<<

Non-equilibrium Band Broadening, Gap and Band Inversion in Black

H. Hedayat1, A. Ceraso1,2, G. Soavi3,4,S. Akhavan3,A. Cadore3,C. Dallera2, G. Cerullo1,2,A.C. Ferrari3, E. Carpene1 1Institute for Photonics and IFN-CNR, piazza Leonardo da Vinci 32, 20133 Milano, Italy 2Dipartimento di Fisica, Politecnico di Milano, 20133 Milano, Italy 3Cambridge Centre, University of Cambridge, Cambridge CB3 0FA, UK and 4Institute for Solid State , Abbe Center of Photonics, Friedrich-Schiller University Jena, 07743 Jena, Germany

Black phosphorous (BP) is a layered semiconductor with high carrier mobility, anisotropic optical response and wide bandgap tunability. In view of its application in optoelectronic devices, under- standing transient photo-induced effects is crucial. Here, we investigate by time- and angle-resolved photoemission BP in its pristine state and in the presence of Stark splitting, chemically induced by Cs ad-sorption. We show that photo-injected carriers trigger bandgap renormalization and a concurrent valence band flattening caused by Pauli blocking. In the biased sample, photo- excitation to a long-lived (ns) surface photovoltage of few hundreds mV that counterbalances the Cs-induced surface band bending. This allows us to disentangle bulk from surface electronic states and to clarify the mechanism underlying the band inversion observed in bulk samples.

Black phosphorus (BP) is a layered semiconductor with theoretical work was done on 1L- and few-layers (FL) outstanding physical properties such as high carrier mo- BP[5, 6, 10–12, 14, 18], ARPES measurements were only bility (up to∼ 104cm2V−1s−1 in the monolayer(1L))[1], performed on bulk crystals cleaved with no control large electronic/optical anisotropies (reflectance and DC on sample thickness[16, 17, 25, 28, 30]. The depth sensi- conductance can vary by a factor∼ 2 − 4 with in-plane tivity of photoemission critically depends on the orientation)[2, 3] and excellent mechanical properties energy hν (it can range from 1L at hν ∼ 100eV to several (1L-BP can sustain tensile strain up to∼30%)[4]. Its nm at hν ∼ 6eV)[31]. ARPES cannot ignore the presence direct bandgap depends on the number of layers[5, 6], of underlying bulk states, and the comparison with 1L- ranging from∼ 0.4eV in bulk[7] to ∼2eV in 1L-BP, or FL-BP theoretical predictions might be misleading. phosphorene[5], and is sensitive to pressure[8, 9], elec- Here we use TARPES with 6eV probe to in- tric field[10–13] and in-plane strain[14, 15]. Refs.16– vestigate the ultrafast response of photo-excited bulk BP, 18 demonstrated that surface doping by alkali with and without a vertical electric bias, resulting from allows to engineer the gap of BP, leading to surface Cs adsorption, to trigger Stark splitting. Photo-injected band inversion at a critical dopant concentration∼ 0.41L (e)−hole (h) pairs thermalize within a few ps (∼ 9 × 1013cm−2)[16]. This gap tunability is attributed and induce VB broadening. By comparing the dynam- to the so-called giant Stark effect[19, 20], i.e. an electric ics of pristine and alkali-adsorbed samples, we assign the field-induced shift of electronic states, named ”giant” as broadening to carriers drift-diffusion processes, and not it can to gap closure, as confirmed by modelling[6, to the previously claimed Stark-related effects[25]. Our 10, 17, 18] and photoemission experiments[17, 21], and a analysis also reveals∼ 50meV BGR entwined to a tran- pronounced surface depletion at the valence band (VB) sient VB flattening that arises from band filling. In bi- and surface confinement of the conduction band (CB). ased BP, pump photons trigger SPV of a few hundreds mV, in agreement with TARPES measurements at fixed In of the promise of BP for opto-electronic pump-probe delay[32]. We demonstrate that SPV can applications[2, 22, 23], it is important to understand its partially or totally compensate Cs-induced band bend- ultrafast non-equilibrium response. Time- and angle- ing. The bulk sensitivity of our experiment allows us to resolved photoemission spectroscopy (TARPES), exploit- disentangle bulk from surface electronic states and to es- ing a pump-probe scheme, can track the dynamics of the tablish that band inversion is a crossing between surface electronic structure after an ultrashort (tens of fs) opti- arXiv:2007.09754v1 [cond-mat.mes-hall] 19 Jul 2020 CB and bulk VB. Thus, the semiconductor-to-semimetal cal stimulus[24]. To date, only a few TARPES studies transition upon alkali surface doping can be argued only have been performed on BP. The role of photo-induced in 1L- or FL-BP[18], where bulk contributions are absent. band broadening and ionized surface impurities on car- rier dynamics was studied in Ref.25, suggesting the ab- sence of bandgap renormalization (BGR)[26, 27]. On the other hand, the VB shift triggerd by optical excitation RESULTS AND DISCUSSION was attributed to BGR and corroborated by ab-initio calculations in Ref.28, in agreement with resonant tran- BP crystallizes in the orthorhombic structure[7, 33], sient absorption measurements[29]. Although extensive where sp3 orbital hybridization leads to buckled layers 2

(c) (d)

2 (a)

A 3000

g

CB 1

= 514 nm

2000 (eV)

E 0.4 eV

F 0

gap

B E-E VB 1 2g

1000 A (001) (b)

g Intensity (a.u.) Intensity

-1

zig-zag

ZM Z

ZL ZL armchair 0

350 400 450 500

L Z M L Z

-1

Raman Shift (cm )

FIG. 1. (a) BP crystal structure. Each phosphorene layer, normal to the z-axis, is buckled along the x-axis (armchair direction). (b) First BZ of bulk BP. The direct bandgap is located at Z. (c) Schematic electronic band structure around Z. The band dispersions along the three main axes are estimated from the effective masses of Ref.7. (d) RT 514.5nm Raman spectrum of bulk BP normal to the z-axis, Fig.1a. In the bulk, the direct (∼ 10−10mbar). Photo-emitted e are detected by a hemi- bandgap is located at the Z point of the first Brillouin spherical analyzer (Phoibos 100, Specs), with a combined Zone (BZ)[7], Fig.1b, and is strongly anisotropic, Fig.1c. energy resolution∼ 45meV (as estimated from the low en- ◦ The e/h effective masses range from∼ 0.08×m0 along the ergy cut-off of the spectra) and angular acceptance±15 . x-axis (armchair direction, where m0 is the free electron All data are recorded at room temperature (RT). The mass)[7] to ∼ 1×m0 along the y-axis (zigzag direction)[7]. in-plane crystallographic orientation of the sample is de- Here, we consider the armchair (x) direction, where band termined by Low Energy Electron Diffraction and cross- dispersion is largest (smallest effective mass), but our checked by exploiting the in-plane anisotropy of the VB conclusions on band dynamics are unaffected by the in- dispersion seen by ARPES. plane crystallographic orientation because we probe the Cs deposition is performed with a SAES Getters dis- region surrounding the Z symmetry point. penser inside the preparation chamber adjacent the pho- We use bulk BP from HQ graphene. Samples are toemission one[43]. The current through the dispenser characterized by Raman spectroscopy[34–36] utilizing a (∼ 5A) is adjusted to keep∼ 10−9mbar during deposition Renishaw InVia spectrometer equipped with a 50× ob- (with starting pressure∼ 10−10mbar). The doping dose, jective (numerical aperture NA= 0.75) at 514nm. The proportional to the deposition time, is estimated compar- power is kept below 100µW to avoid any possible ing the resulting bandgap with the data in Ref.16. This damage[37]. Before measuring, the surface oxidized layer indicates that 45s correspond to the critical dose∼ 0.351L is pulled off via micromechanical cleavage by tape[38–40]. Cs, that closes the bandgap. The Raman spectrum of BP exhibits three major peaks, Fig.2a shows the RT ARPES map of bulk BP along as shown in Fig.1d. They correspond to in-plane B2g 2 1 the armchair direction (see Fig.1), following excitation and Ag and out-of-plane Ag vibrational modes[40, 41]. 1 2 by a 30fs pulse at 1.82eV. The map, recorded at posi- Our BP flakes show Ag,B2g and Ag ∼ 361.9, 439.1, and −1 tive, i.e. after pump arrival, pump-probe delay∼+2.4ps, 466.7 cm , respectively, in good agreement with bulk reveals also the normally unoccupied CB. Fig.2b re- BP literature[39–42]. ports the differential ARPES maps obtained subtract- TARPES is performed with the setup described in ing the photoemission spectrum acquired before pump Ref.43. The laser source, based on a Yb system (Pharos, arrival from spectra at two selected pump-probe delays. Light Conversion) and a non-collinear optical paramet- Red color represents photo-induced increase of spectral ric amplifier, provides ultrashort pump (hν = 1.82eV, weight, while blue represents depletion. The crosses la- pulse duration 30fs) and probe (hν = 6eV, pulse duration beled 1 to 4 in Fig.2a mark relevant features (band ex- 70fs) pulses at 80kHz repetition rate. p-polarized pump trema and band sides), and their temporal is and probe beams, focused respectively on spots∼ 135 reported in Fig.2c. While the VB (labels 1-3) shows a and ∼ 60µm in diameter, impinge at 45◦ on the sam- prompt (< 100fs) response to pump pulses, the CB (la- ple cleaved in-situ under ultrahigh vacuum conditions bel 4) has a slow build-up, on the ps scale, indicating 3

