M′2MxXyene (M′ and M are the early transitional metals and X is carbide with x = 1 for y = 2 and x = 2 for y = 3) are the ordered double transitional metal layered carbides derived from their parent MAX phases M′2MxAlXyene by a wet chemical etching method. Their oxides are predicted to have topological properties for which they should be annealed at around 800 °C in an oxygen background. This paper includes the new ablated plasma thrust method for the ionization and adsorption of oxygen on the M′2MxXyene substrate in the pulsed laser deposition chamber. We have found that the background pressure has a negative effect and the substrate temperature has a positive effect on plume expansion. The density profile of the background gas is highly affected by deposition temperature. Similarly, it is found that the density of plasma generated by longer wavelengths is not affected significantly due to the inverse bremsstrahlung process. A shorter wavelength produces the bremsstrahlung process as photoionization takes place. At a certain time (200 ns), the pressure of background gas and plasma pressure are equal (snow-plogh effect) so that all the wavelengths then produce electrons (highest for shorter wavelengths), thereby increasing its density. The energy necessary for the oxidation of the substrate is provided by the energy of the ablated species. The adsorption is assured by the reflective high electron energy diffraction technique, and it is found that the ambient gas pressures p = 0.1 mbar and 0.2 mbar are appropriate for the adsorption process. The obtained M′2MxXyene oxides can be used for their topological test.

The discovery of high-temperature superconductivity is the most awaited moment for material scientists. MXene has recently been the most promising material in several aspects of graphene. MXene has the general formula Mn+1Xn, with n = 1, 2, and 3, derived from the MAX phase, where M is an early transitional metal, X is carbide or nitride, and A is the A block element.1 We have used Mo2TixCyene (x = 1 for y = 2 and x = 2 for y = 3; Mo and Ti are the outer and inner transitional metal layers), the ordered double transitional metal layered carbides for their oxidation.2,3 However, there are many challenges in synthesizing the oxides of MXenes due to their sensitive layered properties.4 For growing their oxides and analyzing their surface properties, Pulsed Laser Deposition (PLD) and Angle-Resolved Photoemission Spectroscopy (ARPES) are the most powerful and appropriate techniques. The Ablated Plasma (AP)-thrust method, as a new concept, is adopted for the oxidation of the MXene surface. In this method, the MXene material used as a substrate is oxidized with atomic oxygen that is derived from the oxygen molecule filled as background gas in the PLD chamber. The energy needed for the atomicity of oxygen is provided by the thermal energy of the ablated plume in the form of plasma5,6 originated from the target. The proper ionization and thrusting by the plume at a stopping distance nearly equal to the substrate target distance leave an appropriate environment for the oxidation. The atomic oxygen will occupy the hollow site2 above the carbon atoms in the double-ordered MXenes. The oxidized sample surface can be checked by Reflective High-Energy Electron Diffraction (RHEED) and fast photography (ICCD—Intensified Charge Coupled Device) approaches. The ambient gas pressure for oxidation is found to be 0.1 and 0.2 mbar for an input laser energy of 400 mJ, an energy density of 2 J cm−2, 300 °C, and a pulse repetition rate 2 Hz. The oxidized MXene can be cleaved in situ through the proper crystallographic axes for its flat and plane surface for the ARPES test requirement. Theoretically, it has been shown that the oxidized MXene has topological insulating properties.7 We are expecting its experimental verification.

PLD is one of the effective (simple and fast) techniques used for thin film deposition and nanomaterial synthesis. It can produce a scalable sample in 10–15 min. Laser technology used in PLD has made it more effective in depositing complex stoichiometry multi-element and multi-layered oxides, nitrides, superlattices, etc.8 Materials such as YBa2Cu3O7−δ (YBCO) single crystals, without a natural cleaving plane, have problems in their characterization. This paper focuses on MXene oxide synthesis by AP-thrust PLD. As discussed theoretically, the surface of an (Ordered Double Transition Metal Layered) ODTML MXene has hollow sites above the carbon atom with a required formation energy of 4.40 eV for Mo2TiC2O2, and the formation energy for Mo2Ti2C3O2 is a little bit higher than that.9 For a 2 × 2 × 1 supercell, the formation energy of oxygen vacancy is obtained as follows:

Ef=EM8M4C8O7+12E(O2)EM8M4C8O8,
(1)

where E(X) is the total energy of system X.10 The oxygen formed by the plasma plume occupies the site and forms MXene oxide.

