The atomic and electronic dynamics in the topological insulator (TI) Bi2Te3 under strong photoexcitation were characterized with time-resolved electron diffraction and time-resolved mid-infrared spectroscopy. Three-dimensional TIs characterized as bulk insulators with an electronic conduction surface band have shown a variety of exotic responses in terms of electronic transport when observed under conditions of applied pressure, magnetic field, or circularly polarized light. However, the atomic motions and their correlation between electronic systems in TIs under strong photoexcitation have not been explored. The artificial and transient modification of the electronic structures in TIs via photoinduced atomic motions represents a novel mechanism for providing a comparable level of bandgap control. The results of time-domain crystallography indicate that photoexcitation induces two-step atomic motions: first bismuth and then tellurium center-symmetric displacements. These atomic motions in Bi2Te3 trigger 10% bulk bandgap narrowing, which is consistent with the time-resolved mid-infrared spectroscopy results.

Three-dimensional topological insulators (TIs) which are characterized as bulk insulators with an electronic-conduction surface band1–4 have yielded a variety of exotic electronic-transport responses under magnetic field or circular polarized light.5–11 A capacitor-type device with thin-film TIs has been proposed for exciton condensates wherein quasi-particle excitation carries fractional charge.12,13 TIs have attracted considerable attention in both scientific and technical application fields. One of the interesting challenges regarding the technical application of TIs is the artificial and transient tuning of their electronic properties. The modification of electronic bandgaps at the surface and in the bulk of TIs has been demonstrated with static methods that use dopants or the application of pressure.13–16 However, the artificial and transient modification of bandgaps12,13 in TIs via photoexcitation has not thus far been examined. The bulk bandgap of a TI, e.g., Bi2Te3, has been characterized as a direct gap at cryogenic temperatures, which is “buried” by the indirect gap formed at room temperature.17 Therefore, the temperature dependence of the Bi2Te3 bandgap from cryogenic temperatures to room temperature is large. However, the temperature dependence of the indirect bandgap in Bi2Te3 above room temperature is negligibly small (see Fig. S1 of the supplementary material18). In this context, the responses of TIs under photoexcitation should be useful not only in understanding the nature of the electronic energy excitation and relaxation pathways but also in utilizing the ensuing photoinduced atomic displacements as a means to modify the electronic properties. The electronic dynamics in TIs under interband photoexcitation have been investigated with time- and angle-resolved photoelectron spectroscopy (tr-ARPES)19,20 and time-resolved mid-infrared (MIR) spectroscopy.21 The electrons photoexcited beyond the bandgap are immediately scattered into the bulk conduction band and partially into the surface state. The electrons in the bulk conduction band persistently relax into the bulk valence band through the surface state. A recent report has suggested that the balance between electrons and holes on the surface of TIs can strongly affect the electronic band structure of the system.22 In contrast, coherent phonons have been observed in both the bulk and surface state of TIs under relatively weak photoexcitation.23,24 The correlation of the electronic state dynamics to atomic motions under strong photoexcitation, particularly in the bulk, has not been considered in the dynamics of photoexcited TIs. Herein, we show that photoinduced atomic motions can be correlated to change in the electronic bandgap of TIs. We also examine the possibility of the artificial modification of the electronic structure in Bi2Te3 by interband photoexcitation via atomic motions occurring along the excited state surface.

Near-ultraviolet (NUV) (3.1 eV) pump and MIR (0.12–0.22 eV) probe experiments were performed in the reflection mode on a Bi2Te3 bulk single-crystal. The experimental details of this optical pump-probe setup are presented elsewhere.25 The incident angle of the pump and probe light was set nearly parallel to the surface normal of the sample. The pulse durations of the NUV and MIR pulses were 100 fs and <1 ps, respectively. The repetition rate and the incident fluence of the NUV pump pulse were 500 Hz and 1.7 mJ/cm2, respectively. The absorption pump fluence (FA) was calculated to be 1.0 mJ/cm2, and the reflectivity of Bi2Te3 was measured to be 41%.

Time-resolved electron diffraction (Tr-ED) measurements were next performed in the transmission mode. The experimental setup of the compact DC-accelerated electron diffraction is provided elsewhere.26 NUV light (3.1 eV) was focused to a 280-μm spot size on a single-crystalline Bi2Te3 ultrathin film (30 nm thick). The incident laser fluence was 0–9.0 mJ/cm2. From the transmission and reflectivity of the Bi2Te3 thin film measured to be 15% and 41%, respectively, FA was determined to be 0–4.0 mJ/cm2. The acceleration voltage of the probe electron pulses was 75 keV under a DC electric field. Photoinduced structural changes inside the material were investigated with electron pulses containing 1.25 × 104 electrons confined to a 100-μm-diameter spot incident on the Bi2Te3 sample. The pulse duration of the electron beam was <1.5 ps as measured by the plasma method.27 Diffracted and directly transmitted electrons were focused with a magnetic lens onto a 1:2 fiber-coupled charge-coupled device (CCD) camera (iKon-L HF, Andor) coated with a P43 (Gd2O2S:Tb) phosphor scintillator. To acquire one electron diffraction image, 1–4 × 104 shots of electron pulses were collected at repetition rates of 1 kHz below the FA value of 2 mJ/cm2 and 500 Hz above the FA value of 2 mJ/cm2 to avoid sample damage and accumulation of heat. The literature28,29 indicates that the carrier dynamics of this class of topological insulator in 30-nm-thick-film correspond to that of a bulk sample rather than the surface of a sample. In the following discussion, we consider the Bi2Te3 ultrathin film to be a bulk sample.