+2.4 ps +2.4 ps +50 fs

max

0.5 + 0.2

4

(c) (a) (b)

x5

0 CB 0.1

3 -

0.0

min (eV)

0.0 1 F

VB 2

2

-0.1 E-E 3 diff. intensity intensity diff. 1

4 (relative to VB peak) VB to (relative

-0.2

-0.5

-0.1 0.0 0.1 -0.1 0.0 0.1 -0.1 0.0 0.1 0 1 2 3 4 5

delay (ps)

k (1/ Å)

x

+2.4 ps VB dispersion

-0.10

(d) (e) (f) 0/Å

-0.1

VB at k = 0/ Å

x

-0.15 1.0

0.9 I/I

0

0.8

C1: BGR -0.2 0.044/ Å (eV) (eV) F

C2: BMS -0.20 F

fit E-E E-E VB int. (I/I )

0

-0.3

-0.25

-0.3 ps VB at k = 0.044/ Å

x

1.0

VB 0 ps

I/I

0.9

0 +2.4 ps (norm. EDCs)

0.8

-0.4 -0.30

0.00 0.04 0.00 0.04 0 1 2 3 4 5

delay (ps) k (1/ Å)

x

FIG. 2. (a) ARPES map of BP along the x-axis, Fig.1, at pump-probe delay +2.4ps, where also CB is populated. (b) Differential spectra at two selected delays: red represents enhancement, blue is depletion. The crosses labeled 1 to 4 mark spectral features whose dynamics is shown in panel (c) (solid lines are guides to the eye). (d) Close-up of VB maps at +2.4ps delay. EDCs are normalized to their maxima. The symbols represent the fitted band dispersions (see Methods). (e) Comparison of VB dispersions at -0.3ps (circles), 0ps (squares) and +2.4ps (triangles) delays. The black dashed line is a parabolic fit. (f) VB −1 −1 dynamics at kx = 0A˚ (open squares) and kx = 0.044A˚ (solid squares). Lines are model results from a phenomenological fit (see text), while solid circles are the VB spectral intensities integrated in the rectangular areas of panel (e). The pump fluence is 0.4mJ/cm2 (corresponding to∼ 5 × 1019cm−3 photo-injected carriers, see Methods). All measurements performed at RT that it is indirectly populated. Photo-excited e decay to their maxima in order to enhance the band shape. from higher energy levels reached by the excitation pulse Through fitting (see Methods) we quantify the band dis- towards the CB bottom via e − e[44] and e−[45] persion, shows as open symbols. Repeating the procedure scattering processes. Due to the bandgap, the enhanced at each pump-probe delay, we can track the evolution of intensity on the VB upper border (labels 2,3) is ascribed the VB structure. Fig.2e reports the VB dispersions at to photo-induced broadening and/or shift of the band three selected delays: −0.3ps (circles), 0ps (squares) and edge. Such enhancement has a longer lifetime on the VB +2.4ps (triangles). There is a∼ 20meV red-shift at zero sides (label 2) compared to the VB center (label 3). pump-probe delay and a subsequent photo-induced re- These carrier dynamics are in agreement with previous duction of the band curvature at positive delay, pointing TARPES investigations on BP[25, 28, 30]. Here, we fo- at an increased h effective mass. The latter is consistent cus on two open issues: BGR and VB broadening. Fig.2d with the persistent spectral enhancement at the VB sides is a zoom-in of the VB ARPES maps at +2.4ps delay. in Fig.2b (label 2). However, while the parabolic fit of The energy distribution curves (EDCs) are normalized the VB dispersion at negative delay is good (black dash 4 line), providing a h effective mass mVB ∼ 0.065m0 (m0 investigations[25, 28]. Our analysis confirms the presence being the free e mass) consistent with literature[7, 25], of both BGR and BMS. the curves at zero and positive delays cannot be fitted In order to understand the physics underlying the VB by simple . This suggests two alternative sce- broadening, we analyze the photo-induced response of narios: (i) loss of parabolic shape, particularly evident BP under the effect of a vertical electric bias, chemi- at positive delay, due to a flattening of the top VB (for cally induced by Cs adsorption. Since alkali atoms are −1 kx < 0.03A˚ ), rather than increase of effective mass; (ii) e donors[53], a n-type surface region forms after dop- pump pulses induce an upward shift (∼ 50meV) of the ing, leading to band bending[54]. All energy levels bend VB toward EF , signature of BGR[26, 27]. The flattening downwards when moving from bulk to surface (it would counterbalances the BGR on the very top of the VB. be upwards for acceptor surface doping[54]). Owing to Further evidence is provided by the temporal evolu- the giant Stark effect, in BP the CB bending is en- tions in Fig.2f. Open symbols represent the dynamics hanced relative to VB[11, 16, 18], eventually causing −1 of the VB at kx = 0A˚ , while solid squares refer to surface band inversion at sufficiently high Cs coverage −1 kx = 0.044A˚ (see the corresponding dash-dotted ver- (> 0.351L)[16]. Our TARPES analysis reveals that pho- tical lines in Fig.2e). A phenomenological model based toexcitation of Cs-adsorbed BP induces up to∼ 0.25V on two components, labeled C1 and C2, reproduces the SPV[32, 55]. The built-in potential generated by sur- observed behaviors. The positive component C1 (or- face doping spatially separates photo-injected e and h. ange dashed line) shifts the VB towards EF and rep- With downward band bending, e migrate to the surface, resents BGR. As photo-excited e and h rearrange in while h move towards the bulk, developing a dipole field real space, their energies in the respective bands reduce (and potential) that neutralizes the alkali-induced bend- as a consequence of screened exchange and correlation ing. Such dipole field can extend outside the sample, effects[26, 27, 46], shrinking the bandgap. This is inde- accelerating photo-emitted e (thus shifting all energy lev- pendent of kx, with a maximum amplitude∼ 47 ± 5meV, els) even if they are emitted before pump arrival, provided a rise time∼ 280 ± 100fs, and a decay time∼ 4.8 ± 0.8ps. they have not escaped the region in vacuum where the The negative component C2 (green dotted line) has a field spreads (see Methods for details). maximum amplitude∼ −36±5meV, a pulsewidth-limited Fig.3a shows three ARPES maps of BP measured af- rise time (< 100fs) and a kx-dependent decay (∼ 1ps at ter 45s Cs exposure (∼ 0.351L, see Methods) at negative −1 −1 kx = 0.044A˚ , ∼ 6ps at kx = 0A˚ ). To clarify its pump-probe delay∼ −1ps and with increasing pump flu- origin we evaluate the temporal evolution of the VB in- ence. Without pump (left map) the Cs-induced modifi- tensity by integrating the photoemission spectral weight cations can be assessed by comparing with Fig.2a. Not in the small rectangular areas in Fig.2e. The result, dis- only the top of the VB shifts down by∼ 0.3eV, but also played as green dots in Fig.2f, reproduces the VB tran- the CB minimum is now roughly touching the VB max- sient depletion (the vertical axis on the right-hand-side imum, corresponding to a downward bending∼ 0.6eV of Fig.2f shows the relative intensity I/I0, normalized to and to the apparent closure of the gap. The middle I0 at negative delay). There is a one-to-one correspon- and right maps of Fig.3a show how increasing the pump dence with the component C2. As photo-excited carriers fluence Fpump leads to a rigid, non-linear shift of the (e in CB and h in VB) thermalize, they fill states at whole ARPES spectrum to higher energy, caused by the respective band edges. h gather at the VB maxi- SPV. Also EF shifts accordingly. The saturation of SPV mum, as indicated by the slower intensity recovery at vs Fpump is shown in Fig.3b. The fluence on the top −1 kx = 0A˚ . h accumulation reduces the ARPES spec- axis is converted into photo-injected carrier density on tral weight and, consequently, the VB edge shifts down- the bottom axis (see Methods) and the data are fit- ward to higher binding energy. Our pump fluence gen- ted by a phenomenological logarithmic model[56, 57]: 19 −3 erates photo-carrier density< 10 cm , largely exceed- φSPV = αkBT/e ln(1 + neh/p0). Here, kBT = 25 meV is 16 −3 ing the equilibrium carrier concentration∼ 10 cm (see the thermal energy at RT, neh is the photo-generated car- 16 −3 Methods). The depletion of the VB top and the concur- rier density, α = 1.33±0.05 and p0 = 2.8±0.5×10 cm ring filling of the CB bottom lead to a blue-shift of the are fitting parameters (the analog of the ideality factor in optical absorption edge, known as Burstein-Moss shift a Schottky diode[57] and the equilibrium majority carrier (BMS)[47, 48] caused by Pauli blocking[49]. This was density, respectively). observed in BP, suggesting its use as active material for Transient photo-induced effects are reported in Fig.3c, mid-infrared optoelectronic devices, such as tunable in- where we compare the VB dynamics, i.e. the energy frared emitters[50] and tunable optical modulators[51]. of its maximum EV and the edge width wVB deduced BMS and BGR can have similar and opposite ampli- from the fitting procedure, before (open squares) and af- tudes for a given photo-carrier density[52], nearly can- ter (solid circles) Cs exposure, for the same excitation 2 celing each other. The coexistence of these two com- fluence∼ 0.4 mJ/cm . The evolution of EV for pris- pensating phenomena explains the diverging conclusions tine BP is the same as Fig.2f. Apart from the different on the occurrence of BGR in BP from previous TARPES binding energies, the photo-induced dynamics of the VB 5