The plasma plume expansion in vacuum was studied by Anisimov et al.11 and Singh and Narayan,12 while that in background gas was described by Predtechensky and Mayorov (external dynamics of the contact front)13 and Arnold et al. (internal dynamics and background gas).14 The highest deposition rate of PLD is of the order of 1020 atoms cm−2 s−1.15,16 This gives the higher degree of supersaturation according to the following equation:

Δμ=κBTlnRRo,
(2)

where κB, R, and Ro are the Boltzmann constant, actual deposition rate, and rate at equilibrium temperature T. In our case, the adsorption of oxygen is considered. So, we are concerned with the adsorption rate rather than the growth rate. Instead, it is a strong reference for us.

The Mo2TiC2 and Mo2Ti2C3 sheets were used as substrates for oxidation in the PLD system. The PLD system used at the RRCAT, Indore, was Lambda Physik Complex 201 where we used a bulk aluminum target17 of 1 diameter and 1–2 mm thickness. The delaminated M′2MxXyene substrate can be adjusted from room temperature to 850 °C, with a repetition rate in the range of 1–10 Hz and a maximum pulse energy of 600 mJ, and the target to substrate distance18,19 can be varied from 4 to 10 cm. It uses a high-power KrF excimer (∼248 nm, ∼20 ns, ∼10 Hz, and ∼108 Wcm−2) to ablate the rotating target material. It can be adjusted with operating wavelengths of 992, 496, and 248 nm with a pulse duration of nearly 8 ns with the help of a delay generator. The energy density of the laser is controlled by adjusting the mask size and magnification. A multitarget deposition chamber with six targets deposits thin film with a computer-controlled system. Materials such as complex ceramic oxides, high-temperature superconductors, Heusler’s alloys, and silicon carbides can be easily treated in this chamber. The plasma plume is recorded using an ICCD camera (only data are taken here without images), and the deposition on the substrate was recorded by high-pressure RHEED. We can observe all these phenomena in the PLD chamber through the view port as shown in Fig. 1. The adsorption site of oxygen on the surface is the hollow site above the carbon atom, as shown in Fig. 2, of the trigonal-hexagonal crystal structure with space group P3̄m1. Figures 2(a) and 2(b) show the top and side views of Mo2TiC2 and Mo2Ti2C3 and their oxides, respectively. The sources of oxygen in the chamber are (a) background oxygen, (b) target or plasma oxygen, and (c) substrate oxygen.20 The study by Chen et al. on O-18 claimed that the oxygen in the substrate was predominant on increasing background oxygen pressure.21 The measuring device was focused 1 mm apart from the target, and an aluminum spectral line at 281.62 nm was taken for density calculation.

FIG. 1.

Schematic diagram of the PLD chamber at the RRCAT.

FIG. 1.

Schematic diagram of the PLD chamber at the RRCAT.

Close modal
FIG. 2.

Adsorption of oxygen at the hollow site above the carbon atom of (a) Mo2TiC2 and (b) Mo2Ti2C3.2 

FIG. 2.

Adsorption of oxygen at the hollow site above the carbon atom of (a) Mo2TiC2 and (b) Mo2Ti2C3.2 

Close modal

1. Ambient gas pressure effect on plume propagation

At the laser wavelength of 248 nm, an energy fluence of 2 J cm−2, an exposure time of ∼8 ns, and a substrate to target distance of 4 cm at room temperature, the different sets of data were recorded with the help of the ICCD camera at different times and pressures. They were then plotted in a graph as shown in Fig. 3. Initially, at a time of ≤1 µs, a position of ≤1 cm, and a velocity of 12 km/s, the expansion is nearly linear. At around 2 cm, the expansion breaks due to the oxygen pressure. At lower pressure (p = 0.05 mbar), the plume easily reaches the substrate after 10 µs. However, at higher pressure (p = 0.2 mbar), the plume does not reach the substrate even at 20 µs and ultimately stops showing the snow-plough effect after the energy of ablated species is transferred to the background gas.

FIG. 3.

Graph of the distance traveled by the plume front vs time at three different oxygen background gas pressures.

FIG. 3.

Graph of the distance traveled by the plume front vs time at three different oxygen background gas pressures.

Close modal

The graph of plume front velocity (u) against perpendicular distance from the target surface R is shown in Fig. 4. In this figure, the velocity decreases sharply along with the distance profile due to the effect of background pressure. The graph of the distance traveled by the plume front against time at three different substrate temperatures and the plot of substrate temperature and corresponding plume front velocity are shown in Figs. 5 and 6, respectively.

FIG. 4.

Graph of plume front velocity (u) vs perpendicular distance from the target surface R.

FIG. 4.

Graph of plume front velocity (u) vs perpendicular distance from the target surface R.

Close modal
FIG. 5.

Graph of the distance traveled by the plume front vs time at three different substrate temperatures.

FIG. 5.

