

where Tx stands for a general surface

termination.26 Ti3C2Tx, is the most intensively studied to date among the discovered MXenes. The Ti3C2Tx interlayers have interactions like graphite or MoS227 and is intrinsically of electronic conductivity.28 Ti3C2Tx MXene stores charge through pseudocapacitance induced by surface functional groups-involved bonding/debonding in H2SO4 electrolyte.17,18,29 Termination plays important roles in this charge storage process. Many efforts have been devoted to tune the surface groups to modulate the electrochemical properties of MXene. For example, MXenes, after cation intercalation30,31 and/or annealing treatment,32 get an improvement in the electrochemical performance, which is largely attributed to the decrease of the content of –F groups. So far, an in-depth and comprehensive research about the relationship between structural feature and fundamental electrochemical properties of Ti3C2Tx 3 ACS Paragon Plus Environment

ACS Nano 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 33

MXene remains lacking, which critically impedes the process of exploring MXene and MXene-based hybrid electrodes for electrochemical applications. In order to figure out the correlation between termination and electrochemical performance, we should firstly obtain the MXenes with different termination species and proportions. It is worth noting that the termination species and proportions are dependent on the treatment conditions.33–37 Lower HF concentrations always result in a larger –O groups to –F groups ratio.33 NMR spectra of Ti3C2Tx MXene demonstrated that HF synthesised material has almost four times as much –F termination as the LiF–HCl synthesised material.36 Importantly, the capacitive performance is highly depended on the synthesis methods by means of selectively etching off Al atomistic layers from the laminated Ti3AlC2 precursor (Table S1). It is thus of scientific importance to make an insight into the huge discrepency in electrochemical capacitance. Importantly, etching is the first step to prepare MXenes electrodes. Once the relationship between the structure and electrochemical performance is confirmed, one can directly optimize the etching environment to achieve higher capacitance instead of conducting further subsequent treatments such as annealing in a variety of environments32 or intercalation.30,31,38 Herein, we comprehensively studied the correlation between the structure and electrochemical performance of two kinds of Ti3C2Tx MXenes obtained by etching Ti3AlC2 precursor in two aqueous HF solutions with different concentrations (6 mol/L and 15 mol/L), followed by ultrasonication. Even the acidic etchant and the sonication method are identical, the capacitance recorded in the 6 mol/L HF-etched Ti3C2Tx MXene (denoted as Ti3C2Tx–6M) electrode is nearly 2 times that of the 15 mol/L HF-etched Ti3C2Tx MXene (Ti3C2Tx–15M) 4 ACS Paragon Plus Environment

Page 5 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

electrode. Such huge difference in capacitance is attributed to the distinction in –O functional groups content confirmed by in situ Raman spectroscopy and X-ray photoelectron spectroscopy (XPS). Besides, more H2O molecule between the Ti3C2Tx layers facilitates more hydrogen ions accessible to the active sites of Ti3C2Tx–6M, giving a dramatic increase in gravimetric performance. RESULTS AND DISCUSSION

The preparation of Ti3C2Tx–6M and Ti3C2Tx–15M followed the method described in the experimental section. X-ray diffraction (XRD) patterns demonstrated the extraction of Al from the laminated Ti3AlC2 precursor for the two samples investigated in this study (Figure S1). In order to figure out whether the two etching conditions would affect the electrochemical performance of the samples, we carefully examined their capacitive performance of the MXenes electrodes prepared from Ti3C2Tx–6M and Ti3C2Tx–15M. The fabrication of Ti3C2Tx electrodes followed the method reported previously (Figure S2).18,39 Figure 1a plots the cyclic voltammetry (CV) curves of the Ti3C2Tx–6M and Ti3C2Tx–15M electrodes in H2SO4 electrolyte. The more rectangular and symmetric shape of the CV curve indicates a more superior capacitive performance of the Ti3C2Tx–6M electrode. Additionally, a much higher capacitance was obtained in the Ti3C2Tx–6M electrode than the Ti3C2Tx–15M electrode. The difference in capacitance between the two electrodes is as high as 192 F/g (Figure 1b). First, we examined the specific surface areas of the two samples. As shown in Figure 2, the Brunauer–Emmett–Teller (BET) specific surface areas of the Ti3C2Tx particulates obtained in 5 ACS Paragon Plus Environment

ACS Nano 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 33

6 mol/L HF and 15 mol/L HF are determined to be 3.5 and 40.5 m2/g by nitrogen adsorption, respectively. Notably, the accordion-like MXene particulates with more microscopically visible slits can adsorb more N2 gas. For the purpose of film electrode preparation, the MXene particulates were delaminated by ultrasonic treatment. With the ultrasonication, the specific surface area of Ti3C2Tx–15M (25.6 m2/g) is still lager than that of Ti3C2Tx–6M (16.2 m2/g). Large specific surface area typically leads to large capacitances for carbon-based materials in which electrical double layer capacitance dominates the capacitances.19 In contrast, Ti3C2Tx–6M with smaller BET specific surface area exhibited much higher capacitance. Thus, pseudocapacitance predominates in capacitance for Ti3C2Tx–6M with smaller BET specific surface area yet much higher capacitance, which is relevant to the surface functional groups on the MXene. In order to figure out whether the large difference in electrochemical performance comes from the surface functional groups, we conducted the in situ electrochemical Raman spectroscopy measurements to monitor the electrochemical processes of Ti3C2Tx–6M and Ti3C2Tx–15M electrodes in H2SO4. Herein a two-electrode open device was adopted for the measurements. Figure 3 shows that Raman band on the negatively charged electrode varies orderly and reversibly with the change of voltage in H2SO4 electrolyte. To observe the change of Raman peaks explicitly and directly, corresponding Lorentzian peak-fitting of representative bands in the range of 530–770 cm–1 is performed through PeakFit software (Figure 4). Based on the result previously reported,18,40 also as summarized in Table S2, the mode (590 and 726 cm–1) comes from the vibrations of atoms in Ti3C2O2, the band at 708 cm– 1

is assigned to out-of-plane vibrations of C atoms in Ti3C2O(OH), and the bands at 630 and 6 ACS Paragon Plus Environment