VB dynamics 2 (-1 ps delay)

pump fluence F ( J/cm )

pump 2

2 2 2

(F = 400 J/cm )

0 J/cm 24 J/cm 485 J/cm pump max

0.5

(a) (c) (b) (eV) F

min - E - V

0.0 SPV (V) E (eV)

F SPV SPV E-E

pristine

Cs-doped VB shift (eV)

Cs-doped VB logarithmic fit -0.5

w (no pump)

-0.1 0.0 0.1 -0.1 0.0 0.1 -0.1 0.0 0.1

19 -3

delay (ps) n (x10 cm ) k (1/ Å)

eh ||

FIG. 3. (a) ARPES maps after 45s Cs exposure at three pump fluences and negative pump-probe delay∼ −1ps. The asymmetric intensity of the CB is a matrix element effect in the photoemission process[25]. The shift of the spectra is the result of SPV. (b) SPV vs photo-injected carrier density (bottom scale)/pump fluence (top scale). (c) VB maximum (top) and broadening (bottom) of pristine (open squares), and biased BP (solid circles) vs delay. (d) Scheme of photoemission process for pristine BP, (e) Cs-doped BP and (f) pumped, Cs-doped BP. The upper inset in each panel sketches the expected ARPES spectrum, −1 the lower reports the measured EDCs at k|| = 0 A˚ . φBB : band bending potential, φSPV : surface photo-voltage.

−1 maximum (upper graph) before and after Cs doping are kx = 0A˚ is shown in the lower inset. very similar. Thus, BGR and BMS appear to coexist in the biased sample. More challenging to understand is Fig.3e depicts the situation after alkali adsorption. Cs the dynamics of the edge width (lower graph). Photo- surface states are ionized donors and therefore lie above injection in pristine BP leads to a prompt VB broaden- EF . As a consequence, the n-doped surface region of ing (wVB nearly doubles within∼150fs), as for Ref.25. In the sample is characterized by the downward bending biased BP, before optical excitation, the VB width is al- potential φBB. Owing to Stark effect, the CB bending most twice the pristine case (dark yellow horizontal line is enhanced with respect to the VB one. At sufficient in Fig.3c), while after pumping (solid circles) it drops, doping (> 0.11L according to Ref.16), the minimum of following the dynamics in the unbiased sample. the CB at the surface falls below EF (green circle) and can be observed in static photoemission. Focusing on These behaviors can be understood with the help of the VB, e photoemitted from the surface (brown circle) Figs.3d-f that sketch the photoemission process under have the largest binding energy and determine the main specific situations. Fig.3d depicts the pristine BP case. peak of photoemission (see EDC in the lower inset of All energy levels (CB, VB, EF and vacuum level Evac) Fig.3e). However, with the 6eV photons used here, bulk are represented as horizontal lines since no band bending sensitivity is enhanced due to the large e mean free path is present. Probe photons hνprobe promote bound e from (∼ 10nm)[31]. Accordingly, bulk VB states (dark yel- occupied VB states to free e that can travel in vacuum low circle) contribute to the spectral weight with lower towards the analyzer where the ARPES map is recorded, binding energy. The enhanced VB width after Cs doping as sketched in the upper inset. The measured EDC at is caused by these bulk states, see upper inset of Fig3e, 6

min max CB (lin. fit)

VB onset

15 s, +1.5 ps 15 s, -0.5 ps

15 s 0.4 300

-0.5 ps

+1.5 ps

0.2

200

0.0

100

-0.2

(a) (d) (g)

0

30 s, -0.5 ps 30 s, +1.5 ps

0.4 30 s

100

-0.5 ps

+1.5 ps

(eV) 0.2

50

0.0 onset(+) E-E

-0.2

(b) (e) intensity PES (h)

0

45 s, +1.5 ps 45 s, -0.5 ps

45 s 0.4

40

-0.5 ps

+1.5 ps

0.2

20

0.0

-0.2

(c) (f) (i)

0

0 3 6 9 0 3 6 9 -0.2 0.0 0.2

emission angle (deg.) E-E (eV)

onset(+)