Graph of the distance traveled by the plume front vs time at three different substrate temperatures.

Close modal
FIG. 6.

Plot of substrate temperature vs corresponding plume front velocity.

FIG. 6.

Plot of substrate temperature vs corresponding plume front velocity.

Close modal

2. Effect of substrate temperature on plume propagation

In this case, the background pressure is kept constant, say, at p = 0.2 mbar, and the plume front expansion is observed for substrate temperatures of 300, 600, and 900 °C with the help of ICCD images. The related data were recorded and are plotted in a graph as shown in Fig. 4. This figure shows that there is a significant increase in the velocity (kinetic energy) of the contact front with higher substrate temperature. The rate of change increases for higher temperatures, i.e., up to 300 °C, and the change was ≈0.03–0.05 eV, and from 800 to 900 °C, the change in the energy was ≈0.2–0.3 eV. The plot of substrate temperature against corresponding plume front velocity is shown in Fig. 5. The results are similar to those of the previous one.

3. Pulsed laser energy and wavelength effect in plasma ablation

An Yttrium Aluminum Garnet (YAG) excimer is used, and we recorded the data obtained from the images captured by fast photography (ICCD camera) at first 0–30 ns under the effect of different wavelengths with different energies on the interaction that takes place in the process of adsorption.22,23 The fundamental (996 nm), second (496 nm), and fourth harmonic (248 nm) wavelengths with a full wave half maximum of 6 ns were focused onto the 2 mm thick aluminum target (spot size, 100 mm) through a plano-convex quartz lens to produce plasma. The laser pulse energies were 10, 20, and 40 mJ, respectively. Their values per unit cm2 are 153, 318, and 636 J, respectively. The energy after the explosion is 10%–15% of the input laser energy.24,25 

The distance–time plot for the spherical plume (n = 3) for wavelengths of (a) 248 nm, (b) 496 nm, and (c) 992 nm is shown in the Figs. 7(a)7(c) respectively. The curves obtained for laser energies of 10 and 20 mJ are fitted well on spherical Taylor–Sedov solution,26 which agree well with the input laser energies. The plasma generated by a longer wavelength of 996 nm with a higher energy of 40 mJ is more cylindrical than that generated by shorter ones. This is due to the phenomenon of the Inverse Bremsstrahlung (IB) process induced by longer wavelength radiation. If the process was carried by a shorter wavelength, then the process is bremsstrahlung due to photoionization (PI) as studied by Chang et al. in 1996.23,27 The photoionization process involving higher temperature does not work in our surface-sensitive substrate. So, the input laser with a longer wavelength and larger energy fluence is appropriate for the adsorption process.

FIG. 7.

Distance–time plot for the spherical plume for wavelengths (a) 248 nm, (b) 496 nm, and (c) 992 nm.

FIG. 7.

Distance–time plot for the spherical plume for wavelengths (a) 248 nm, (b) 496 nm, and (c) 992 nm.

Close modal

Similarly, the R–t plot for a pulse energy of 10 mJ at 992, 496, and 248 nm laser ablation (obtained from ICCD images) is shown in Fig. 8. The best fits given in Fig. 8 are obtained using the drag model. The blast model is not fitted. The result agrees well for 992 nm but deviates for 248 and 496 nm early just opposite to the previous case.

FIG. 8.

Position–time plot of 992, 496, and 248 nm ablation for a pulse energy of 10 mJ.

FIG. 8.

Position–time plot of 992, 496, and 248 nm ablation for a pulse energy of 10 mJ.

Close modal

The temporal advancement of the electron density comparison for the three wavelengths of 248, 496, and 992 nm for each energy is shown in Figs. 9(a)9(c). The electron density decreases abruptly for all wavelengths and energies. This is due to the background oxygen. As the ablated plasma plume pressure and background pressure at around 200 ns become equal, all the wavelengths of different energies decayed to increase the electron number density (4 ± 0.5 × 1016 cm−3). The electron density at 992 nm was the lowest and that at 248 nm was the highest. Moreover, the increase in laser energy increases the electron number density. The highest density for 10 and 40 mJ is 5.5 × 1017 cm−3 and 1.2 × 1018 cm−3 for 248 nm and 4.2 × 1017 cm−3 and 1.18 × 1018 cm−3 for 496 nm, respectively. The density of plasma generated by a longer wavelength of 992 nm is not affected significantly.

FIG. 9.

Graph of the density of the electron vs time for (a) 10 mJ, (b) 20 mJ, and (c) 40 mJ.

FIG. 9.

Graph of the density of the electron vs time for (a) 10 mJ, (b) 20 mJ, and (c) 40 mJ.