Page 7 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

672 cm–1 belong to the vibrations of atoms in Ti3C2(OH)2. Prior to applying voltage (at 0 V), the Raman peak intensity of Ti3C2O2 for Ti3C2Tx–6M electrode is higher than that of Ti3C2Tx– 15M electrode, which implies that –O terminations in Ti3C2Tx–6M are more than those in Ti3C2Tx–15M. When scanned from 0 V to –0.4 V, whether for Ti3C2Tx–6M electrode or for Ti3C2Tx–15M electrode, the peaks at 590 and 726 cm−1 in Ti3C2O2 gradually weaken, whereas the mode at 708 cm−1 of Ti3C2O(OH) strengthens, and the mode (630 and 672 cm–1) in Ti3C2(OH)2 almost keeps intact, which indicates that both electrodes all undergo the transformation from Ti3C2O2 to Ti3C2O(OH) upon applying voltage. An obvious distinction is that this transformation is more dramatic in Ti3C2Tx–6M, which results from that Ti3C2Tx– 6M possesses more –O terminations that involved in bonding/debonding and induced pseudocapacitance. For the bands in the range of 150–300 cm–1, similar mode evolution trend was observed (Figure S3). Hence, the increase of capacitance for Ti3C2Tx–6M electrode is attributed to the higher –O group content in a large extent. To shed light on the relationship between surface functional groups and capacitances quantitatively, we characterized the two samples by XPS in terms of surface chemical composition and chemical state.41 Among the three functional groups of –F, –OH, and –O, the functional groups of –O are involved in electrochemical reaction to store energy according to the above discussed and previously reported18 in situ electrochemical Raman spectroscopy analysis. So, we focus on the discrepancy of the proportion of –O groups between Ti3C2Tx–6M and Ti3C2Tx–15M. To avoid the influence of any contamination on the surface of samples from ambient environment, we recorded the XPS spectra of Ti3C2Tx–6M and Ti3C2Tx–15M after sputtering (Figure S4). Figure 5 shows the deconvolution of Ti 2p, O 7 ACS Paragon Plus Environment

ACS Nano 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 33

1s XPS spectra for Ti3C2Tx–6M and Ti3C2Tx–15M, respectively. The Ti 2p spectra are deconvolved with three doublets (Ti 2p 2p3/2–Ti 2p1/2) with a set area ratio equal to 2:1. The confirmation of peak position of moieties is shown in Figure S5. According to the Ti 2p spectra and O 1s spectra of both samples, the proportion of the moieties containing –O group in the whole spectrum for Ti3C2Tx–6M is all higher than those for Ti3C2Tx–15M (Table 1), which is consistent with the result of Raman spectroscopy analysis (Figure 4). Combining the O1s spectra with the percentage of O atom in the Ti3C2Tx unit, we can obtain the ratio of –O group in the Ti3C2Tx unit (Table S3). The proportion determined from C–Ti–O in O1s spectrum for Ti3C2Tx–6M is 49%, which is 6% higher than that for Ti3C2Tx–15M. Additionally, one Ti3C2Tx unit possesses 1.9 O atoms for Ti3C2Tx–6M while one Ti3C2Tx unit has 1.5 O atoms for Ti3C2Tx–15M. Based on the above results, we can deduce that the quantity of –O groups in one Ti3C2Tx unit is 0.93 for Ti3C2Tx–6M and that is 0.65 for Ti3C2Tx–15M. The difference in –O functional groups content between them is 0.28. According to the result of in situ electrochemical Raman spectroscopy measurements, the electrochemical reaction upon discharging follows the equation 1

1

2

2

൫M–Ox ൯ + x e− + x H+ → M−O1x (OH)1x 2

(1)

2

In the case of x is 0.28, the specific capacitance value is 167 F/g. It means that Ti3C2Tx–6M electrode would give rise to the capacitance of 167 F/g more than Ti3C2Tx–15M electrode in H2SO4 solution. The value is very close to the difference in specific capacitances experimentally measured herein (192 F/g). In the above discussion, we do not take into account the –F functional group due to its low content and instable chemical nature. Once Ti3C2Tx MXene is all ideally terminated with –O functional group (wherein x is 2), after 8 ACS Paragon Plus Environment