FIG. 4. RT ARPES maps at (a-c) −0.5ps and (d-f) +1.5ps delay for increasing Cs exposure times (15, 30, 45s, top to bottom). ◦ (g-i) Corresponding EDCs at 0 emission angle (kx = 0). In all maps, zero energy is referred to the VB onset at positive delay 2 (labeled Eonset(+)). Pump fluence is∼0.4mJ/cm marked by the vertical arrows in the EDC. pump, since e are emitted before excitation, thus pre- Fig.3f shows the pump effect on Cs-doped BP. As SPV serving the un-pumped spectral shape (as can be con- develops, all energy levels at the surface shift upward, firmed by comparing the red and dark yellow EDCs in compensating the bending potential. Here, a distinction the lower inset). Instead, at positive pump-probe de- between negative and positive pump-probe delays must lay SPV develops before photoemission, surface and bulk be made. If probe photons precede the pump pulse, e states level to similar binding energies (see upper inset) are emitted before SPV sets in. A few ps after pump- and the photoemission spectrum resembles the pristine ing, the photovoltage creates a dipole field that extends BP case with sharper VB edge, as experimentally ob- in vacuum, accelerating the traveling free e and rigidly served in the dark red EDC in the lower inset. Based shifting the photoemission spectrum. Within the exper- on these considerations, the dynamics of the VB width imental uncertainty, the corresponding EDC is identical, in Fig.3c (red curve) represents the SPV temporal on- apart from the energy shift, to that recorded without set which, in turn, reflects the dynamics of e − h space 7 separation that builds up the dipole field[58]. set of Fig.3f) and resulting in a sharper VB edge. The Having clarified the physical origin of the observed crossing of VB and CB at −0.5ps with the largest Cs spectral features, we can now explain the ubiquitous dose (Fig.4c) is the overlap of bulk VB and surface CB, photo-induced VB broadening. The close resemblance resolved thanks to the bulk sensitivity of our ARPES of the VB edge temporal evolution in pristine and Cs- photon energy. The CB binding energy and line shape doped samples at positive delay, Fig.3c (bottom), points do not significantly vary upon pumping. We only observe at a common mechanism, regardless of bias. Pump pulses a loss of spectral weight after optical excitation, due to produce identical e and h distributions that can largely photo-induced depletion. Considering that CB bending exceed the equilibrium densities (see Methods) and de- is more pronounced than VB, due to the Stark effect, cay exponentially with depth. Density gradients trigger and that SPV affects all bands equally, we should ex- carriers diffusion. In bulk BP, the mobility of h in the pect a well discernible photo-induced sharpening of CB direction normal to the surface is larger than e[7]. Since at positive delay, which is not observed. The width in- mobility is proportional to the diffusion coefficient via variance of the CB vs delay confirms its strong surface Einstein’s relation[53], h diffuse faster than e. Thus, im- confinement, in agreement with Refs.16–18, 21. Thus, mediately after pumping, h move to the bulk, leaving e our data demonstrate that band crossing in chemically at the sample surface even in the absence of a vertical biased bulk BP should be claimed with caution: only in bias. As e and h spatially separate, an outward-bound 1L or FL-BP (≤ 4 layers[18]) band inversion could be dipole field develops producing a potential, called Dem- seen, since bulk contributions are absent. ber photovoltage[55], that can alter band binding ener- gies at the surface. The build-up of this photovoltage is as fast as charge separation, and it can happen on a sub−100fs timescale (see Methods). Due to the bulk CONCLUSIONS sensitivity of our photoemission setup, we detect these alterations as band broadening. The orientation of the dipole field opposes the diffusion process. Eventually, We employed time- and angle-resolved photoemission charge separation (thus spectral broadening) will stop, spectroscopy to investigate the transient electronic dy- as for Fig.3c. Further support to this explanation is pro- namics of bulk BP. Our measurements show photo- vided by two experimental observations. (i) The Dem- induced bandgap renormalization entwined to VB flat- ber photovoltage saturates logarithmically with photo- tening, caused by Pauli blocking. Applying a vertical carrier density[55] (see Methods) and (ii) VB broadening electric bias, chemically induced by Cs ad-sorption, we in pristine BP shows the same logarithmic behavior with showed that the ubiquitous VB broadening is due to respect to pump fluence[25]. photo-carriers ambipolar drift-diffusion. Both VB and Thus, the occurrence of band inversion must be cau- (to a larger extent) CB experience surface bending upon tiously claimed in biased BP. Figs.4a-f show ARPES doping, that can be counterbalanced by a surface pho- maps recorded at negative (−0.5ps, a-c) and positive tovoltage prompted by laser irradiation. This allows us (+1.5ps, d-f) pump-probe delays for increasing Cs expo- to unequivocally discriminate bulk vs surface states, to sure times (15, 30, 45s, top to bottom). The correspond- establish the surface localization of the conduction band −1 ing EDCs at kx = 0A˚ are in Figs.4g-i. To harmonize and to elucidate the occurrence of band inversion in bulk the comparison, the zero binding energy is referred to samples. For any application of black phosphorus in- the onset of the VB at positive delay (labeled Eonset(+)). volving hot carrier injection, transient changes will affect Owing to Stark effect[16, 18], as Cs concentration in- device performance. Our analysis reveals a rich and mul- creases, the CB minimum downshifts and falls within the tifaceted photo-induced band dynamics that might help onset of the VB at negative delay for the highest doping, design opto-electronic devices. Since broad gap tunabil- reached at 45s (Fig.4c). This corresponds to the critical ity by alkali atoms ad-soprtion has been demonstrated in Cs coverage∼0.35 1L identified in Ref.16 at which VB transition-metal dichalcogenides[59, 60], our results will and CB cross. Comparing negative and positive delays, be relevant to a wider class of layered materials. we observe that, after photo excitation, the VB down- shifts, seemingly leading to a larger bandgap. Recalling Figs.3e-f, this is the effect of SPV that compensates the Cs-induced bending potential. The apparent VB shift ACKNOWLEDGEMENTS is due to its edge sharpening. Before pumping, the VB onset is determined by bulk states having smaller (less negative) binding energy as compared to surface states We acknowledge funding from EU Graphene Flagship, (see upper inset of Fig.3e). After excitation, SPV neu- ERC Grants Hetero2D, GSYNCOR, EPSRC Grants tralizes band bending, thus bulk states ”align” to surface EP/K01711X/1, EP/K017144/1, EP/N010345/1, states, appearing at larger binding energy (see upper in- EP/L016087/1, and Italian PRIN project 2017BZPKSZ. 8

+2.4 ps

-1

data (-250 fs)

VB (k = 0 Å ) (a) (c) (d) (e) x

0.4 a(E): simul. (E = -0.09 eV)

V

-250 fs

g(E): gaussian (w = 0.1 eV)

0 fs

a(E) g(E): convolution

CB

0.3

+0.29 w

VB

(arb. un.) (arb. 0.0

E

VB (norm. EDCs)

V PES intensity PES

E

C (eV) F

(b)

-0.15 E-E

-1

-0.2

CB (k > 0 Å )

x

data:

+3 ps PES intensity (arb. un.) (arb. intensity PES -250 fs un.) (arb. intensity PES

gauss fit

0 fs (arb. un.) (arb.

fit:

linear background 1st derivative derivative 1st

( gauss/ E)

( gauss/ E)

-0.4

0.00 0.03 0.06

k (1/ Å) E-E (eV) E-E (eV) E-E (eV)

x F (d) F F

−1 FIG. 5. (a) VB EDCs of pristine BP at kx = 0 A˚ for −250fs (black) and 0fs (red) pump-probe delays. (b) Numerical derivative of EDCs shown in (a): the solid lines are fits according to Eq.1. The resulting peaks are reproduced as shaded areas −1 in panel (a), with the corresponding peak positions EV and peak widths wVB . (c) EDC of the CB at kx = 0.04A˚ and +3ps delay. The solid line is a gaussian fit centered at EC . (d) ARPES maps and dispersion fits of CB (top) and VB (bottom) at 2.4ps delay. Band extrema are indicated. (e) Simulation of VB photoemission spectrum at negative delay. The spectral function a(E) is peaked at EV − EF = −0.09eV and decays exponentially at larger binding energy. Once convoluted with a gaussian profile g(E) (line width w = 0.1eV), it matches experiments.

METHODS timates of peak amplitude. Fig.5c reports the fitting rou- −1 tine for the CB (at +3ps delay and kx 6= 0 A˚ , where SPECTRAL ANALYSIS AND FITTING it is more evident). In this case, the peak is modeled by a gaussian profile with a linear background, since the To quantify band dispersions and photo-induced ef- spectral feature is well-defined. The CB binding energy fects, we employ a fitting routine of VB and CB measured is given by the position of the gaussian peak (EC ). Simi- with TARPES. larly to VB, the CB dispersion is obtained repeating the −1 fit at various kx. Fig.5a plots the EDCs of the VB at kx = 0A˚ for −250fs (black) and 0fs (red) pump-probe delays. Photo-induced depletion and shift/broadening of the ELECTRONIC AND OPTICAL PROPERTIES OF VB are seen. The peak asymmetry, caused by sec- BP ondary e (inelastic scattering events in the photoemission process[54]), makes the exact peak identification ambigu- Using the fitting routine previously described, we es- ous. Therefore, we employ a different approach. We first timate the BP bandgap E = 0.44 ± 0.01eV, with VB compute the EDCs numerical derivative, then we fit the g maximum∼ 0.15eV below E and CB minimum∼ 0.29eV resulting curves with the analytical derivative of a gaus- F above E , see Fig.5d. Data refer to positive pump- sian profile: F probe delay∼2.4ps when CB is populated. This is the