Close modal

For the laser energy of 100 mJ, the pressure of 0.26 mbar, and the substrate–target distance of 4.5 cm, a set of data for the density of the atoms at different contact front distances is taken and plotted in the graph as shown in Fig. 10. In this figure, the density of oxygen is relatively low and is almost independent of the distance from the target to the contact front, while that for the ablated aluminum vapor slowly decreases with the distance. Similarly, the rate of decrease in temperature for vapor is higher than that for oxygen, which decreases rapidly than the rate of decrease in temperature of oxygen with the target substrate distance as shown in Fig. 11. The temperature of the ablated material above the substrate temperature of 7000 K decreases below 1000 K at a distance below 4 cm from the substrate but that for oxygen is still high. This decrease shows the possibility of adsorption of oxygen on the substrate.

FIG. 10.

Graph of the density of the atoms vs the position of the contact front near the substrate.

FIG. 10.

Graph of the density of the atoms vs the position of the contact front near the substrate.

Close modal
FIG. 11.

Temperature of the atoms vs position of the contact front near the substrate.

FIG. 11.

Temperature of the atoms vs position of the contact front near the substrate.

Close modal

The graph of the temperature of the atoms in the shock wave vs. the pressure of the background oxygen near the substrate is shown in Fig. 12. As the background oxygen pressure is increased, its temperature within the thin layer increases, while that of vapor decreases rapidly. This indicates that the ablated materials transfer their kinetic and thermal energies to the oxygen molecules, and condense back to the target. This helps in the ionization of oxygen molecules for the production of oxygen atoms that are easily absorbed. Similarly, the energy of atoms near the substrate plotted against the background pressure is shown in Fig. 13. The energy of oxygen is high and that of vapor is low but decreases in the same way with an increment of pressure.

FIG. 12.

Graph of the temperature of the atoms vs pressure of the background oxygen near the substrate.

FIG. 12.

Graph of the temperature of the atoms vs pressure of the background oxygen near the substrate.

Close modal
FIG. 13.

Energy of the atoms vs pressure of background oxygen near the substrate.

FIG. 13.

Energy of the atoms vs pressure of background oxygen near the substrate.

Close modal

The background pressure can control the energy of the impinging particles from being incident onto the surface. The density of the atoms of the laser plume within the shock layer is of the order of 1016/cm2, and the temperature is below 1000 K or 727 °C. This condition is favorable for the oxidation of the metal atoms, condensation, and the formation of the clusters. This temperature is equally suitable for the oxidation of MXene up to which all of its properties are preserved. These data are in good agreement with the previous literature.13 

Reflection High Energy Electron diffraction (RHEED) characterizes crystalline surfaces with the help of a diffraction pattern with a spot as maximum intensity.28 Cleaving or coating with passive oxides helps to get a clean surface,29 which can be removed by annealing. The RHEED technique considers both reciprocal and real space observations.30 The RHEED pattern at four different pressures [(1) p = 0.05 mbar, (2) p = 0.1 mbar, (3) p = 0.2 mbar, and (4) p = 0.3 mbar] for Mo2TiC2 and Mo2Ti2C3 is studied in this work. The energy of the laser is 400 mJ with surface energy density F = 2 J cm−2, target to substrate distance dT–S = 4 cm, and pulse repetition rate ν = 2 Hz.

1. RHEED oscillation and diffraction patterns at p = 0.05 mbar

The pressure is adjusted to p = 0.05 mbar, and Ek is in the order of 10 eV. The stopping distance (average path before thermalization) for the plume is larger than the deposition distance, i.e., dstop ≫ dT–S. The strong damping in the oscillation (0, 0) indicates the transition to a multilayer system in both Mo2TiC2 and Mo2Ti2C3 as shown in Figs. 14 and 15, respectively. In addition, the growth is so focused on one point that even adsorption is not completed on one unit cell. The deposition takes place on the top of the island. In this condition, aluminum is also deposited on the surface, resulting in a useless surface. However, the surfaces of both the MXenes look similar. The reconstruction of the pattern is due to the post nucleation process. At this pressure, the circular spots (0, 0) and (0, ±1) are distorted to rods, indicating the more disordered surface of both MXenes as shown in Figs. 16 and 17, respectively. The oscillating pattern gives the deposition rate. The kinetic energy of the ablated particles is so high that the surface uniformity and flatness are destroyed. Both the initial and final patterns of are different.

FIG. 14.

RHEED oscillations during adsorption in Mo2TiC2.

FIG. 14.

RHEED oscillations during adsorption in Mo2TiC2.

Close modal
FIG. 15.

RHEED oscillations during adsorption in Mo2Ti2C3.

FIG. 15.

RHEED oscillations during adsorption in Mo2Ti2C3.