Page 9 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

discharging, an ultrahigh capacitance of 1190 F/g is predicted. By means of sophiscatedly tuning the surface chemistry of Ti3C2Tx MXene, this huge theoretical capacitance is reasonably expected. In addition to the difference in –O functional groups content, there exists another siginificant difference of proportion of H2O molecule between the Ti3C2Tx–6M and Ti3C2Tx– 15M interlayers, as shown in Figure 5c,d and Table 1. We know that the incorporation of confined fluid molecules into the layered and 2D materials interlayer can improve ion transport and electrochemical activity.42 To further shed light on the reason of the difference in electrochemical performance between the two samples, it is necessary to distinguish the difference of the amount and state of confined water between the both MXene layers. The amount of water is sensitive to the degree of drying.43,44 Hence, the analysis of both MXenes at different moisture levels can help us to further understand the hydration effect between Ti3C2Tx layers. As shown in Figure 6a, with the increase of drying temperature, the weights all decrease above 3% whether for Ti3C2Tx–6M or Ti3C2Tx–15M. Generally, the drying below 200 °C does not lead to the change of the surface termination.36 In addition, the weight of MXene should remain fairly unchanged even though the surface functional groups are converted to one another owing to the close molar weights for O (16), OH (17), and F (19). Therefore, the weight decrease upon drying is reasonably ascribed to the evaporation of interlayer water. The loss of water always gives rise to the different interlayer spacing, which is confirmed by XRD. As shown in Figure 6b, when the samples dried at room temperature (RT), the distance between MXene layers of the Ti3C2Tx–6M is significantly larger than that of the Ti3C2Tx–15M, indicating that the unit sheets of Ti3C2Tx–6M MXene are more 9 ACS Paragon Plus Environment

ACS Nano 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 33

separated. This is consistent with the fact that etching in more concentrated HF (22.6 mol/L) gives rise to decreased interlayer spacing.45 For film electrode preparation, the Ti3C2Tx–6M and Ti3C2Tx–15M particulates were delaminated by ultrasonication. We note that this treatment however does not lead to the increase in the interlayer distance (Figure S1). One possibility for the increase of interlayer spacing comes from the difference in surface functional groups between both samples. This possibility can be excluded by the simulation of XRD patterns for Ti3C2Tx with different terminations (Figure S6). More –O terminations alone in Ti3C2Tx do not increase interlayer spacing. So, the increased interlayer spacing maybe originate from the presence of water intercalated between the MXene layers. Upon drying at elevated temperature, the (0002) reflections all significantly shift to similar high angles for both MXenes, indicating that the interlayer distances of two samples are comparable after drying at 120 °C. As discussed above, the drying below 200 °C does not change the surface termination and the amount of water is sensitive to the drying temperature. Consequently, the change in interlayer spacing is ascribed to the volatility of the water molecules intercalated between Ti3C2Tx interlayers. The interlayer spacing difference between room temperature drying and elevated temperature (120 °C) drying for Ti3C2Tx–6M is 3.2Å while that for Ti3C2Tx–15M is only 1.1 Å. This distinction for these two samples underlines that more water molecule is accommodated between the Ti3C2Tx–6M MXene interlayers. Note that the weight loss difference from ambient drying to drying at 120 °C for Ti3C2Tx–6M is however comparable to Ti3C2Tx–15M (4.8% for Ti3C2Tx–6M versus 3.6% for Ti3C2Tx–15M). This discrepancy probably results from that the microscopically visible slits as seen in Figure 2 accommodate an unnegligible amount of water for the Ti3C2Tx–15M. 10 ACS Paragon Plus Environment

Page 11 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Having firmly established the amount of water molecules intercalated between Ti3C2Tx layers, to gain an insight into the nature of water, we further characterized the state of water by means of proton nuclear magnetic resonance (NMR). It had been widely accepted that 1H NMR relaxation times are essential parameters to monitor the state of water in a variety of material systems.46,47 In this study, the 1H NMR time-domain spectra were recorded to distinguish the identities of H-containing species in Ti3C2Tx–6M and Ti3C2Tx–15M. Figure 6c depicts the continuous distribution of spin–spin relaxation time, T2, for the two samples. Generally, there are three recognizable states of water molecular in materials, i.e., bonded water, confined water, and free water, which are distinguishable by the time-domain parameter. 48,49 For Ti3C2Tx–6M, it shows a broad population in the time-domain region of 1.0–3.0 ms. A similar population in the range of 0.5–2 ms can be seen for Ti3C2Tx–15M. The populations in these ranges are less mobile, coming from –OH groups with variable degrees of hydrogen bonding. The two 1H spectra show a more mobile population in the region of 7– 200 ms, ascribed to H2O as their intensities are highly sensitive to the drying conditions (RT or 120 °C). The signal in the intermediate time-domain region of 7–200 ms is reasonably assigned to confined water (in the present case, it is referred as to interlayer water). Here we should note that the interlayer water in Ti3C2Tx–6M MXene with a time-domain range of 25– 200 ms is more easily relaxed than that in Ti3C2Tx–15M MXene whose time-domain range is in the range of 7–75 ms. This unambiguously demonstrates that it is easier to diffuse between MXene interlayers for hydrogen ion in Ti3C2Tx–6M than in Ti3C2Tx–15M. In order to figure out whether a larger width of interlayer slit in which more layers of water with high mobility intercalated can store more energy, we comprehensively investigated the 11 ACS Paragon Plus Environment

ACS Nano 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 33

electrochemical performance of Ti3C2Tx–6M electrode at different moisture levels. As shown in Figure S7, the electrode dried at low temperature dose exhibit a higher capacitance. We know that hydrogen ion transport is assisted by H2O molecule. More H2O lying between the Ti3C2Tx interlayers reasonably facilitates hydrogen ions accessible to the active sites of MXenes. As a result, an improvement in gravimetric performance for Ti3C2Tx–6M electrode was observed as its wide interlayer slit can accommodate more water molecule. The excellent capacitance of Ti3C2Tx–6M electrode is also understood from a kinetic point of view. In general, the capacitance of an electrode may come from the contribution of diffusion-limited process and that of surface capacitive effect-limited process. These effects can be characterized by analyzing the CV data. To shed light on the charge storage kinetics, we analyzed the relationship between current and scan rates. Assuming power-law dependence of the current i on sweep rate ν, gives rise to i = aν b