N-type Bi2Te3 single crystals were grown by means of the modified Bridgman technique, as described in Ref. 6. The carrier density of the sample was found to be 5 × 1018 cm−3. The surface of the bulk Bi2Te3 single crystal was cleaved with Scotch tape, and a fresh surface c-axis was used for the optical experiments. For the Tr-ED studies, the Bi2Te3 single crystal was cut to a 30-nm thickness perpendicular to the c-axis with a diamond knife on a microtome (UC7, Leica). The ultrathin Bi2Te3 film was “picked-up” with a commercial Cu TEM (#400) grid.

First-principles calculations within the framework of density functional theory (DFT) were performed with the use of the full-potential linearized augmented plane-wave (FP-LAPW) method with the generalized gradient approximation/Perdew–Burke–Ernzerhof functional (GGA/PBE) implemented in the WIEN2k code.30 Spin-orbit coupling was included as a second variation step. In order to describe the photoexcited states of Bi2Te3, the original rhombohedral crystal structure31 was converted to orthorhombic symmetry (a = 14.3362 Å, b = 57.5988 Å, c = 8.277 Å) with the space group B2 (origin choice 5). The atomic coordinates at 0, 10, and 100 ps used for the calculations are listed in Tables S1–S7.18 These coordinates were measured by time-resolved electron diffraction. The plane-wave cutoff energy was set to RMTKmax = 9, where RMT = 2.5 for both Bi and Te. The Brillouin zone was sampled with the Monkhorst–Pack scheme32 with a Γ-centered 35 × 6 × 35 k-point mesh, which achieved reasonable convergence (<1 meV) of the total energy and the bandgap.

NUV pump and MIR probe reflective spectroscopy were used to characterize the carrier dynamics on the relevant time scales. Figure 1(a) presents the differential reflectivity (ΔR/R) as a function of the probe photon energy at various time delays. As shown in Figs. 1(a) and 1(b), the signature of ΔR/R changes at a probe photon energy of approximately 0.13 eV. The decrease and recovery time constants in the transient reflectivity from Bi2Te3 vary from 10 to 3 ps and 60 to 15 ps, respectively, in the probe photon energy range of 0.12–0.22 eV (Fig. S2 provides more detail on the decrease and recovery time constants18). When the ground-state electrons undergo direct excitation to the far-from-equilibrium state, holes are created at the edge of the bulk valence band, and the electrons in the surface state recombine into the bulk valence band to “fill” the holes. The excited electrons are scattered into the edge of the bulk conduction bands and then relaxed into the bulk valence band through the surface state. The variation in the time constants in the probe energy as shown in Figs. 1(a) and 1(b) reflects carrier dynamics in the surface state in a manner similar to that explained in Ref. 21.

FIG. 1.

(a) Two-dimensional plot of the normalized reflectivity as a function of probe photon energy at various time delays. (b) Typical time-resolved reflectivity ΔR/R at probe photon energies below (0.125 eV) and above (0.18 eV) the bandgap.

FIG. 1.

(a) Two-dimensional plot of the normalized reflectivity as a function of probe photon energy at various time delays. (b) Typical time-resolved reflectivity ΔR/R at probe photon energies below (0.125 eV) and above (0.18 eV) the bandgap.

Close modal

The indirect bandgap of Bi2Te3 was confirmed to be 0.13 eV by transmission spectroscopy (Fig. S118) and an optical conductivity spectrum calculated using the Kramers–Kronig (K–K) transform of the reflectivity (Figs. 2(a) and 2(b)). For the K–K transformation, we measured reflectivity in the energy range of 0.07–2.4 eV (see Fig. S318) and assumed simulated reflectivity obtained using the conventional Drude model below 0.07 eV and ω2 extrapolation above 2.4 eV. The figure shows a “bump” structure in the reflectivity spectrum at 0.14–0.15 eV (blue arrow). We obtained more details on this structure when we plot the simulated data based on the simple Drude model (dashed line). A comparison between the optical conductivity spectra calculated with the K–K transform from the experimental and simulated data reveals a kink structure (blue arrow in Fig. 2(b)) that is related to the bump structure in reflectivity. For the low-energy extrapolation method, we examined several models and parameters, but we could not find significant differences with respect to the position of the kink structure in the optical conductivity spectra shown in Fig. 2(b). The reflectivity spectrum modulation on the time scale of the first 10 ps is particularly noteworthy. Figure 3 presents the transient reflectivity in the probe-photon energy range of 0.12–0.22 eV. From the K–K analysis results, we note that the bandgap of Bi2Te3 is strongly related to the kink structure (0.14–0.15 eV), indicated by blue arrows in Figs. 2(a) and 3. This kink structure, marked in Fig. 3, is clearly shifted toward a lower energy (10 meV) in the time-resolved spectra. The penetration depth of the MIR light in Bi2Te3 is 400 nm (above the bandgap) and 700 nm (below the bandgap), whereas that of NUV light is approximately 30 nm, which is considerably different.33–35 This mismatch of the penetration depth may affect the spectral shape of the reflectivity of the photoinduced state, however, we could not consider the effect of the mismatch on the pump-probe results because of the ambiguity of the spectral shape of the photoinduced state as well as the relatively small dispersion of the penetration depth in the photon energy range of our measurement. At present, we believe the observed shift of the kink structure in the reflectivity spectrum suggests the band-gap narrowing on the time scale of 10 ps. We present more detailed observations of the photoinduced dynamics on the results of the time-resolved electron diffraction.