Close modal
FIG. 16.

RHEED patterns of (a) the starting surface of Mo2TiC2 and (b) the final surface of Mo2TiC2.

FIG. 16.

RHEED patterns of (a) the starting surface of Mo2TiC2 and (b) the final surface of Mo2TiC2.

Close modal
FIG. 17.

RHEED patterns of (a) the starting surface of Mo2Ti2C3 and (b) the final surface of Mo2Ti2C3.

FIG. 17.

RHEED patterns of (a) the starting surface of Mo2Ti2C3 and (b) the final surface of Mo2Ti2C3.

Close modal

2. RHEED oscillation and diffraction pattern at p = 0.1 mbar

RHEED oscillatory behavior during adsorption in Mo2TiC2 under p = 0.1 mbar and Mo2Ti2C3 under p = 0.1 mbar is shown in Figs. 18 and 19, respectively. RHEED patterns of the starting surface of Mo2TiC2 and the final surface of Mo2TiC2 are shown in Figs. 20(a) and 20(b), respectively. RHEED patterns of the starting surface of Mo2Ti2C3 and the final surface of Mo2Ti2C3 are shown in Figs. 21(a) and 21(b), respectively.

FIG. 18.

RHEED oscillations during adsorption in Mo2TiC2 under p = 0.1 mbar.

FIG. 18.

RHEED oscillations during adsorption in Mo2TiC2 under p = 0.1 mbar.

Close modal
FIG. 19.

RHEED oscillations during adsorption in Mo2Ti2C3 under p = 0.1 mbar.

FIG. 19.

RHEED oscillations during adsorption in Mo2Ti2C3 under p = 0.1 mbar.

Close modal
FIG. 20.

RHEED pattern of (a) the starting surface of Mo2TiC2 and (b) the final surface of Mo2TiC2.

FIG. 20.

RHEED pattern of (a) the starting surface of Mo2TiC2 and (b) the final surface of Mo2TiC2.

Close modal
FIG. 21.

RHEED patterns of (a) the starting surface of Mo2Ti2C3 and (b) the final surface of Mo2Ti2C3.

FIG. 21.

RHEED patterns of (a) the starting surface of Mo2Ti2C3 and (b) the final surface of Mo2Ti2C3.

Close modal

3. RHEED oscillation and diffraction pattern at p = 0.2 mbar

RHEED oscillatory behavior during adsorption in Mo2TiC2 under p = 0.2 mbar and Mo2Ti2C3 under 0.2 mbar is shown in Figs. 22 and 23, respectively. RHEED patterns of the starting surface of Mo2TiC2 and the final surface of Mo2TiC2 are shown in Figs. 24(a) and 24(b), respectively. RHEED patterns of the starting surface of Mo2Ti2C3 and the final surface of Mo2Ti2C3 are shown in Figs. 25(a) and 25(b), respectively.

FIG. 22.

RHEED oscillations during adsorption in Mo2TiC2 under p = 0.2 mbar.

FIG. 22.

RHEED oscillations during adsorption in Mo2TiC2 under p = 0.2 mbar.

Close modal
FIG. 23.

RHEED oscillations during adsorption in Mo2Ti2C3 under p = 0.2 mbar.

FIG. 23.

RHEED oscillations during adsorption in Mo2Ti2C3 under p = 0.2 mbar.

Close modal
FIG. 24.

RHEED patterns of (a) the starting surface of Mo2TiC2 and (b) the final surface of Mo2TiC2.

FIG. 24.

RHEED patterns of (a) the starting surface of Mo2TiC2 and (b) the final surface of Mo2TiC2.

Close modal
FIG. 25.

RHEED patterns of (a) the starting surface of Mo2Ti2C3 and (b) the final surface of Mo2Ti2C3.

FIG. 25.

RHEED patterns of (a) the starting surface of Mo2Ti2C3 and (b) the final surface of Mo2Ti2C3.

Close modal

4. RHEED oscillation and diffraction pattern at p = 0.3 mbar

For the pressure of 0.01 mbar, there are three spots and reciprocal lattice rods seen on the zeroth order Laue circle with the respective point (0, 0) at the center and (0, −1) and (0, +1) on the left and right, respectively [Figs. 20(a) and 21(a)]. There are three sharp brighter and circular spots and little stretched rods in the final diffraction better than the initial one as shown in Figs. 20(b) and 21(b). This confirms the flatness and smoothness of the substrate surface. There is an intensity variation during growth at the (0, 0) spot, showing an oscillatory behavior as shown in Figs. 18 and 19 for Mo2TiC2 and Mo2Ti2C3, respectively. The intensity continuously increases up to 800s (nearly 14 min), indicating the improvement in the smoothness and flatness of the surface compared with the initial one. The kinetic energy of the ablated species is calculated from the plume dynamic analysis and found to be in the order of a few eV. Here, dstop ≈ dT–S.