(2)

where a and b are adjustable values. In particular, a b-value of 0.5 is an indication of the diffusion-controlled process, while a value of 1 represents a surface capacitive storage.50,51 Herein for the Ti3C2Tx–6M electrode, the b-values are about 1.0 at the measured potential, demonstrating that the current comes primarily from the surface capacitive effects-controlled process, whereas the b-values for the Ti3C2Tx–15M electrode are around 0.8, which suggests a diffusion and surface capacitive effect jointly limited behavior as afore-discussed (Figure S8a). Due to the fact that water-richer interlayers minimize ion transport limitations, surface capacitive effect was dominated in Ti3C2Tx–6M electrode, which is beneficial for the rate performance. Electrochemical impedance spectroscopy (EIS) data also can support this 12 ACS Paragon Plus Environment

Page 13 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

argument. In Figure S8b, the impedance spectra include a high-frequency region with a semicircle arc and a low-frequency region with a straight line. The value of charge transfer resistance (Rct) determined from the diameter of the semicircle in the high frequency range52 of the Ti3C2Tx–6M electrode is lower than that of the Ti3C2Tx–15M electrode (Figure S9, Table S4 and Table S5). Generally, Rct is significantly negatively correlated with the electrolyte accessible area.32 The lower value of Rct for the Ti3C2Tx–6M electrode means it has the larger electroactive surface area to make hydronium ion contact with. Additionally, the nearly vertical line in the low-frequency region and a constant-phase element (CPE) with a fractional exponent α = 0.96 (Table S4) suggest that the Ti3C2Tx–6M electrode behaves more closely as an ideal capacitor.

CONCLUSION In summary, we have comprehensively compared the structure and electrochemical performance of two electrodes in which a much higher capacitance was recorded in Ti3C2Tx– 6M electrode than Ti3C2Tx–15M electrode. Using a combination of spectroscopic, structural and electrochemical characterization tools, we demonstrated that the huge difference in capacitance between the two Ti3C2Tx MXenes came from two aspects: First, the increased content of –O functional groups in the Ti3C2Tx–6M electrode; Second, the more H2O molecule intercalated between the Ti3C2Tx interlayers enables more hydrogen ions accessible to active sites of Ti3C2Tx–6M electrode, giving a dramatic increase in gravimetric performance. The knowledge established here not only reasonably elucidate the long-standing discrepancy in gravimetric capacitance for Ti3C2Tx MXene,13,16,23,30,31,38 but also more importantly offers a practical guideline for the further improvement of capacitance. 13 ACS Paragon Plus Environment

ACS Nano 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

For instance, one can tune the surface chemistry through etching in lower concentration of HF or other milder systhesis methods and/or expand the interlayer spacing by accommodating more H2O molecules between the MXene interlayers. The MXene with larger interspaces and larger amount of water not only can achieve the high capacitance, but also can behave excellent cycling performance.53

EXPERIMENTAL SECTION Synthesis of Ti3C2Tx MXene. The porous Ti3AlC2 monolith was prepared by means of solid-liquid reaction.54 In a typical synthesis, two pieces of the monolith with a total weight of roughly 2 gram were immersed in 10 mL of 6 and 15 mol/L HF solution (Sinopharm Chemical Reagent CO., Ltd, Shanghai, China) at RT for 72 h and 48 h, respectively. The resulting powders were vacuum filtrated and washed several times with deionized water. Then, deionized water was added into the as-separated wet sediment, followed by sonication in a pulse mode for 1 h, and centrifuged at 2,000 rpm for 30 min to remove the large particulates. After decantation, the black Ti3C2Tx MXenes colloidal supernatants were obtained for further use. Electrochemical Measurements. The electrode fabrication followed the dropping-mild baking method.18,39 The Ni foam was employed as current collector and the Ti3C2Tx MXene film uniformly coated on the skeleton of the foam as the active material. The mass loading of MXene is about 1.2 mg/cm2 and the thickness of the Ti3C2Tx film is statistically measured to be 1.0 ± 0.3 µm for both samples (Figure S2). All electrochemical measurements were performed in 3-electrode cells, where Ti3C2Tx–6M and Ti3C2Tx–15M served as working electrode respectively, platinum was used as counter electrode, and Ag/AgCl in saturated KCl 14 ACS Paragon Plus Environment

Page 14 of 33

Page 15 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

as a reference in order to precisely control electrochemical potentials. The electrolyte was 1 mol/L H2SO4. CV and EIS measurements were all performed using an electrochemical workstation (PARSTAT 2273, Princeton Applied Research). The CV curves were recorded using scan rates from 2 to 100 mV/s. EIS investigations were performed at 0 V with a 10 mV amplitude between 10 mHz and 100 kHz. Characterization of Ti3C2Tx MXenes. Microstructural morphology studies of Ti3C2Tx paticulates were conducted by scanning electron microscope (LEO Supra35, Zeiss). XRD patterns were collected on an powder diffractometer (D/max-2400, Rigaku) using Cu Kα radiation (λ=1.5406 Å) in the range 2θ = 5–80° with a step of 0.02°. BET specific surface areas were characterized with N2 as adsorption gas at 77 K on a physisorption analyzer (ASAP 2020, Micromeritics). The sample for BET specific surface area measurement was prepared by drying at 60 °C overnight. XPS was performed by a surface analysis system (ESCALAB250, Thermo VG company, USA) using monochromatic Al-K (1486.6 eV) radiation. The data was collected before and after Ar+ beam (3 kV) sputtering for 60 s. Component peak-fitting of XPS spectra was performed by using an XPS curve fitting software. NMR experiments were performed under 0.5 T applied magnetic field using a low field nuclear magnetic resonance spectrometer (MesoMR, Niumag Corporation, Shanghai, China). A corresponding resonance frequency for 1H of 23.311 MHz was used for measurement. The volume of the sample is around 2 mL. The spin-spin relaxation time, T2, was collected through Carr-Purcell-Meiboom-Gill (CPMG) sequences. The pulse width was 3 µs, sequence repetition time was 5s and the dwell time between data was 5 µs. The NMR