" 2 # 2 (VB)peak-to-(CB)peak energy gap. Refs.28, 62, 63 sug- − 2(E−EV )   − 2(E−EV ) ∂ w2 4A(E − EV ) w2 VB VB gest to use the (VB)onset-to-(CB)onset as definition of Ae = − 2 e ∂E wVB bandgap. This would give Eg ∼ 0.3eV. However, consid- (1) ering the combined energy-time resolution of our exper- where EV is the binding energy, wVB is the width (2σ) iments, the use of band onsets underestimates the gap. and A is the peak amplitude. Fig.5b reports the results. On the other hand, Eg = 0.44 ± 0.01eV is slightly larger The fits in the yellow-shaded region provide an accurate than the commonly reported values (0.3−0.4eV)[64]. The determination of the high-energy side of the original VB discrepancy might be caused by the kz sensitivity of pho- peaks, as testified by the reconstructed gaussian profiles toemission, related to our photon energy. Owing to the in Fig.5a (colored shaded areas), and overcome the peak strong band dispersion along the ΓZ crystallographic di- asymmetry issue. To deduce the VB dispersion, this rection (see Fig.4c), our 6eV probe photon might corre- procedure is repeated for various kx. Using the Shirley spond to a kz slightly away from the Z-point, thus de- method[61] to remove the incoherent background leads tecting a larger bandgap. We also show a VB spectral to very similar peak positions and widths, but poorer es- simulation assuming a gaussian line width w ∼ 0.1eV, 9 smaller than the measured one (wVB ∼ 0.15eV) convo- luted with a possible spectral function a(E), Fig.5e. a(E) decays exponentially at large binding energy and peaks at EV − EF = −0.09eV. Once convoluted with a gaus- sian profile, it provides an excellent fit of the experimen- tal spectrum, placing the VB maximum at EV − EF = −0.09eV instead of −0.15eV, with Eg = 0.38eV. Regardless of the exact value of Eg, equilibrium carrier concentrations in BP at RT can be estimated[65]: n0 = −(EC −EF )/kB T −(EF −EV )/kB T NC e (e in CB), p0 = NV e 2 −(EC −EV )/kB T (h in VB) and ni = n0p0 = NC NV e , 3/2 19 −3 with NC = 2.5(mC /m0) × 10 cm , NV = 3/2 19 −3 2.5(mV /m0) × 10 cm . Here, EV and EC are the energies of the VB maximum and CB minimum, mV FIG. 6. Band filling effect. (a) In equilibrium VB is fully and mC the respective effective masses and ni is the in- trinsic carrier density. Due to the very similar e and occupied and CB empty. Photo-excitation promotes e to CB leaving h in VB. (b) After intraband thermalization, the VB h effective masses in BP (mC ∼ mV ∼ 0.23m0)[7], we maximum is depleted (i.e. filled by h) and the CB minimum get N ∼ N ∼ 2.8 × 1018cm−3. Using the measured 0 C V is occupied, leading to a larger bandgap Eg defined by the bandgap and VB maximum, we estimate the equilibrium corresponding quasi-Fermi levels e and h densities (for comparison, we use values obtained from the fitting procedure and from the convolution ex- ample of Fig.5e) in TableI.

0.25

−3 −3 −3 (a) (b) Eg[eV] EV − EF [eV] ni[cm ] p0[cm ] n0[cm ] SPV

e data 0.44 ± 0.01 -0.15 4.2 × 1014 6.9 × 1015 2.6 × 1013 data exp. decay dipole 0.20 = =

0.38 -0.09 1.4 × 1015 7.7 × 1016 2.6 × 1012 kin E

0.15 TABLE I. Eg bandgap, EV − EF VB binding energy, ni in- trinsic carrier density, p0 and n0 estimated h (in VB) and e

(in CB) densities at RT. 0.10 (no pump) = = pump) (no V 16 −3 0.05

2 2 The majority carrier density p0 ∼ 10 cm agrees (t) = exp (-t/ ) (d) = (1- d d + ) 0 d

0 (t)-E

=1.2 ±0.2 ns

with that obtained from the fluence dependence of SPV V R 185 m d

0.00 (Fig.2b). It is instructive to compare this value with the E -150 -100 -50 0 50 100 0 100 200 300

traveled distance "d" ( m) photo-induced e − h density. According to the optical delay "t" (ps) properties of BP[7], with pump photon hν ∼ 1.82eV (λ ∼ 680nm) and electric field polarized along the armchair FIG. 7. Measured PES kinetic energy of VB as a function of (a) pump-probe delay and (b) distance of the photo-emitted direction, the constant√ is ε ∼ 12 + 2i, which leads to refractive index n = ε ∼ 3.5 + 0.3i, reflectivity e from sample surface. Dash line is the model from Eq.2 R = (|n−1|/|n+1|)2 ∼ 0.3 and absorption length 1/α = λ/4πIm(n) ∼ 180nm[66]. The incident pump fluence 2 Fpump[J/cm ] can be converted into density of photo- −3 generated e − h pairs neh[cm ] using the relation[66] neh = Fpump(1−R)α/hν (assuming quantum efficiency of the mild thermal carrier populations), Fig.6a. Illumina- 1). Similarly, the increase of the lattice temperature can tion promotes e to CB, leaving h in VB. As carries ther- be estimated as[66]: ∆T = Fpump(1 − R)αM/ρcp, where malize, they occupy the respective band edges, Fig.6b. ρ ∼ 2.7g/cm3[67] is the mass density, M ' 31g/mol is the molar mass and cp ' 21J/mol K is the BP lattice specific heat[67]. With pump fluence∼ 0.5mJ/cm2 we The resulting out-of-equilibrium charge distributions 19 −3 obtain neh ∼ 7 × 10 cm and ∆T ∼ 11K. Although can be defined by the so-called quasi-Fermi levels for the lattice temperature is hardly affected by the laser ir- CB (EqF C ) and VB (EqF V )[55, 68], both departing radiation, neh (photo excitation) p0 (majority carrier from the equilibrium Fermi energy EF [69]. State occu- 0 density). Therefore, with our laser fluence we are always pancy increases the optical bandgap to Eg = EqF C − in a strong electronic photo-excitation regime, and state EqF V (Burstein-Moss effect[47, 48] caused by Pauli filling effects cannot be ignored. At equilibrium, the VB blocking[70]) pushing down (up) the VB (CB) edge. This is almost fully occupied, and the CB empty (apart from explains the origin of component C2 in Fig.1f. 10

FIG. 8. (a) Simulation of e/h diffusion after optical excitation, according to Eqs.3-5. The colored surface represents [δp(z, t) − δn(z, t)]/neh as a function of delay t and depth z (blue color: p < n, red color p > n). Charge separation induces a dipole field Ed that counterbalances diffusion. (b) Surface charge imbalance at z = 0 as a function of delay at 300K (green-solid) and 1000K (orange-dashed)