The (0, 0) spot intensity (at p = 0.2 mbar) is varied with adsorption time and better than that at p = 0.1 mbar. The (0, 0) RHEED intensity shows an oscillatory nature almost with the same amplitude, indicating the adsorption of good quality again as shown in Figs. 22 and 23. The damping indicates the ending of the adsorption process. The first four maxima indicate the quality adsorption of four monolayers (ML). The kinetic energy is in the order of 0.1 eV, and dstop ≤ dT–S. At this pressure, the ablated material does not reach the substrate rather it thrusts and ionizes the oxygen atoms. The respective initial and final RHEED patterns are shown in Figs. 24 and 25, respectively.

The pressure is changed to p = 0.3 mbar, and the estimated kinetic energy is in the order of 0.01 eV, which is not sufficient even to ionize the oxygen molecules into oxygen atoms. The plume stopping distance is much smaller than the deposition distance, i.e., dstop ≪≪ dT −S. The mobility of oxygen ions is very low. The rare oscillation at (0, 0) indicates no adsorption in both Mo2TiC2 and Mo2Ti2C3 as shown in Figs. 26 and 27, respectively. At this pressure, the circular spots (0, 0) and (0, ±1) are distorted into rods, indicating the more disordered surface of both MXenes as shown in Figs. 28 and 29, respectively. The kinetic energy of the ablated particles is so small that there is a slight effect on the surface uniformity and flatness.

FIG. 26.

RHEED oscillations during adsorption in Mo2TiC2 under p = 0.3 mbar.

FIG. 26.

RHEED oscillations during adsorption in Mo2TiC2 under p = 0.3 mbar.

Close modal
FIG. 27.

RHEED oscillations during adsorption in Mo2Ti2C3 under p = 0.3 mbar.

FIG. 27.

RHEED oscillations during adsorption in Mo2Ti2C3 under p = 0.3 mbar.

Close modal
FIG. 28.

RHEED pattern of (a) the starting surface of Mo2TiC2 and (b) the final surface of Mo2TiC2.

FIG. 28.

RHEED pattern of (a) the starting surface of Mo2TiC2 and (b) the final surface of Mo2TiC2.

Close modal
FIG. 29.

RHEED patterns of (a) the starting surface of Mo2Ti2C3 and (b) the final surface of Mo2Ti2C3.

FIG. 29.

RHEED patterns of (a) the starting surface of Mo2Ti2C3 and (b) the final surface of Mo2Ti2C3.

Close modal

The kinetic energy of the species impinging onto the substrate (Ek), dstop vs dT–S, Ek vs EA, and the nature of oscillation are summarized in Table I.31 

TABLE I.

Experimental data obtained for different pressures.

p (mbar)Ek (eV)dstop vs dT–SEk vs EAOscillations
0.05 10 eV dstop > dT–S Ek > EA Strongly damped 
0.1 1 eV dstop ∼ dT–S Ek ∼ EA Oscillation 
0.2 0.1 of eV dstop < dT–S Ek < EA Oscillation 
0.3 0.01 of eV dstop < dT–S Ek ≪ EA No oscillation 
p (mbar)Ek (eV)dstop vs dT–SEk vs EAOscillations
0.05 10 eV dstop > dT–S Ek > EA Strongly damped 
0.1 1 eV dstop ∼ dT–S Ek ∼ EA Oscillation 
0.2 0.1 of eV dstop < dT–S Ek < EA Oscillation 
0.3 0.01 of eV dstop < dT–S Ek ≪ EA No oscillation 

It can be seen from this table that the kinetic energy of the ablated species has an inverse relation with the deposition pressure. The thermal energy obtained from the kinetic energy is given by Eth = kBT and has a contribution to the mobility of the oxygen atoms to be adsorbed. In addition, it damages the surface with non-recoverable defects.32–36 The pressures of 0.1 and 0.2 mbar for a pulse energy of 400 mJ with an energy density of 2 J cm−2 are good for the adsorption of oxygen.