15 ACS Paragon Plus Environment

ACS Nano 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 33

data were analysed using the Niumag NMR inverse software. The in situ Raman spectroscopy measurements followed the previously reported method.18 Computation Method. The Cambridge Sequential Total Energy Package55 was used to carry out the density functional theory calculations. The calculations include a complete geometry optimization40 and a subsequent core level spectra calculation. Core level spectra were calculated based on the as-obtained ground-state structures. In the core level spectra calculation, the electronic exchange energy was treated as GGA-PBE.56 In the irreducible Brillouin zone, the Monkhorst-Pack scheme57 with 9 × 9 × 1 k points meshes were used and the individual spacing was less than 0.05 Å−1. On-the-fly generated (OTFG) pseudopotentials58 were used to calculate the core level spectra. The cutoff energy was set to 610 eV, the convergence for energy was assigned to be as 1.0 × 10−10 eV/atom and Fermi level was smeared by 0.1 eV.

ASSOCIATED CONTENT Supporting Information The Supporting Information is available free of charge on the ACS Publications website at DOI: Additional supporting figures and tables (PDF). AUTHOR INFORMATION Corresponding Author * Xiaohui Wang. E-mail: [email protected]. ORCID Xiaohui Wang: 0000-0001-7271-2662 16 ACS Paragon Plus Environment

Page 17 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

ACKNOWLEDGMENT

This work was supported by the Youth Innovation Promotion Association, Chinese Academy of Sciences (CAS) under grant No.2011152, and Shenyang National Laboratory for Materials Science, Institute of Metal Research, CAS, and by Special Program for Applied Research on Super Computation of the NSFC-Guangdong Joint Fund (the second phase) under Grant No. U1501501.

17 ACS Paragon Plus Environment

ACS Nano 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 33

Figure 1. Capacitive performances of Ti3C2Tx–6M and Ti3C2Tx–15M electrodes in 1 mol/L H2SO4 electrolyte. (a) CV curves collected at a scan rate of 20 mV/s. (b) Gravimetric capacitances at different scan rates. Note that the capacitance of Ti3C2Tx–6M is significantly greater than that of Ti3C2Tx–15M at each scan rate.

18 ACS Paragon Plus Environment

Page 19 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Figure 2. Nitrogen adsorption/desorption isotherms of (a,b) as-prepared Ti3C2Tx particulates and (c,d) those treated by ultrasonication. The Ti3C2Tx MXenes are prepared in (a,c) 6 mol/L HF and (b,d) 15 mol/L HF, respectively. Insets show their typical scanning electron microscopy images of as-prepared Ti3C2Tx particulates.

19 ACS Paragon Plus Environment

ACS Nano 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 33

Figure 3. In situ Raman spectra recorded on negative electrode of Ti3C2Tx–6M and Ti3C2Tx– 15M in the H2SO4 electrolyte. Note that the Raman bands are voltage-dependent in a reversible manner.

20 ACS Paragon Plus Environment

Page 21 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Figure 4. Selected Raman shift window and Lorentzian fits of the Raman bands centered at 590, 630, 672, 708 and 726 cm–1 for (a) Ti3C2Tx–6M and (b) Ti3C2Tx–15M. Note that the band at 726 cm–1 weakens while the mode at 708 cm–1 strengthens dramatically when the potential sweeps from 0 V to −0.4 V for the Ti3C2Tx–6M electrode compared with Ti3C2Tx– 15M.

21 ACS Paragon Plus Environment

ACS Nano 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 22 of 33

Figure 5. Component peak-fitting of XPS spectra of Ti3C2Tx–6M and Ti3C2Tx–15M in the (a,b) Ti 2p region and (c,d) O 1s region. The proportion of the moieties containing –O functional group in the Ti 2p spectrum or O 1s spectrum for Ti3C2Tx–6M is higher than those for Ti3C2Tx–15M.

22 ACS Paragon Plus Environment

Page 23 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Figure 6. (a) Normalized weight ( W/WRT, where WRT is the weight of the sample dried at 25 °C) of Ti3C2Tx–6M and Ti3C2Tx–15M dried at different temperatures. Insets show the schematics of Ti3C2Tx–6M and Ti3C2Tx–15M dried at RT. (b) (0002) peaks of XRD patterns show that the interlayer spacing shrinks upon drying at 120 °C both for Ti3C2Tx–6M and Ti3C2Tx–15M, whereas the shringkage extent is more obvious for Ti3C2Tx–6M. (c) 1H time-domain nuclear magnetic resonance spectra of as-synthesized Ti3C2Tx MXenes and those dried at 120 °C overnight.