FIG. 2.

(a) Mid-infrared reflectivity spectrum. The blue arrow indicates the bump positions. The simulated data (black dashed line) exhibit reflectivity without the kink structure. (b) The optical conductivity calculated using the K–K transformation of the reflectivity spectrum. The red solid line and black dashed line represent the experimental and simulated data, respectively. The black arrow indicates the bandgap energy.

FIG. 2.

(a) Mid-infrared reflectivity spectrum. The blue arrow indicates the bump positions. The simulated data (black dashed line) exhibit reflectivity without the kink structure. (b) The optical conductivity calculated using the K–K transformation of the reflectivity spectrum. The red solid line and black dashed line represent the experimental and simulated data, respectively. The black arrow indicates the bandgap energy.

Close modal
FIG. 3.

Time evolution of the reflectivity from photoexcited Bi2Te3. The black and red dotted lines are visual guides showing the kink structure and the transient bandgap narrowing, respectively, as indicated by arrows. The symbol (*) indicates the mid-infrared absorption by water.

FIG. 3.

Time evolution of the reflectivity from photoexcited Bi2Te3. The black and red dotted lines are visual guides showing the kink structure and the transient bandgap narrowing, respectively, as indicated by arrows. The symbol (*) indicates the mid-infrared absorption by water.

Close modal

We performed Tr-ED experiments that provided direct structural information on the evolution of the atomic configuration.26,36–38 Figures 4(a) and 4(b) show an electron diffraction pattern from the 30-nm-thick 〈0001〉-oriented Bi2Te3 single-crystalline film and a simulated diffraction pattern obtained with the WinHREM software (based on the dynamical diffraction theory with multislice methods39), respectively. We analyzed the most significant diffraction spots. The resulting Tr-ED intensities are presented in Figs. 5(a)5(c) for the { 1 2 ̄ 10 } and { 3 3 ̄ 00 } Bragg reflections from the Bi2Te3 film photoexcited at an FA value of 1.5 mJ/cm2. All the Tr-ED intensities are shown in Fig. S4.18 The Tr-ED intensities shown in Figs. 5(a) and 5(b) are the products of three contributions: Debye–Waller effects, acoustic phonons, and atomic displacements within the unit cell. The intensity of the electron diffraction (I) can be expressed via the following equation:40–42 

(1)

where K is the scattering vector, G(K) is the Laue function, fj(K) is the atomic scattering factor, Dj(K) is the Debye–Waller factor, and rj is the equilibrium position of the jth atom in the unit cell. Because the intensity modulation by the acoustic phonon is derived from the mismatch of the Bragg condition, the contribution of the acoustic phonon is only related to (G(K)). On the other hand, the thermal effects and the effects of atomic displacements within the unit cell yield inversion-symmetric modulation in the electron diffraction intensity. The resulting temperature rise can be independently calculated from the thermal expansion coefficients (1.3 × 105 K−1)43,44 and the shifts of the spot positions in the electron diffraction pattern (Fig. 5(d)). The value of the temperature rise from thermal expansion at 100 ps was calculated to be 100 K and to reach approximately 200 K on the time scale of nanoseconds. This temperature increase was also indicated by the diffuse scattering background45,46 (see Fig. S518). The Debye–Waller factor can be expressed as a function of the mean square displacement (Uj) as follows:

(2)

The isotropic thermal displacement (U) can be calculated using the coefficient of linear contraction (ρ), which is derived from the bulk modulus,44 and the atomic distance (dA = 4.38 Å) with the following equation:47,48

(3)

where kB and T are the Boltzmann constant and lattice temperature, respectively. Taking into account the room temperature of 300 K, the temperature rise of approximately 100 K corresponded to decreases in the electron diffraction intensity of 1% and 3% of their original values for the { 1 2 ̄ 10 } and { 3 3 ̄ 00 } reflections.

FIG. 4.

(a) Typical electron diffraction pattern from Bi2Te3 thin films. (b) Electron diffraction pattern from Bi2Te3 simulated with the software WinHREM based on dynamic theory.

FIG. 4.

(a) Typical electron diffraction pattern from Bi2Te3 thin films. (b) Electron diffraction pattern from Bi2Te3 simulated with the software WinHREM based on dynamic theory.

Close modal
FIG. 5.

(a)–(c) The time evolution of the electron diffraction intensity from { 1 2 ̄ 10 } planes corresponding to spots 1 and 2 and that from { 3 3 ̄ 00 } plane corresponding to spot 7 at FA = 1.5 mJ/cm2. The solid blue, green, and red lines indicate the contributions of the Debye–Waller thermal effect, acoustic phonon, and symmetric atomic displacement, respectively. (d) The peak shift of the { 1 2 ̄ 10 } spots positions in the time-resolved electron diffraction experiments and temperature rise. The lattice constant of the undisturbed { 1 2 ̄ 10 } plane is 0.2192 nm. (e) The Fourier transform of the acoustic phonon contribution. The blue lines in (a)–(c) were obtained independently from the red line in (d) and Eq. (2). To obtain the red lines in (a)–(c), double exponential fitting curves were fitted to the summations of the intensities of { 1 2 ̄ 10 } diffraction spots minus the blue lines. The green lines in (a) and (b) were obtained by fitting exponentially increasing and decreasing oscillating equations to the raw data after subtraction of the red and blue lines. The black lines in (a)–(c) are the sum totals of the red, blue, and green lines.