A new ablated plasma thrust method is successfully used for the ionization and adsorption of oxygen on the MXene substrate in the PLD chamber. It is found that the background pressure has a negative effect and the substrate temperature has a positive effect on plume expansion. The density profile of the background gas is highly affected by deposition temperature. Similarly, the plasma generated by longer wavelengths is more cylindrical than that generated by shorter ones due to the inverse bremsstrahlung process. The transfer of the kinetic energy of the confined plasma plume to thermal energy is very important for the ionization and activation of molecular oxygen. From the RHEED study, it is found that the ambient gas pressures p = 0.1 mbar and 0.2 mbar are appropriate for the adsorption process of the oxygen atoms to be adsorbed at the hollow sites of Mo2TiC2 and Mo2Ti2C3. This might be the first approach in which we use ablated species only for ionizing and pressurizing the background gas onto the substrate for adsorption in the form of complex MXene oxide. The microstructural study shows close agreement with the RHEED measurement.

We thank Dr. S. N. Jha and his team for their support in obtaining experimental data at the RRCAT, Indore.

The authors have no conflicts to disclose.

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
Y.
Gogotsi
and
B.
Anasori
, “
The rise of MXenes
,”
ACS Nano
13
(
8)
,
8491
8494
(
2019
).
2.
D.
Parajuli
and
K.
Samatha
, “
MXene as topological insulator
,”
J. Emerging Technol. Innovative Res.
6
(
4
),
689
706
(
2019
).
3.
B.
Anasori
 et al, “
Two-dimensional, ordered, double transition metals carbides (MXenes)
,”
ACS Nano
9
(
10
),
9507
9516
(
2015
).
4.
M.
Seredych
 et al, “
High-temperature behavior and surface chemistry of carbide MXenes studied by thermal analysis
,”
Chem. Mater.
31
(
9
),
3324
3332
(
2019
).
5.
D.
Parajuli
and
J. J.
Nakarmi
, “
High frequency power absorption in inverse bremsstrahlung process in plasma
,”
J. NPS
22
(
1
),
14
17
,
2006
.
6.
D.
Parajuli
and
R.
Khanal
, “
Construction of high voltage source for plasma production
,”
J. Nepal Phys. Soc.
23
(
1
),
44
47
(
2007
).
7.
M.
Khazaei
,
A.
Ranjbar
,
M.
Arai
, and
S.
Yunoki
, “
Topological insulators in the ordered double transition metals M2′ M C2 MXenes (M′ = Mo, W; M = Ti, Zr, Hf)
,”
Phys. Rev. B
94
(
12
),
125152
(
2016
).
8.
B.
Wang
and
P.
Král
, “
Chemically tunable nanoscale propellers of liquids
,”
Phys. Rev. Lett.
98
(
26
),
266102
(
2007
).
9.
Y.-W.
Cheng
,
J.-H.
Dai
,
Y.-M.
Zhang
, and
Y.
Song
, “
Two-dimensional, ordered, double transition metal carbides (MXenes): A new family of promising catalysts for the hydrogen evolution reaction
,”
J. Phys. Chem. C
122
(
49
),
28113
28122
(
2018
).
10.
W.
Jiang
 et al, “
Universal descriptor for large-scale screening of high-performance MXene-based materials for energy storage and conversion
,”
Chem. Mater.
30
(
8
),
2687
2693
(
2018
).
11.
S. I.
Anisimov
,
B. S.
Luk’yanchuk
, and
A.
Luches
, “
An analytical model for three dimensional laser plume expansion into vacuum in hydrodynamic regime
,”
Appl. Surf. Sci.
96–98
,
24
32
(
1996
).
12.
R. K.
Singh
and
J.
Narayan
, “
Pulsed-laser evaporation technique for deposition of thin films: Physics and theoretical model
,”
Phys. Rev. B
41
(
13
),
8843
8859
(
1990
).
13.
M. R.
Predteceensky
and
A. P.
Mayorov
, “
Expansion of laser plasma in oxygen at laser deposition of HTSC films: Theoretical model
,”
Appl. Supercond.
1
(
10–12
),
2011
2017
(
1993
).
14.
N.
Arnold
,
J.
Gruber
, and
J.
Heitz
,
Appl. Phys. A
69
,
S87
S93
(
1999
).
15.
G.
Rijnders
,
The Initial Growth of Complex Oxides: Study and Manipulation
(
University of Twente
,
2001