23 ACS Paragon Plus Environment

ACS Nano 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 33

Table 1. Fractions of the fitted moieties in Ti 2p and O 1s XPS spectra for Ti3C2Tx–6M and Ti3C2Tx–15M. Ti 2p

O 1s

Fraction (%) Ti–C (MXene)

C–Ti–OH

C–Ti–O

Ti–O–H

Ti–O

H2 O

Ti3C2Tx–6M

31

26

43

33

49

18

Ti3C2Tx–15M

33

39

28

45

43

12

24 ACS Paragon Plus Environment

Page 25 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

REFERENCES (1) Gogotsi, Y.; Simon, P. True Performance Metrics in Electrochemical Energy Storage. Science 2011, 334, 917–918. (2) Yang, Z. G.; Zhang, J. L.; Kintner-Meyer, M. C. W.; Lu, X. C.; Choi, D.; Lemmon, J. P.; Liu, J. Electrochemical Energy Storage for Green Grid. Chem. Rev. 2011, 111, 3577–3613. (3) Simon, P.; Gogotsi, Y. Materials for Electrochemical Capacitors. Nat. Mater. 2008, 7, 845–854. (4) Zhai, Y. P.; Dou, Y. Q.; Zhao, D. Y.; Fulvio, P. F.; Mayes, R. T.; Dai, S. Carbon Materials for Chemical Capacitive Energy Storage. Adv. Mater. 2011, 23, 4828–4850. (5) Rauda, I. E.; Augustyn, V.; Dunn, B.; Tolbert, S. H. Enhancing Pseudocapacitive Charge Storage in Polymer Templated Mesoporous Materials. Accounts Chem. Res. 2013, 46, 1113– 1124. (6) Sassin, M. B.; Chervin, C. N.; Rolison, D. R.; Long, J. W. Redox Deposition of Nanoscale Metal Oxides on Carbon for Next-Generation Electrochemical Capacitors. Accounts Chem. Res. 2013, 46, 1062–1074. (7) Augustyn, V.; Simon, P.; Dunn, B. Pseudocapacitive Oxide Materials for High-Rate Electrochemical Energy Storage. Energy Environ. Sci. 2014, 7, 1597–1614. (8) Hu, C. C.; Chang, K. H.; Lin, M. C.; Wu, Y. T. Design and Tailoring of the Nanotubular Arrayed Architecture of Hydrous RuO2 for Next Generation Supercapacitors. Nano Lett. 2006, 6, 2690–2695. (9) Toupin, M.; Brousse, T.; Belanger, D. Charge Storage Mechanism of MnO2 Electrode Used in Aqueous Electrochemical Capacitor. Chem. Mat. 2004, 16, 3184–3190. 25 ACS Paragon Plus Environment

ACS Nano 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 33

(10) Augustyn, V.; Come, J.; Lowe, M. A.; Kim, J. W.; Taberna, P. L.; Tolbert, S. H.; Abruna, H. D.; Simon, P.; Dunn, B. High-Rate Electrochemical Energy Storage through Li+ Intercalation Pseudocapacitance. Nat. Mater. 2013, 12, 518–522. (11) Prasad, K. R.; Miura, N. Electrochemically Deposited Nanowhiskers of Nickel Oxide as a High-Power Pseudocapacitive Electrode. Appl. Phys. Lett. 2004, 85, 4199–4201. (12) Zhang, C.; Xing, J.; Fan, H. W.; Zhang, W. K.; Liao, M. Y.; Song, Y. Enlarged Capacitance of TiO2 Nanotube Array Electrodes Treated by Water Soaking. J. Mater. Sci. 2016, 52, 3146–3152. (13) Lukatskaya, M. R.; Kota, S.; Lin, Z. F.; Zhao, M. Q.; Shpigel, N.; Levi, M. D.; Halim, J.; Taberna, P. L.; Barsoum, M. W.; Simon, P.; Gogotsi, Y. Ultra-High-Rate Pseudocapacitive Energy Storage in Two-Dimensional Transition Metal Carbides. Nat. Energy 2017, 6, 17105. (14) Anasori, B.; Lukatskaya, M. R.; Gogotsi, Y. 2D Metal Carbides and Nitrides (MXenes) for Energy Storage. Nature Rev. Mat. 2017, 2, 16098. (15) Zhang, X.; Zhang, Z.; Zhou, Z. MXene-Based Materials for Electrochemical Energy Storage. J. Energy Chem. 2018, 27, 73–85. (16) 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. (17) Lukatskaya, M. R.; Bak, S. M.; Yu, X.; Yang, X. Q.; Barsoum, M. W.; Gogotsi, Y. Probing the Mechanism of High Capacitance in 2D Titanium Carbide Using In Situ X-Ray Absorption Spectroscopy. Adv. Energy Mater. 2015, 5, 1500589. 26 ACS Paragon Plus Environment

Page 27 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

(18) Hu, M. M.; Li, Z. J.; Hu, T.; Zhu, S. H.; Zhang, C.; Wang, X. H. High-Capacitance Mechanism for Ti3C2Tx MXene by In Situ Electrochemical Raman Spectroscopy Investigation. ACS Nano 2016, 10, 11344–11350. (19) Lukatskaya, M. R.; Mashtalir, O.; Ren, C. E.; Dall’Agnese, Y.; Rozier, P.; Taberna, P. L.; Naguib, M.; Simon, P.; Barsoum, M. W.; Gogotsi, Y. Cation Intercalation and High Volumetric Capacitance of Two-Dimensional Titanium Carbide. Science 2013, 341, 1502– 1505. (20) Wang, Y.; Dou, H.; Wang, J.; Ding, B.; Xu, Y. L.; Chang, Z.; Hao, X. D. Three-Dimensional Porous MXene/Layered Double Hydroxide Composite for High Performance Supercapacitors. J. Power Sources 2016, 327, 221–228. (21) Wen, Y. Y.; Rufford, T. E.; Chen, X. Z.; Li, N.; Lyu, M. Q.; Dai, L. M.; Wang, L. Z. Nitrogen-Doped Ti3C2Tx MXene Electrodes for High-Performance Supercapacitors. Nano Energy 2017, 38, 368–376. (22) Hu, M. M; Li, Z. J.; Li, G. X.; Hu, T.; Zhang, C.; Wang, X. H. All-Solid-State Flexible Fiber-Based MXene Supercapacitors. Adv. Mater. Technol. 2017, 2, 1700143. (23) Fu Q. S.; Wen J.; Zhang N.; Wu L. L.; Zhang M. Y.; Lin S. Y.; Gao H.; Zhang X. T. Free-Standing Ti3C2Tx Electrode with Ultrahigh Volumetric Capacitance. RSC Adv. 2017, 7, 11998. (24) 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.