FIG. 5.

(a)–(c) The time evolution of the electron diffraction intensity from { 1 2 ̄ 10 } planes corresponding to spots 1 and 2 and that from { 3 3 ̄ 00 } plane corresponding to spot 7 at FA = 1.5 mJ/cm2. The solid blue, green, and red lines indicate the contributions of the Debye–Waller thermal effect, acoustic phonon, and symmetric atomic displacement, respectively. (d) The peak shift of the { 1 2 ̄ 10 } spots positions in the time-resolved electron diffraction experiments and temperature rise. The lattice constant of the undisturbed { 1 2 ̄ 10 } plane is 0.2192 nm. (e) The Fourier transform of the acoustic phonon contribution. The blue lines in (a)–(c) were obtained independently from the red line in (d) and Eq. (2). To obtain the red lines in (a)–(c), double exponential fitting curves were fitted to the summations of the intensities of { 1 2 ̄ 10 } diffraction spots minus the blue lines. The green lines in (a) and (b) were obtained by fitting exponentially increasing and decreasing oscillating equations to the raw data after subtraction of the red and blue lines. The black lines in (a)–(c) are the sum totals of the red, blue, and green lines.

Close modal

Acoustic phonon vibrations at a frequency of 35 GHz (Fig. 5(e)) were also observed in the intensity of the diffraction spot from the { 1 2 ̄ 10 } planes. In general, the sound velocity in a crystal is calculated as a function of the frequency of the acoustic phonon vibrations and film thickness.49,50 In this study, the frequency of the acoustic mode (35 GHz) and film thickness (30 nm) yielded a sound velocity of 2100 m/s, which corresponds to the longitudinal-mode sound velocity (2000–2200 m/s) in Bi2Te3.51,52 Since these acoustic phonon modulations have opposite phases in opposite planes, this effect of the acoustic phonons in the transient electron intensity can be canceled out by adding these opposite-phased oscillations. The remaining inversion six-symmetric contribution is attributed to the atomic displacements in the unit cell. As shown in Figs. 5(a) and 5(b), there is a two-step change: a decrease in the Tr-ED intensity by 5%–6% during the first 10 ps, and a subsequent gradual recovery over 10–100 ps. The symmetric-atomic-displacement-related component in the intensity of the higher-order diffraction spot { 3 3 ̄ 00 } exhibits a 25% initial drop and a slight subsequent recovery (Fig. 5(c)).

To determine the transient atomic coordinates in the photoexcited Bi2Te3 crystal, we performed model calculations with the simulation package WinHREM. The electron pulses pass through the sample, parallel to the c-axis of the Bi2Te3 crystal; therefore, the contributions along the c-axis (see Fig. S618) are responsible for significant uncertainty in the determination of the disordered atomic coordinates. For the purpose of this discussion, we used the diffraction spots from the { 1 2 ̄ 10 } and { 3 3 ̄ 00 } planes. In principle, two atomic coordinates in the ab plane can be determined with two independent diffraction spots. A quintuple layer (QL), known as a functional periodic unit of this class of TIs, consists of five atomic layers (Te1, Bi, Te2, Bi, and Te1). The Te1 atom occupies a different site than the Te2 atom from a crystallographic viewpoint. We assumed center-symmetric motions of Bi and Te1 atoms toward or away from the center axis of the unit cell with fixing of the Te2 atom. Simulations based on dynamic theory were performed by changing the atomic coordinates of the Bi and Te1 atoms. The calculations yielded differential electron diffraction intensities from the { 1 2 ̄ 10 } and { 3 3 ̄ 00 } planes as Bi atoms moved in relation to the Te2 atoms upon fixing of the Te1 atoms (Fig. S7(a)) and movement of the Te1 atoms in relation to the Te2 atoms with fixing of the Bi atoms (Fig. S7(b)).18 The intensities of the diffraction spots decrease as the Bi atoms move; however, the motion of the Te1 atoms makes no significant contribution to the intensities of the diffraction spots. Thus, the first-step changes observed in Figs. 5(a) and 5(b) (10 ps) were concluded to have been induced primarily by the motions of the Bi atoms. For the purpose of further elaboration of the atomic displaced coordinates, the two-step changes in the diffraction intensity were characterized (Figs. S7(c) and S7(d)).18 Transient atomic coordinates were determined to minimize R2(dBi, dTe) as expressed by the following equation:

(4)

where dBi and dTe are the displacements of the Bi and Te1 atoms, respectively, and Ii,exp and Ii,sim(dBi, dTe) are the experimental and simulated electron diffraction intensities of the ith index (i.e., { 1 2 ̄ 10 } and { 3 3 ̄ 00 } ), respectively. Figures 6(a)6(e) depict the R2(dBi, dTe) maps as functions of the displacements of the Bi and Te1 atoms at the representative time delays. The local minimum of the map is the transient atomic position, and the displacements of the Bi and Te1 atoms are summarized in Fig. 6(f). As the figure shows, the Bi and Te1 atoms move independently, and the motion observed in the first 10 ps corresponds to large Bi (∼8 pm) and small Te1 (∼2 pm) displacements toward or away from the center axis. The atomic motion over 10–100 ps is the Te1 (∼2 pm) motion along the same direction and reversal of the Bi (∼2 pm) displacement. The directions of these motions correspond to the fundamental-mode E1g and E2g phonons (see Fig. S8).18 These atomic displacements increase linearly with the absorption laser fluence at FA < 1.0 mJ/cm2 and are saturated at an FA value of 1.0 mJ/cm2. As shown in Fig. 7, the damage threshold of the ultrathin Bi2Te3 sample was determined to be 2.5 mJ/cm2. In the FA range from 1.0 to 2.0 mJ/cm2, the atomic dynamics of Bi and Te are identical. The atomic displacements observed with Tr-ED are summarized in Fig. 8, along with the QL structure of Bi2Te3. The disordered atomic coordinates are listed in Tables S1–S7.18 

FIG. 6.

Bi and Te1 displacements derived from the time-resolved electron diffraction intensity and simulated electron diffraction intensity at (a) 0 ps, (b) 2.4 ps, (c) 4.8 ps, (d) 8.4 ps, and (e) 24 ps, respectively. The black arrows in (a) indicate the directions of motions of the Eg1 and Eg2 phonon modes. The white arrows in (b)–(e) indicate the motions of Bi and Te1 atoms obtained from the analysis. (f) The displacements of Bi and Te1 atoms from their original positions as a function of time delay.

FIG. 6.

Bi and Te1 displacements derived from the time-resolved electron diffraction intensity and simulated electron diffraction intensity at (a) 0 ps, (b) 2.4 ps, (c) 4.8 ps, (d) 8.4 ps, and (e) 24 ps, respectively. The black arrows in (a) indicate the directions of motions of the Eg1 and Eg2 phonon modes. The white arrows in (b)–(e) indicate the motions of Bi and Te1 atoms obtained from the analysis. (f) The displacements of Bi and Te1 atoms from their original positions as a function of time delay.

Close modal
FIG. 7.

The absorption laser fluence dependence on the atomic motions at the time delay of 10 ps.

FIG. 7.

The absorption laser fluence dependence on the atomic motions at the time delay of 10 ps.

Close modal
FIG. 8.

Crystal structure of Bi2Te3. The quintuple layer (QL) consists of five atomic layers (Te1, Bi, Te2, Bi, and Te1). The motions of Bi and Te1 after the photoexcitation (10 ps and 100 ps) are indicated with the red arrows.

FIG. 8.

Crystal structure of Bi2Te3. The quintuple layer (QL) consists of five atomic layers (Te1, Bi, Te2, Bi, and Te1). The motions of Bi and Te1 after the photoexcitation (10 ps and 100 ps) are indicated with the red arrows.

Close modal

We discuss the correlation between the atomic motions and the change in the electronic structure of the Bi2Te3 system by comparing the results from the Tr-ED and NUV pump and MIR probe reflection spectroscopy. The transient electronic structure modulated by the disordered atoms can be calculated by DFT with the atomic coordinates obtained by Tr-ED. The DFT total energy increased by 1.0 meV per formula unit (f.u.) in 10 ps and relaxed to 0.2 meV/f.u. in 100 ps. These values correspond to excited electron densities of 4.5 × 1019 and <1 × 1019 cm−3, respectively, in the bulk conduction band where the photoexcited electron and hole density for NUV light is approximately 1 × 1021e cm−3. The density of the states obtained by the DFT calculation is shown in Figs. 9(a) and 9(b). The energy bandgap is 0.12 eV, which agrees with the spectroscopic measurements. The bandgap becomes approximately 5–10 meV narrower at 10 ps after photoexcitation. These results suggest that a fraction of the photoexcited and repopulated holes in the bulk valence band (around 5% of excited electrons and holes) create new atomic coordinates, which are related to the Bi motions occurring over 10 ps. These atomic motions cause bandgap narrowing, which is consistent with the NUV pump and MIR probe reflection spectroscopy results. Atomic displacements often reduce the crystal symmetry of a sample, which leads to the split of degenerated orbital level into several levels. This effect is called “crystal field splitting.”53,54 This crystal field splitting invokes the broadening of the conduction and valence bands, and results in bandgap narrowing unless the center value of the valence and conduction bands shifts. The bandgap renormalization effects55 via hot electrons might also induce bandgap narrowing; however, these hot-electron effects should start to occur within 3 ps when the hot electron is in the excited state. The time duration corresponding to the phenomena observed in our case was too large (10 ps) for the phenomena to be considered to be due only to hot-electron effects. Hot-hole effects56 triggering these atomic motions can indirectly form one of the possible mechanisms for bandgap narrowing.

FIG. 9.

(a) Density of states (DOS) calculated from unperturbed Bi2Te3. (b) Magnified view (gray box in (a)) of the DOS calculated from the displaced atomic coordinates obtained from the time-resolved electron diffraction.

FIG. 9.

(a) Density of states (DOS) calculated from unperturbed Bi2Te3. (b) Magnified view (gray box in (a)) of the DOS calculated from the displaced atomic coordinates obtained from the time-resolved electron diffraction.