).
16.
M.
Huijben
and
M.
Huijben
, “
Tuning electronic properties by atomically controlled growth
,” Ph.D. thesis,
University of Twente
,
The Netherlands
,
2006
.
17.
H. B.
Profijt
,
S. E.
Potts
,
M. C. M.
van de Sanden
, and
W. M. M.
Kessels
, “
Plasma-assisted atomic layer deposition: Basics, opportunities, and challenges
,”
J. Vac. Sci. Technol. A
29
(
5
),
050801
(
2011
).
18.
T.
Ohnishi
,
K.
Shibuya
,
T.
Yamamoto
, and
M.
Lippmaa
, “
Defects and transport in complex oxide thin films
,”
J. Appl. Phys.
103
(
10
),
103703
(
2008
).
19.
J. D.
Ferguson
,
G.
Arikan
,
D. S.
Dale
,
A. R.
Woll
, and
J. D.
Brock
, “
Measurements of surface diffusivity and coarsening during pulsed laser deposition
,”
Phys. Rev. Lett.
103
(
25
),
256103
(
2009
).
20.
D.
Dijkkamp
 et al, “
Preparation of Y-Ba-Cu oxide superconductor thin films using pulsed laser evaporation from high Tc bulk material
,”
Appl. Phys. Lett.
51
(
8
),
619
621
(
1987
).
21.
J.
Chen
 et al, “
Tracing the origin of oxygen for La 0.6 Sr 0.4 MnO 3 thin film growth by pulsed laser deposition
,”
J. Phys. D: Appl. Phys.
49
(
4
),
045201
(
2015
).
22.
S.-B.
Wen
,
X.
Mao
,
R.
Greif
, and
R. E.
Russo
, “
Laser ablation induced vapor plume expansion into a background gas. II. Experimental analysis
,”
J. Appl. Phys.
101
(
2
),
023115
(
2007
).
23.
A. E.
Hussein
,
P. K.
Diwakar
,
S. S.
Harilal
, and
A.
Hassanein
, “
The role of laser wavelength on plasma generation and expansion of ablation plumes in air
,”
J. Appl. Phys.
113
(
14
),
143305
(
2013
).
24.
S. S.
Harilal
,
T.
Sizyuk
,
A.
Hassanein
,
D.
Campos
,
P.
Hough
, and
V.
Sizyuk
, “
The effect of excitation wavelength on dynamics of laser-produced tin plasma
,”
J. Appl. Phys.
109
(
6
),
063306
(
2011
).
25.
T. Y.
Choi
and
C. P.
Grigoropoulos
, “
Plasma and ablation dynamics in ultrafast laser processing of crystalline silicon
,”
J. Appl. Phys.
92
(
9
),
4918
4925
(
2002
).
26.
S. G.
Taylor
, “
The formation of a blast wave by a very intense explosion I. Theoretical discussion
,”
Proc. R. Soc. London, Ser. A
201
(
1065
),
159
174
(
1950
).
27.
J. J.
Chang
and
B. E.
Warner
, “
Laser-plasma interaction during visible-laser ablation of methods
,”
Appl. Phys. Lett.
69
(
4
),
473
475
(
1996
).
28.
A.
Ichimiya
and
P. I.
Cohen
,
Reflection High Energy Electron Diffraction
(
Cambridge University Press
,
Cambridge, UK
,
2004
).
29.
A.
Dobson
and
P. J.
Howie
,
Surface and Interface Characterization by Electron Optical Methods
(
Plenum Press
,
New York
,
1988
).
30.
E. N.
Kaufmann
,
Characterization of Materials
, 2nd ed. (
John Wiley & Sons, Inc.
,
2012
).
31.
M.
Lippmaa
,
N.
Nakagawa
,
M.
Kawasaki
,
S.
Ohashi
, and
H.
Koinuma
, “
Growth mode mapping of SrTio3 epitaxy
,”
Appl. Phys. Lett.
76
(
17
),
2439
2441
(
2000
).
32.
P. R.
Willmott
,
R.
Herger
,
C. M.
Schlepütz
,
D.
Martoccia
, and
B. D.
Patterson
, “
Energetic surface smoothing of complex metal-oxide thin films
,”
Phys. Rev. Lett.
96
(
17
),
176102
(
2006
).
33.
M. E.
Taylor
and
H. A.
Atwater
, “
Monte Carlo simulations of epitaxial growth: Comparison of pulsed laser deposition and molecular beam epitaxy
,”
Appl. Surf. Sci.
127–129
,
159
163
(
1998
).
34.
H.
Jenniches
,
M.
Klaua
,
H.
Höche
, and
J.
Kirschner
, “
Comparison of pulsed laser deposition and thermal deposition: Improved layer-by-layer growth of Fe/Cu(111)
,”
Appl. Phys. Lett.
69
(
22
),
3339
3341
(
1996
).
35.
H.
Brune
,
G. S.
Bales
,
J.
Jacobsen
,
C.
Boragno
, and
K.
Kern
, “
Measuring surface diffusion from nucleation island densities
,”
Phys. Rev. B
60
(
8
),
5991
6006
(
1999
).
36.
E.
Vasco
and
C.
Zaldo
, “
Growth Kinetics of epitaxial Y stabilized ZrO2 films deposited on InP
,”
J. Phys.: Condens. Matter
16
(
5
),
8201
8211
(
2004
).