SURFACE PHOTOVOLTAGE (SPV) to deduce the SPV lifetime[58]. At negative delay t < 0, e are emitted before pump arrival. They travel in vacuum 2 with kinetic energy m0v /2 (and velocity v) covering a SPV has been extensively investigated since the distance d = |vt| from the sample surface. At that point, early 1950s[71–74]. It is ubiquitous in doped the pump pulse reaches the sample and triggers SPV. semiconductors[75]. The basic ingredient is the pres- The resulting dipole field (and dipole potential) extends ence of (intrinsic or doping-induced) surface/interface in vacuum, accelerating the traveling . The gain states acting as e donors or acceptors. The consequence in kinetic energy ∆Ekin is proportional to the intensity of is the formation of a space-charge layer. The result- dipole potential φ at distance d, i.e. ∆Ekin(d) = eφ(d). ing band bending is estimated solving Poisson’s equation Therefore, mapping ∆E vs distance provides the spa- 2 + − kin ∇ φBB ∝ p−n+ND −NA (φBB is the built-in potential, tial profile of the dipole potential in front of the surface. + − p, n, ND and NA are h, e, donor and acceptor densities, Fig.7a depicts the change of VB binding energy (rela- respectively)[75]. Light absorption with photon energy tive to that measured without pump, i.e. with no SPV) larger than the gap generates e − h pairs and the electric in Cs-doped BP vs delay. It embodies the gained kinetic field in the space-charge layer spatially separates opposite energy as a function of delay. At t > 0 it mildly de- charges, leading to band flattening. The band bending is cays and an exponential fit provides the lifetime of SPV compensated by the photo-injection. This is the SPV ef- τd = 1.2 ± 0.2ns (the temporal measurement window is fect. Following Ref.55, several approaches have been de- limited by the travel range of our delay stage). Fig.7b veloped. Here, we point out some aspects involving time- reports the same data plotted as a function of traveled resolved photoemission techniques. The logarithmic de- distance d = |vt| for t < 0, and v deduced from the pendence of SPV with photo-injected carrier density was measured kinetic energy of the VB without pump. The suggested in Ref.74, and adapted to time-resolved exper- dashed red line at d > 0 is the fit with the electrostatic iments in Refs.56, 57, 76. Although some constraints ap- potential φ generated by a uniform dipole distribution on ply when dealing with time-resolved photoemission[77], a disc of radius R[77]: the technique can provide useful . (i) The SPV measured at negative pump-probe delay allows one  d  φ(d) = φ0 1 − √ (2) to retrieve the spatial dependence of the electric field d2 + R2 (and potential) outside the sample surface[77]. (ii) The band energy shift measured at positive delay allows one The resulting disc diameter 2R is∼ 3 times larger that the 11 pump spot size. This is due to the non-linear saturating from current balance[78]: behavior of SPV with light intensity. Even at the border ∂δn(z, t) of the laser spot, where laser intensity is weaker, the SPV J (z, t)/e = µ (δn + n )E + D (6) n n 0 d n ∂z might be as large as in the spot center, resulting in an ∂δp(z, t) apparent disc diameter larger than the nominal laser spot Jp(z, t)/e = µp(δp + p0)Ed − Dp (7) size. ∂z The field Ed that neutralizes diffusion corresponds to an overall vanishing current, i.e. Jn + Jp = 0, leading to: VB BROADENING IN PRISTINE BP ∂δp ∂δn ∂δp Dp ∂z − Dn ∂z (Dp − Dn) ∂z Ed = ≈ µp(δp + p0) + µn(δn + n0) (µp + µn)δp + µpp0 The pump pulse produces identical e (δn) and h (δp) (8) distributions that can largely exceed the equilibrium den- In the last term of Eq.8 we use the fact that δp ∼ δn  sities (p0, n0), especially at the surface (i.e. δp, δn  p0  n0 for a p-doped sample. When moving from bulk p0  n0 for p-doped samples), and decay exponentially to surface, the dipole field Ed builds up a potential φD, with depth. The density gradient triggers carriers diffu- called Dember photovoltage[55]: sion. For simplicity, here we will ignore recombination Z ∞ Z ∞ (D − D ) ∂δp and drift. According to Ref.7, in bulk BP h mobility p n ∂z φD = − Eddz = − dz (9) (µp) along the z-axis, normal to the surface, is higher 0 0 (µp + µn)δp + µpp0 than e (µ ). Since mobility is proportional to the diffu- D − D  µ + µ δp(0, t) n = p n ln 1 + p n (10) sion coefficient[53], h diffuse faster than e. Thus, after µp + µn µp p0 pumping, e tend to accumulate at the sample surface, 2 2 while h move to the bulk. This can be simulated via With µp = 550cm /Vs[7], µn = 400cm /Vs[7] and 2 3 diffusion equation[78]: δp/p0 ∼ 10 ÷ 10 , we obtain φD ∼ (0.8 ÷ 1.2)kBT/e. A transient photo-induced electronic T∼ 103K would re- ∂δn(z, t) ∂2δn(z, t) sult in φD ∼ 0.1V, compatible with the measured VB = Dn (3) broadening. According to Eq.10, φ also saturates loga- ∂t ∂z2 D rithmically with photo-injected carrier density, in agree- ∂δp(z, t) ∂2δp(z, t) ment with Ref.25. Thus, the photo-induced dynamics of = Dp 2 (4) ∂t ∂z the VB peak width in pristine BP can be ascribed to δp(z, 0) = δn(z, 0) = n exp(−α|z|) (5) eh transient charge separation, in with the case of doped BP. This explains the dynamics in Fig.2c. Where Dn = kBT µn/e, Dp = kBT µp/e; neh and α are the photo-excited carrier density and the pump absorp- tion coefficient, respectively. The use of the absolute value |z| in Eq.5 ensures no diffusion through the sur- face. Eqs.3-4, with initial condition given by Eq.5, can [1] J. Qiao, X. Kong, Z. X. Hu, F. Yang, W. Ji, High- be analytically solved to obtain the carrier distributions mobility transport anisotropy and linear dichroism in few-layer black phosphorus. Nat. Comm. 5, 4475 (2014). over time and depth. δr(z, t) = [δp(z, t) − δn(z, t)]/neh is relevant for us. δr > 0 indicates h excess, while δr < 0 [2] F. Xia. H, Wang, Y. Jia, Rediscovering black phosphorus as an anisotropic layered material for optoelectronics and represents e excess. The temporal and depth evolution electronics. Nat. Commun. 5, 4458 (2014). of δr(z, t) are reported in Fig.8a. [3] W. Xin, H. B. Jiang, T. Q. , X. G. Gao, S. N. Chen, e accumulate at the sample surface within a few ps B. Zhao, J. J. Yang, Z. B. Liu, J. G. Tian, C. L. Guo, (blue color, δr < 0), while h move deeper in the bulk Optical anisotropy of black phosphorus by total internal (red color, δr > 0). The estimated diffusion coefficients reflection. Nano Mater. Sci. 1 304 (2019). refer to RT. Laser pumping can induce a much higher [4] Q. Wei and X. Peng, Superior mechanical flexibility of phosphorene and few-layer black phosphorus. Appl. electronic T[79], increasing the diffusion coefficient and Phys. Lett. 104, 251915 (2014). speeding up e accumulation at the surface. Fig.8b shows [5] V. Tran, R. Soklaski, Y. Liang, L. Yang, Layer-controlled the surface charge ratio δr at z = 0 as a function of band gap and anisotropic excitons in few-layer black delay for 300K (blue) and 1000K (red). As e and h sep- phosphorus. Phys. Rev. B 89, 235319 (2014). arate, a dipole field develops. The area marked in light [6] A. N. Rudenko, S. Yuan, M. I. Katsnelson, Toward a re- blue in Fig.8a represents the surface region of the sample alistic description of multilayer black phosphorus: From with predominantly negative charge, while the light red GW approximation to large-scale tight-binding simula- tions. Phys. Rev. B 92, 085419 (2015). area refers to the region where positive charge prevails. [7] A. Morita, Semiconducting black phosphorus. Appl. The electric dipole field Ed, marked by the large black Phys. A 39, 227 (1986). arrow, arises between these two regions and counterbal- [8] Z. J. Xiang, G. J. Ye, C. Shang, B. Lei, N. Z. Wang, K. ances charge separation. We can estimate the dipole field S. Yang, D. Y. Liu, F. B. Meng, X. G. Luo, L. J. Zou, Z. 12