Close modal

In summary, we observed that photoinduced atomic displacements in Bi2Te3 affect its bandgap, which is reflected in the DFT calculation. The observed dynamics illustrate the very nature of the photoexcited electronic relaxation pathways and the strong coupling between the electronic structure and atomic configuration of TIs. Bulk bandgap modification of TI films with intra- and interband photoexcitation in the range from THz to UV could play an important role in newer TI applications, particularly for modulating their properties on ultrafast time scales. To better understand electron–lattice coupling in TIs, it would be useful to observe the atomic motions along the c-axis with the use of ultrabright X-ray diffraction and spectroscopic methods.57,58 As computing methods to calculate the excited states of these complex materials are further developed, ab initio molecular dynamics will also provide a more detailed microscopic understanding of these electronically coupled atomic motions.59 In the present context, we have characterized both the transient electronic structure and atomic motions involved in photoinduced bandgap changes and demonstrated a causal connection to the atomic motions’ amenable control. The tabletop combination of time-resolved MIR spectroscopy and electron diffraction methodologies can provide significant insight into material science and condensed matter physics in strongly correlated materials, or even in chemistry, where information on electronic and atomic dynamics is important for designing new functional materials.60,61

This work was mainly supported by the Japan Science Technology Agency (JST), PRESTO, for a project entitled “Molecular technology and creation of new functions.” K.O. acknowledges JST-PRESTO for the project “Chemical conversion of light energy.” K.G.N. and Y.K. acknowledge JST-CREST for the project of “Enhancing Applications of Innovative Optical Science and Technologies by making Ultimate Use of Advanced Light Sources.” S.-y.K. acknowledges Grants-in-Aid for Scientific Research (Grant No. 15H02103) from the Japan Society for the Promotion of Science and the Ministry of Education, Culture, Sports, Science and Technology. S.K. and R.J.D.M. were supported by the Max Planck Society. The authors acknowledge Dr. Dongfang Zhang at the Max Planck Institute for the Structure and Dynamics of Matter and Professor Masahiro Kitajima at the Tokyo Institute of Technology for discussions.