27 ACS Paragon Plus Environment

ACS Nano 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 28 of 33

(25) Alhabeb, M.; Maleski, K.; Anasori, B.; Lelyukh, P.; Clark, L.; Sin, S.; Gogotsi, Y. Guidelines for Synthesis and Processing of Two-Dimensional Titanium Carbide (Ti3C2Tx MXene). Chem. Mat. 2017, 29, 7633–7644. (26) Naguib, M.; Mochalin, V. N.; Barsoum, M. W.; Gogotsi, Y. 25th Anniversary Article: MXenes: A New Family of Two-Dimensional Materials. Adv. Mater. 2014, 26, 992–1005. (27) Hu, T.; Hu, M. M.; Li, Z. J.; Zhang, H.; Zhang, C.; Wang, J. Y.; Wang, X. H. Interlayer Coupling in Two-Dimensional Titanium Carbide MXenes. Phys. Chem. Chem. Phys. 2016, 18, 20256–20260. (28) Hu, T.; Zhang, H.; Wang, J. Y.; Li, Z. J.; Hu, M. M.; Tan, J.; Hou, P. X.; Li, F.; Wang, X. H. Anisotropic Electronic Conduction in Stacked Two-Dimensional Titanium Carbide. Sci. Rep. 2015, 5, 16329. (29) Srimuk, P.; Kaasik, F.; Krüner, B.; Tolosa, A.; Fleischmann, S.; Jäckel, N.; Tekeli, M. C.; Aslan, M.; Suss, M. E.; Presser, V. MXene as a Novel Intercalation-Type Pseudocapacitive Cathode and Anode for Capacitive Deionization. J. Mater. Chem. A 2016, 4, 18265–18271. (30) Li, J.; Yuan, X. T.; Lin, C.; Yang, Y. Q.; Xu, L.; Du, X.; Xie, J. L.; Lin, J. H.; Sun, J. L. Achieving High Pseudocapacitance of 2D Titanium Carbide (MXene) by Cation Intercalation and Surface Modification. Adv. Energy Mater. 2017, 7, 1602725. (31) Dall'Agnese, Y.; Lukatskaya, M. R.; Cook, K. M.; Taberna, P. L.; Gogotsi, Y.; Simon, P. High Capacitance of Surface-Modified 2D Titanium Carbide in Acidic Electrolyte. Electrochem. Commun. 2014, 48, 118–122.

28 ACS Paragon Plus Environment

Page 29 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

(32) Rakhi, R. B.; Ahmed, B.; Hedhili, M. N.; Anjum, D. H.; Alshareef, H. N. Effect of Postetch Annealing Gas Composition on the Structural and Electrochemical Properties of Ti2CTx MXene Electrodes for Supercapacitor Applications. Chem. Mat. 2015, 27, 5314– 5323. (33) Wang, H. W.; Naguib, M.; Page, K.; Wesolowski, D. J.; Gogotsi, Y. Resolving the Structure of Ti3C2Tx MXenes through Multilevel Structural Modeling of the Atomic Pair Distribution Function. Chem. Mat. 2016, 28, 349–359. (34) Magne, D.; Mauchamp, V.; Celerier, S.; Chartier, P.; Cabioc'h, T. Site-Projected Electronic Structure of Two-Dimensional Ti3C2 MXene: The Role of the Surface Functionalization Groups. Phys. Chem. Chem. Phys. 2016, 18, 30946–30953. (35) Harris, K. J.; Bugnet, M.; Naguib, M.; Barsoum, M. W.; Goward, G. R. Direct Measurement of Surface Termination Groups and Their Connectivity in the 2D MXene V2CTx Using NMR Spectroscopy. J. Phys. Chem. C 2015, 119, 13713–13720. (36) Hope, M. A.; Forse, A. C.; Griffith, K. J.; Lukatskaya, M. R.; Ghidiu, M.; Gogotsi, Y.; Grey, C. P. NMR Reveals the Surface Functionalisation of Ti3C2 MXene. Phys. Chem. Chem. Phys. 2016, 18, 5099–5102. (37) Hu, T.; Li, Z. J.; Hu, M. M.; Wang, J. M.; Hu, Q. M.; Li, Q. Z.; Wang, X. H. Chemical Origin of Termination-Functionalized MXenes: Ti3C2T2 as a Case Study. J. Phys. Chem. C 2017, 121, 19254–19261. (38) Mashtalir, O.; Lukatskaya, M. R.; Kolesnikov, A. I.; Raymundo-Piñero, E.; Naguib, M.; Barsoum, M. W.; Gogotsi, Y. Effect of Hydrazine Intercalation on Structure and Capacitance of 2D Titanium Carbide (MXene). Nanoscale 2016, 8, 9128–9133. 29 ACS Paragon Plus Environment