Sun, Y. Zhang, X. H. Chen, Pressure-Induced Electronic [24] U. Bovensiepen, H. Petek, M. Wolf (Editors), Dynam- Transition in Black Phosphorus. Phys. Rev. Lett. 115, ics at Solid State Surfaces and Interfaces: Volume 1 - 186403 (2015). Current Developments (Wiley-VCH, Weinheim 2010)- [9] R. Fei, V. Tran, L. Yang, Topologically protected Dirac [25] Z. Chen, J. Dong, E. Papalazarou, M. Marsi, C. Gior- cones in compressed bulk black phosphorus. Phys. Rev. getti, Z. Zhang, B. Tian , J.-P. Rueff, A. Taleb-Ibrahimi, B 91, 195319 (2015). L. Perfetti, Band Gap Renormalization, Carrier Multi- [10] Q. Liu, X. Zhang, L. B. Abdalla, A. Fazzio, A. Zunger, plication, and Stark Broadening in Photoexcited Black Switching a Normal Insulator into a Phosphorus. Nano Lett. 19, 488 (2019). via Electric Field with Application to Phosphorene. Nano [26] C. Klingshirn, Semiconductor Optics, ch. 21. (Springer, Lett. 15, 1222 (2014). Heidelberg, 2005). [11] Y. Liu, Z. Qiu, A. Carvahlo, Y. Bao, H. Xu, S. J. R. [27] N. Peygambarian, S. W. Koch, A. Mysyrowicz, Introduc- Tan, W. Liu, A. H. Castro Neto, K. P. Loh, J. Lu, Gate- tion to Semiconductor Optics, ch. XIII. (Prentice Hall, Tunable Giant Stark Effect in Few-Layer Black Phospho- University of Michigan, 1993). rus. Nano Lett. 17, 1970 (2017). [28] S. Roth, A. Crepaldi, M. Puppin, G. Gatti, D. Bug- [12] K. Dolui, S. Y. Quek, Quantum-confinement and Struc- ini, I. Grimaldi, T. R. Barrilot, C. A. Arrell, F. Fras- tural Anisotropy result in Electrically-Tunable Dirac setto, L. Poletto, M. Chergui, A. Marini, M. Grioni, Cone in Few-layer Black Phosphorous. Sci. Rep. 5, 11699 Photocarrier-induced band-gap renormalization and ul- (2015). trafast charge dynamics in black phosphorus. 2D Mater. [13] B. Deng, V. Tran, Y. Xie, H. Jiang, C. Li, Q. Guo, X. 6, 031001 (2019). Wang, H. Tian, S. J. Koester, H. Wang, J. J. Cha, Q. [29] X. Miao, G. Zhang, F. Wang, H. Yan, M. Li, Layer- Xia, L. Yang, F. Xia, Efficient electrical control of thin- Dependent Ultrafast Carrier and Coherent Phonon Dy- film black phosphorus bandgap. Nat. Commun. 8, 14474 namics in Black Phosphorus. Nano Lett. 18, 3053 (2018). (2017). [30] M. Nurmamat, Y. Ishida, R. Yori, K. Sumida, S. Zhu, [14] Y. Li, S. Yang, J. Li, Modulation of the Electronic Prop- M. Nakatake, Y. Ueda, M. Taniguchi, S. Shin, Y. Aka- erties of Ultrathin Black Phosphorus by Strain and Elec- hama, A. Kimura, Prolonged photo-carriers generated in trical Field. J. Phys. Chem. C 118, 23970 (2014). a massive-and-anisotropic Dirac material. Sci. Rep. 8, [15] A. S. Rodin, A. Carvalho, and A. H. Castro Neto, Strain- 9073 (2018). Induced Gap Modification in Black Phosphorus. Phys. [31] M. P. Seah, W. A. Dench, Quantitative electron spec- Rev. Lett. 112, 176801 (2014). troscopy of surfaces: A standard data for electron [16] J. Kim, S. S. Baik, S. H. Ryu, Y. Sohn, S. Park, B.-G. inelastic mean free paths in solids. Surf. and Interf. Anal- Park, J. Denlinger, Y. Yi, H.J. Choi, K. S. Kim, Observa- ysis, 1, 2 (1979). tion of tunable band gap and anisotropic Dirac semimetal [32] Z. Chen, J. Dong, C. Giorgetti, E. Papalazarou, M. state in black phosphorus. 349, 723 (2015). Marsi, Z. Zhang, B. Tian, Q. Ma, Y. Cheng, J.-P. [17] N. Ehlen, A. Sanna, B. V. Senkovskiy, L. Petaccia, A. Rueff, A. Taleb-Ibrahimi, L. Perfetti, Spectroscopy of V. Fedorov, G. Profeta, A. Gr¨uneis,Direct observation buried states in black phosphorus with surface doping. of a surface resonance state and surface band inversion 2D Mater. 7, 035027 (2020). control in black phosphorus. Phys. Rev. B 97, 045143 [33] S. Endo, Y. Akahama, S. Terada, S. Narita, Growth of (2018). Large Single Crystals of Black Phosphorus under High [18] S.-W. Kim, H. Jung, H.-J. Kim, J.-H. Choi, S.-H. Wei, J.- Pressure. Jap. J. Appl. Phys. 21, L482 (1982). H. Cho, Microscopic mechanism of the tunable band gap [34] A. C. Ferrari, D. M. Basko, Raman spectroscopy as a ver- in -doped few-layer black phosphorus. Phys. satile tool for studying the properties of graphene. Nat. Rev. B 96, 075416 (2017). Nanotech. 8, 235 (2013). [19] K. H. Khoo, M. S. C. Mazzoni, and S. G. Louie, Tuning [35] A. C. Ferrari, Raman spectroscopy of graphene and the electronic properties of boron nitride nanotubes with : Disorder, electron-phonon coupling, doping transverse electric fields: A giant dc Stark effect. Phys. and nonadiabatic effects. Solid State Commun. 143, 47 Rev. B 69, 201401(R) (2004). (2007). [20] M. Ishigami, J. D. Sau, S. Aloni, M. L. Cohen, A. Zettl, [36] J. Kim, J.-U. Lee, J. Lee, H. J. Park, Z. Lee, C. Lee, Observation of the Giant Stark Effect in Boron-Nitride H. Cheong, Anomalous dependence of Ra- Nanotubes. Phys. Rev. Lett. 94, 056804 (2005). man scattering and crystallographic orientation of black [21] B. Kiraly, E. J. Knol, K. Volckaert, D. Biswas, A. N. phosphorus. Nanoscale 7, 18708 (2015). Rudenko, D. A. Prishchenko, V. G. Mazurenko, M. I. [37] J. Lu, J. Wu, A. Carvalho, A. Ziletti, H. Liu, J. Tan, Katsnelson, P. Hofmann, D. Wegner, A. A. Khajetoori- Y- Chen, A. H. Castro Neto, B. Ozyilmaz, C. H. Sow, ans, Anisotropic Two-Dimensional Screening at the Sur- Bandgap Engineering of Phosphorene by Laser Oxida- face of Black Phosphorus. Phys. Rev. Lett. 123, 216403 tion toward Functional 2D Materials. ACS Nano 9, 10411 (2019). (2015). [22] H. Yuan, X. Liu, F. Afshinmanesh, W. Li, G. Xu, J. [38] K. S. Novoselov, D. Jiang, F. Schedin, T. J. Booth, Sun, B. Lian, A. G. Curto, G. Xe, Y. Hikita, Z. Schen, V. V. Khotkevich, S. V. Morozov, A. K. Geim, Two- S.-C. Zhang, X. Chen, M. Brongersma, H. Y. Hwang, Y. dimensional atomic crystals. PNAS 102, 10451 (2005). Cui, Highly anisotropic and robust excitons in monolayer [39] A. Favron, E. Gaufres, F. Fossard, A. L. Phaneuf-L black phosphorus. Nat. Nanotech. 10, 707 (2015). Heureux, N. Y-W. Tang, P. L. Levesque, A. Loiseau, R. [23] P. Chen, N. Li, X. Chen, W.-J. Ong, and X. Zhao, The Leonelli, S. Francoeur, R. Martel, Photooxidation and rising of 2D black phosphorus beyond graphene: syn- quantum confinement effects in exfoliated black phospho- thesis, properties and electronic applications. 2D Mater. rus. Nat. Mater. 14, 826 (2015). 5, 014002 (2018). [40] A. Castellanos-Gomez, L. Vicarelli, E. Prada, J. O. Is- 13