1.
L.
Fu
and
C. L.
Kane
, “
Topological insulators with inversion symmetry
,”
Phys. Rev. B
76
,
045302
(
2007
).
2.
D.
Hsieh
 et al., “
A topological Dirac insulator in a quantum spin Hall phase
,”
Nature
452
,
970
(
2008
).
3.
H.
Zhang
 et al., “
Topological insulators in Bi2Se3, Bi2Te3 and Sb2Te3 with a single Dirac cone on the surface
,”
Nat. Phys.
5
,
438
(
2009
).
4.
Y. L.
Chen
 et al., “
Experimental realization of a three-dimensional topological insulator, Bi2Te3
,”
Science
325
,
178
(
2009
).
5.
D.-X.
Qu
,
Y. S.
Hor
,
J.
Xiong
,
R. J.
Cava
, and
N. P.
Ong
, “
Quantum oscillations and hall anomaly of surface states in the topological insulator Bi2Te3
,”
Science
329
,
821
(
2010
).
6.
Y. L.
Chen
 et al., “
Massive Dirac fermion on the surface of a magnetically doped topological insulator
,”
Science
329
,
659
(
2010
).
7.
P.
Cheng
 et al., “
Landau quantization of topological surface states in Bi2Se3
,”
Phys. Rev. Lett.
105
,
076801
(
2010
).
8.
J. W.
McIver
,
D.
Hsieh
,
H.
Steinberg
,
P.
Jarillo-Herrero
, and
N.
Gedik
, “
Control over topological insulator photocurrents with light polarization
,”
Nat. Nanotechnol.
7
,
96
(
2012
).
9.
T.
Kondo
 et al., “
Anomalous dressing of Dirac fermions in the topological surface state of Bi2 Se3, Bi2Te3, and Cu-doped Bi2 Se3
,”
Phys. Rev. Lett.
110
,
217601
(
2013
).
10.
Y.
Onishi
 et al., “
Ultrafast carrier relaxation through Auger recombination in the topological insulator Bi1.5 Sb0.5Te1.7Se1.3
,”
Phys. Rev. B
91
,
085306
(
2015
).
11.
Y. H.
Wang
,
H.
Steinberg
,
P.
Jarillo-Herrero
, and
N.
Gedik
, “
Observation of Floquet-Bloch states on the surface of a topological insulator
,”
Science
342
,
453
(
2013
).
12.
B.
Seradjeh
,
J. E.
Moore
, and
M.
Franz
, “
Exciton condensation and charge fractionalization in a topological insulator film
,”
Phys. Rev. Lett.
103
,
066402
(
2009
).
13.
S.-Y.
Xu
 et al., “
Topological phase transition and texture inversion in a tunable topological insulator
,”
Science
332
,
560
(
2011
).
14.
S.-Y.
Xu
 et al., “
Hedgehog spin texture and Berry’s phase tuning in a magnetic topological insulator
,”
Nat. Phys.
8
,
616
(
2012
).
15.
T.
Sato
 et al., “
Direct evidence for the Dirac-cone topological surface states in the ternary chalcogenide TlBiSe2
,”
Phys. Rev. Lett.
105
,
136802
(
2010
).
16.
K.
Kirshenbaum
 et al., “
Pressure-induced unconventional superconducting phase in the topological insulator Bi2 Se3
,”
Phys. Rev. Lett.
111
,
087001
(
2013
).
17.
G. A.
Thomas
 et al., “
Large electronic-density increase on cooling a layered metal: Doped Bi2Te3
,”
Phys. Rev. B
46
,
1553
(
1992
).
18.
See supplementary material at http://dx.doi.org/10.1063/1.4955188 for supplementary Figure S1 which shows infrared transmittance spectra from Bi2Te3at temperatures of 287 K and 390 K. In Figure S2, we indicate the time constant of the decrease and recovery in the mid-infrared probe reflectance spectroscopy. Figure S3 shows the infrared-to-visible reflectivity spectrum at room temperature. Figure S4 shows that the time evolution of the electron diffraction intensity from the { 1 2 ¯ 10 } planes corresponds to spot 1–6 and that the time evolution from the { 3 3 ¯ 00 } planes corresponds to spots 7 and 8. Figure S5 shows the time-resolved change in the diffuse scattering background. Figure S6 shows the time-resolved electron intensity from the ( 202 10 ¯ ) diffraction plane. Figure S7 shows atomic displacements derived from the time-resolved electron diffraction. Figure S8 shows the Raman spectrum and the coherent phonon vibrations determined from the conventional near-IR pump-probe experiment. Tables S1–S7 present the atomic coordinates obtained from the time-resolved electron diffraction.
19.
J. A.
Sobota
 et al., “
Ultrafast optical excitation of a persistent surface-state population in the topological insulator Bi2 Se3
,”
Phys. Rev. Lett.
108
,
117403
(
2012
).
20.
M.
Hajlaoui
 et al., “
Time resolved ultrafast ARPES for the study of topological insulators: The case of Bi2Te3
,”
Eur. Phys. J.: Spec. Top.
222
,
1271
(
2013
).
21.
C. W.
Luo
 et al., “
Snapshots of Dirac fermions near the Dirac point in topological insulators
,”
Nano Lett.
13
,
5797
(
2013
).
22.
M.
Hajlaoui
 et al., “
Tuning a Schottky barrier in a photoexcited topological insulator with transient Dirac cone electron-hole asymmetry
,”
Nat. Commun.
5
,
3003
(
2014
).
23.
J. A.
Sobota
 et al., “
Distinguishing bulk and surface electron-phonon coupling in the topological insulator Bi2Se3 using time-resolved photoemission spectroscopy
,”
Phys. Rev. Lett.
113
,
157401
(
2014
).
24.
K.
Norimatsu
 et al., “
Dynamics of all the Raman-active coherent phonons in Sb2Te3 revealed via transient reflectivity
,”
J. Appl. Phys.
117
,
143102
(
2015
).
25.
N.
Fukazawa
 et al., “
Time-resolved infrared vibrational spectroscopy of the photoinduced phase transition of Pd(dmit)2 salts having different orders of phase transition
,”
J. Phys. Chem. C
117
,
13187
(
2013
).
26.
M.
Hada
 et al., “
Cold ablation driven by localized forces in alkali halides
,”
Nat. Commun.
5
,
3863
(
2014
).
27.
M.
Hada
,
K.
Pichugin
, and
G.
Sciain
, “
Ultrafast structural dynamics with table top femtosecond hard x-ray and electron diffraction setups
,”
Eur. Phys. J.: Spec. Top.
222
,
1093
(
2013
).
28.
D.
Hsieh
 et al., “
Selective probing of photoinduced charge and spin dynamics in the bulk and surface of a topological insulator
,”
Phys. Rev. Lett.
107
,