ACS Nano 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 33

(39) Hu, M. M.; Li, Z. J.; Zhang, H.; Hu, T.; Zhang, C.; Wu, Z.; Wang, X. H. Self-Assembled Ti3C2Tx MXene Film with High Gravimetric Capacitance. Chem. Commun. 2015, 51, 13531– 13533. (40) Hu, T.; Wang, J. M.; Zhang, H.; Li, Z. J.; Hu, M. M.; Wang, X. H. Vibrational Properties of Ti3C2 and Ti3C2T2 (T = O, F, OH) Monosheets by First-Principles Calculations: A Comparative Study. Phys. Chem. Chem. Phys. 2015, 17, 9997–10003. (41) Halim, J.; Cook, K. M.; Naguib, M.; Eklund, P.; Gogotsi, Y.; Rosen, J.; Barsoum, M. W. X-Ray Photoelectron Spectroscopy of Select Multi-Layered Transition Metal Carbides (MXenes). Appl. Surf. Sci. 2016, 362, 406–417. (42) Augustyn, V.; Gogotsi, Y. 2D Materials with Nanoconfined Fluids for Electrochemical Energy Storage. Joule 2017, 1, 443–452. (43) Ren, C. E.; Hatzell, K. B.; Alhabeb, M.; Ling, Z.; Mahmoud, K. A.; Gogotsi, Y. Chargeand Size-Selective Ion Sieving through Ti3C2Tx MXene Membranes. J. Phys. Chem. Lett. 2015, 6, 4026–4031. (44) 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. (45) Osti, N. C.; Naguib, M.; Ostadhossein, A.; Xie, Y.; Kent, P. R.; Dyatkin, B.; Rother, G.; Heller, W. T.; van Duin, A. C.; Gogotsi, Y.; Mamontov, E. Effect of Metal Ion Intercalation on the Structure of MXene and Water Dynamics on Its Internal Surfaces. ACS Appl. Mater. Interfaces 2016, 8, 8859–8863.

30 ACS Paragon Plus Environment

Page 31 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

(46) Doona, C. J.; Baik, M. Y. Molecular Mobility in Model Dough Systems Studied by Time-Domain Nuclear Magnetic Resonance Spectroscopy. J. Cereal Sci. 2007, 45, 257–262. (47) Chen, C.; Xu, Z.; Han, Y.; Sun, H. Y.; Gao, C. Redissolution of Flower-Shaped Graphene Oxide Powder with High Density. ACS Appl. Mater. Interfaces 2016, 8, 8000– 8007. (48) Ding, X. L.; Zhang, H.; Wang, L.; Qian, H. F.; Qi, X. G.; Xiao, J. H. Effect of Barley Antifreeze Protein on Thermal Properties and Water State of Dough during Freezing and Freeze-Thaw Cycles. Food Hydrocolloids 2015, 47, 32–40. (49) Li, W. M.; Wang, P.; Xu, X. L.; Xing, T.; Zhou, G. H. Use of Low-Field Nuclear Magnetic Resonance to Characterize Water Properties in Frozen Chicken Breasts Thawed under High Pressure. Eur. Food Res. Technol. 2014, 239, 183–188. (50) Lindström, H.; Södergren, S.; Solbrand, A.; Rensmo, H.; Hjelm, J.; Hagfeldt, A.; Lindquist, S. E. Li+ Ion Insertion in TiO2(Anatase). 2. Voltammetry on Nanoporous Films. J. Phys. Chem. B 1997, 101, 7717–7722. (51) Yan, J.; Ren, C. E.; Maleski, K.; Hatter, C. B.; Anasori, B.; Urbankowski, P.; Sarycheva, A.; Gogotsi, Y. Flexible MXene/Graphene Films for Ultrafast Supercapacitors with Outstanding Volumetric Capacitance. Adv. Funct. Mater. 2017, 27, 1701264. (52) Levi, M. D.; Lukatskaya, M. R.; Sigalov, S.; Beidaghi, M.; Shpigel, N.; Daikhin, L.; Aurbach, D.; Barsoum, M. W.; Gogotsi, Y. Solving the Capacitive Paradox of 2D MXene Using Electrochemical Quartz-Crystal Admittance and In Situ Electronic Conductance Measurements. Adv. Energy Mater. 2015, 5, 1400815.

31 ACS Paragon Plus Environment

ACS Nano 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 32 of 33

(53) Shpigel, N.; Lukatskaya, M. R.; Sigalov, S.; Ren, C. E.; Nayak, P.; Levi, M. D.; Daikhin, L.; Aurbach, D.; Gogotsi, Y. In Situ Monitoring of Gravimetric and Viscoelastic Changes in 2D Intercalation Electrodes. ACS Energy Lett. 2017, 2, 1407–1415. (54) Wang, X. H.; Zhou, Y. C. Solid-Liquid Reaction Synthesis of Layered Machinable Ti3AlC2 Ceramic. J. Mater. Chem. 2002, 12, 455–460. (55) Segall, M. D.; Lindan, P. J. D.; Probert, M. J.; Pickard, C. J.; Hasnip, P. J.; Clark, S. J.; Payne, M. C. First-Principles Simulation: Ideas, Illustrations and the Castep Code. J. Phys.: Condens. Matter 2002, 14, 2717–2744. (56) Perdew, J. P.; Burke, K.; Ernzerhof, M. Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 1996, 77, 3865–3868. (57) Methfessel, M.; Paxton, A. High-Precision Sampling for Brillouin-Zone Integration in Metals. Phys. Rev. B 1989, 40, 3616–3621. (58) Gao, S. P.; Pickard, C. J.; Perlov, A.; Milman, V. Core-Level Spectroscopy Calculation and the Plane Wave Pseudopotential Method. J Phys.: Condens. Matter 2009, 21, 104203.

32 ACS Paragon Plus Environment

Page 33 of 33 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Nano

Table of Contents Graphic

33 ACS Paragon Plus Environment