land, K. L. Narasimha-Acharya, S. I. Blanter, D. J. Groe- Materna, L. Krusin-Elbaum, M. Marsi, Unraveling the nendijk, M. Buscema, G. A. Steele, J. V. Alvarez, Isola- Dirac fermion dynamics of the bulk-insulating topologi- tion and characterization of few-layer black phosphorus. cal system Bi2Te2Se. Phys. Rev. Mater. 2, 104202 (2018). 2D Mater. 1, 025001 (2014). [58] S.-L. Yang, J. A. Sobota, P. S. Kirchmann, Z.-X. Shen, [41] H. B. Ribeiro, M. A. Pimenta, C. J. S. de Matos, Raman Electron propagation from a photo-excited surface: im- spectroscopy in black phosphorus. J. Raman Spectrosc. plications for time-resolved photoemission. Appl. Phys. 49, 76 (2018). A 116, 85 (2014). [42] S. Sugai, I. Shirotani, Raman and infrared reflection spec- [59] M. Kang, B. Kim, S. H. Ryu, S. W. Jung, J, Kim, L. troscopy in black phosphorus. Solid State Commun. 53, Moreschini, C. Jozwiak, E. Rotenberg, A. Bostwick, K. 753 (1985). S. Kim, Universal Mechanism of Band-Gap Engineering [43] F. Boschini, H. Hedayat, C. Dallera, P. Farinello, C. Man- in Transition-Metal Dichalcogenides. Nano Lett. 17, 1610 zoni, A. Magrez, H. Berger, G. Cerullo, E. Carpene, An (2017). innovative Yb-based ultrafast deep ultraviolet source for [60] B. S. Kim, W. S. Kyung, J. J. Seo, J. Y. Kwon, J. D. time-resolved photoemission experiments. Rev. Sci. In- Denlinger, C. Kim, S. R. Park, Possible electric field in- strum. 85, 123903 (2014) duced indirect to direct band gap transition in MoSe2. [44] Y. Shi, N. Gillgren, T. Espiritu, S. Tran, J. Yang, K. Scientific Reports 7, 5206 (2017). Watanabe, T. Taniguchi, C.-N. Lau, Weak localization [61] D. A. Shirley, High-Resolution X-Ray Photoemission and electron-electron interactions in few layer black phos- Spectrum of the Valence Bands of . Phys, Rev. B phorus devices. 2D Mater. 3, 034003 (2016). 5, 4709 (1972). [45] N. Mao, X. Wang, Y. Lin, B. G. Sumpter, Q. Ji, T. [62] E. A. Kraut, R. W. Grant, J. R. Waldrop, S. P. Kowal- Palacios, S. Huang, V. Meunier, M. S. Dresselhaus, W. czyk, Precise Determination of the Valence-Band Edge in A. Tisdale, L. Liang, X. Ling, J. Kong, Direct Observa- X-Ray Photoemission Spectra: Application to Measure- tion of Symmetry-Dependent Electron-Phonon Coupling ment of Semiconductor Interface Potentials. Phys. Rev. in Black Phosphorus. J. Am. Chem. Soc. 141, 18994, Lett. 44, 1620 (1980). (2019). [63] J. Endres, D. A. Egger, M. Kulbak, R. A. Kerner, L. [46] P. Vashishta, R. K. Kalia, Universal behavior of Zhao, S. H. Silver, G. Hodes, B. P. Rand, D. Cahen, L. exchange-correlation energy in electron-hole liquid. Phys. Kronik, A. Kahn, Valence and Conduction Band Densi- Rev. B 25, 6492 (1982). ties of States of Metal Halide Perovskites: A Combined [47] T. S. Moss, The Interpretation of the Properties of In- Experimental-Theoretical Study. J. Phys. Chem. Lett. 7, dium Antimonide. Proc. Phys. Soc. B 67, 775 (1954). 2722 (2016). [48] E. Burstein, Anomalous Optical Absorption Limit in [64] H. Liu, A. T. Neal, Z. Zhu, Z. Luo, X. Xu, D. Tom´ainek, InSb. Phys. Rev. 93, 632 (1954). D. Y. Peide, Phosphorene: an unexplored 2D semicon- [49] T. Low, A. S. Rodin, A. Carvalho, Y. Jiang, H. Wang, F. ductor with a high hole mobility. ACS Nano 8, 4033 Xia, and A. H. Castro Neto, Tunable optical properties (2014). of multilayer black phosphorus thin films, Phys. Rev. B [65] N. W. Ashcroft, N. D. Mermin, Solid State Physics 90, 075434 (2014). (Saunders College, Philadelphia, 1976). [50] W. S. Whitney,M. C. Sherrott, D. Jariwala, W.-H. Lin, [66] M. von Allmen and A.Blatter, Laser-Beam Interactions H. A. Bechtel, G. R. Rossman, and H. A. Atwater, Field with Materials (Springer, Berlin, 1995). Effect Optoelectronic Modulation of Quantum-Confined [67] C. C. Stephenson, R. L. Potter, T. G. Maple, J. C. Mor- Carriers in Black Phosphorus, Nano Lett. 17, 78-84 row, The thermodynamic properties of elementary phos- (2017). phorus The heat capacities of two crystalline modifica- [51] C. Lin, R. Grassi, T. Low, A. S. Helmy, Multilayer Black tions of red phosphorus, of α and β white phosphorus, Phosphorus as a Versatile Mid-Infrared Electro-optic Ma- and of black phosphorus from 15 to 300 K. J. Chem. terial, Nano Lett. 16, 1683 (2016). Thermodynamics 1, 59 (1969). [52] M. Feneberg, S. Osterburg, K. Lange, C. Lidig, B. Garke, [68] M. Shur, Physics of Semiconductor Devices (Prenice R. Goldhahn, E. Richter, C. Netzel, M. D. Neumann, N. Hall, New Jersey, 1990). Esser, S. Fritze, H. Witte, J. Bl¨asing,A. Dadgar, and [69] J. A. Steele, R. A. Lewis, J. Horvat, M. J. B. Nancar- A. Krost, Band gap renormalization and Burstein-Moss row, M. Henini, D. Fan, Y. I. Mazur, M. Schmidbauer, effect in - and germanium-doped wurtzite GaN up M. E. Ware, S.-Q. Yu, G. J. Salamo, Surface effects to 1020 cm−3. Phys. Rev. B 90, 075203 (2014). of vapour-liquid-solid driven Bi surface droplets formed [53] S. M. Sze, K. K. Ng, Physics of Semiconductor Devices during molecular-beam-epitaxy of GaAsBi. Sci. Rep. 6, (Wiley, Hoboken, New Jersey, 2007). 28860 (2016). [54] H. L¨uth, Solid Surfaces, Interfaces and Thin Films [70] R. N. Zitter, Saturated optical absorption through band (Springer, Berlin Heidelberg, 2001). filling in semiconductors Appl. Phys. Lett. 14, 73 (1969). [55] L. Kronik, Y. Shapira, Surface photovoltage phenomena: [71] W. H. Brattain, Evidence for Surface States on Semicon- theory, experiment, and applications. Surf. Sci. Rep, 37 ductors from Change in Contact Potential on Illumina- 1 (1999). tion. Phys. Rev. 72, 345 (1947). [56] D. Br¨ocker, T. Giessel, W. Widdra, Charge carrier dy- [72] W. H. Brattain, J. Bardeen, Surface Properties of namics at the SiO2/Si(100) surface: a time-resolved pho- Germanium- Bell System Tech. J. 32, 1 (1953). toemission study with combined laser and synchrotron [73] C. G. B. Garrett, W. H. Brattain, Physical Theory of radiation. Chem. Phys. 299, 247 (2004). Semiconductor Surfaces. Phys. Rev. 99, 376 (1955). [57] E. Papalazarou, L. Khalil, M. Caputo, L. Perfetti, N. [74] E. O. Johnson, Large-Signal Surface Photovoltage Stud- Nilforoushan, H. Deng, Z. Chen, S. Zhao, A. Taleb- ies with Germanium. Phys. Rev. 111, 153 (1952). Ibrahimi, M. Konczykowski, A. Hruban, A. Wolos, A. [75] W. M¨onch, Semiconductor Surfaces and Interfaces 14

(Springer-Verlag, Berlin, 2001). probe photoelectron spectroscopy for detecting time- [76] B. F. Spencer, D. M. Graham, S. J. O. Hardman, E. resolved surface photovoltage. J. Elec. Spec. Relat. Phe- A. Seddon, M. J. Cliffe, K. L. Syres, A. G. Thomas, S. nom. 185, 152 (2012). K. Stubbs, F. Sirotti, M. G. Silly, P. F. Kirkham, A. [78] J. P. McKelvey, Solid State and Semiconductor Physics, R. Kumarasinghe, G. J. Hirst, A. J. Moss, S. F. Hill, (Harper and Row, New York, 1966). D. A. Shaw, S. Chattopadhyay, W. R. Flavell, Time- [79] E. Carpene, Ultrafast laser irradiation of metals: Beyond resolved surface photovoltage measurements at n-type the two-temperature model. Phys. Rev. B 74, 024301 photovoltaic surfaces: Si(111) and ZnO(1010). Phys. (2006). Rev. B 88, 195301 (2013). [77] S.-I. Tanaka, Utility and constraint on the use of pump-