077401
(
2011
).
29.
Y. D.
Glinka
 et al., “
Ultrafast carrier dynamics in thin-films of the topological insulator Bi2Se3
,”
Appl. Phys. Lett.
103
,
151903
(
2013
).
30.
P.
Blaha
,
K.
Schwarz
,
P.
Sorantin
, and
S. B.
Trickey
, “
Full-potential, linearized augmented plane wave programs for crystalline systems
,”
Comput. Phys. Commun.
59
,
399
(
1990
).
31.
S.
Nakajima
, “
The crystal structure of Bi2Te3−xSex
,”
J. Phys. Chem. Solids
24
,
479
(
1963
).
32.
H. J.
Monkhorst
and
J. D.
Pack
, “
Special points for Brillouin-zone integrations
,”
Phys. Rev. B
13
,
5188
(
1976
).
33.
H.
Cui
,
I. B.
Bhat
,
B.
O’Quinn
, and
R.
Venkatasubramanian
, “
In-situ monitoring of the growth of Bi2Te3 and Sb2Te3 films and Bi2Te3-Sb2Te3 superlattice using spectroscopic ellipsometry
,”
J. Electron. Mater.
30
,
1376
(
2001
).
34.
A.
Zimmer
 et al., “
Optical constants of electroplated Bi2Te3 films by Mueller matrix spectroscopic ellipsometry
,”
Thin Solid Films
516
,
2922
(
2008
).
35.
A. Y.
Morsy
,
S. S.
Fouad
,
E.
Hashem
, and
A. A.
El-Shazly
, “
Optical properties of thermally deposited bismuth telluride in the wavelength range of 2.5–10 μm
,”
Acta Phys. Pol., A
80
,
819
(
1991
).
36.
G.
Sciaini
and
R. J. D.
Miller
, “
Femtosecond electron diffraction: Heralding the era of atomically resolved dynamics
,”
Rep. Prog. Phys.
74
,
096101
(
2011
).
37.
A. H.
Zewail
, “
4D ultrafast electron diffraction, crystallography, and microscopy
,”
Annu. Rev. Phys. Chem.
57
,
65
(
2006
).
38.
M.
Gulde
 et al., “
Ultrafast low-energy electron diffraction in transmission resolves polymer/graphene superstructure dynamics
,”
Science
345
,
200204
(
2014
).
39.
K.
Ishizuka
and
N.
Uyeda
, “
A new theoretical and practical approach to the multislice method
,”
Acta Crystallogr., Sect. A
33
,
740
(
1977
).
40.
M.
Hada
,
K.
Okimura
, and
J.
Matsuo
, “
Characterization of structural dynamics of VO2 thin film on c-Al2O3 using in-air time-resolved x-ray diffraction
,”
Phys. Rev. B
82
,
153401
(
2010
).
41.
L.-M.
Peng
, “
Quasi-dynamical electron diffraction—A kinematic type of expression for the dynamical diffracted-beam amplitudes
,”
Acta Crystallogr., Sect. A
56
,
511
(
2000
).
42.
S.
Nie
,
X.
Wang
,
J.
Li
,
R.
Clinite
, and
J.
Cao
, “
Femtosecond electron diffraction: Direct probe of ultrafast structural dynamics in metal films
,”
Microsc. Res. Tech.
72
,
131
(
2009
).
43.
L. M.
Pavlova
,
Y. I.
Shtern
, and
R. E.
Mironov
, “
Thermal expansion of bismuth telluride
,”
High Temp.
49
,
369
(
2011
).
44.
B.-L.
Huang
and
M.
Kaviany
, “
Ab initio and molecular dynamics predictions for electron and phonon transport in bismuth telluride
,”
Phys. Rev. B
77
,
125209
(
2008
).
45.
R.
Xu
and
T.-C.
Chiang
, “
Determination of phonon dispersion relations by x-ray thermal diffuse scattering
,”
Z. Kristallogr.
220
,
1009
(
2005
).
46.
M. J.
Gutmann
,
G.
Graziano
,
S.
Mukhopadhyay
,
K.
Refson
, and
M.
von Zimmerman
, “
Computation of diffuse scattering arising from one-phonon excitations in a neutron time-of-flight single-crystal Laue diffraction experiment
,”
J. Appl. Crystallogr.
48
,
1122
(
2015
).
47.
A. I.
Frenkel
and
J. J.
Rehr
, “
Thermal expansion and x-ray-absorption fine-structure cumulants
,”
Phys. Rev. B
48
,
585
(
1993
).
48.
T.
Sakuma
 et al., “
Correlation effects among thermal displacements of atoms in KBr
,”
Solid State Ionics
192
,
54
(
2011
).
49.
G.
Moriena
,
M.
Hada
,
G.
Sciaini
,
J.
Matsuo
, and
R. J. D.
Miller
, “
Femtosecond electron diffraction: Preparation and characterization of (110)-oriented bismuth films
,”
J. Appl. Phys.
111
,
043504
(
2012
).
50.
M.
Harb
 et al., “
Excitation of longitudinal and transverse coherent acoustic phonons in nanometer free-standing films of (001) Si
,”
Phys. Rev. B
79
,
094301
(
2009
).
51.
L. W.
da Silva
and
M.
Kaviany
, “
Micro-thermoelectric cooler: Interfacial effects on thermal and electrical transport
,”
Int. J. Heat Mass Transfer
47
,
2417
(
2004
).
52.
K.
Hoefer
 et al., “
Intrinsic conduction through topological surface states of insulating Bi2Te3 epitaxial thin films
,”
Proc. Natl. Acad. Sci. U. S. A.
111
,
14979
(
2014
).
53.
R. G.
Shulman
and
K.
Knox
, “
Interactions of pσ and pπ orbitals in transition element fluorides
,”
Phys. Rev. Lett.
4
,
603
(
1960
).
54.
R. G.
Shulman
and
S.
Sugano
, “
Calculation of the crystal field splitting
,”
Phys. Rev. Lett.
7
,
157
(
1961
).
55.
J.
Shah
,
R. F.
Leheny
, and
C.
Lin
, “
Dynamic Burstein shift in GaAs
,”
Solid State Commun.
18
,
1035
(
1976
).
56.
P. W.
Juodawlkis
and
S. E.
Ralph
, “
Hole-induced transient bandgap renormalization: A mechanism for photo-induced absorption in defect-engineered semiconductors
,”
Appl. Phys. Lett.
76
,
1722
(
2000
).
57.
K. J.
Gaffney
and
H. N.
Chapman
, “
Imaging atomic structure and dynamics with ultrafast x-ray scattering
,”
Science
316
,
1444
(
2007
).
58.
R. J. D.
Miller
, “
Femtosecond crystallography with ultrabright electrons and x-rays: Capturing chemistry in action
,”
Science
343
,
1108
(
2014
).
59.
E. S.
Zijlstra
,
A.
Kalitsov
,
T.
Zier
, and
M. E.
Garcia
, “
Fractional diffusion in silicon
,”
Adv. Mater.
25
,
5605
(
2013
).
60.
V. R.
Morrison
 et al., “
A photoinduced metal-like phase of monoclinic VO2 revealed by ultrafast electron diffraction
,”
Science
346
,
445448
(
2014
).
61.
T.
Ishikawa
 et al., “
Direct observation of collective modes coupled to molecular orbital–driven charge transfer
,”
Science
350
,
1501
(
2015
).

Supplementary Material