optical properties and photocarrier dynamics of bi o se

19
Optical Properties and Photocarrier Dynamics of Bi 2 O 2 Se Monolayer and Nanoplates Shuangyan Liu, 1, * Congwei Tan, 2, * Dawei He, 1 Yongsheng Wang, 1, Hailin Peng, 2, and Hui Zhao 3, § 1 Key Laboratory of Luminescence and Optical Information, Ministry of Education, Institute of Optoelectronic Technology, Beijing Jiaotong University, Beijing 100044, China 2 Center for Nanochemistry, Beijing National Laboratory for Molecular Sciences (BNLMS), College of Chemistry and Molecular Engineering, Peking University, Beijing 100871, China 3 Department of Physics and Astronomy, The University of Kansas, Lawrence, Kansas 66045, United States (Dated: November 11, 2019) Abstract We report a comprehensive experimental study on optical properties and photocarrier dynamics in Bi 2 O 2 Se monolayers and nanoplates. Large and uniform Bi 2 O 2 Se nanoplates with various thicknesses down to the monolayer limit were fabricated. In nanoplates, a direct optical transition near 720 nm was identified by optical transmission, photoluminescence, and transient absorption spectroscopic measurements and was attributed to the transition between the valence and conduction bands in the Γ valley. Time-resolved dif- ferential reflection measurements revealed ultrafast carrier thermalization and energy relaxation processes and a photocarrier recombination lifetime of about 200 ps in nanoplates. Furthermore, by spatially resolv- ing the dierential reflection signal, we obtained a photocarrier diusion coecient of about 4.8 cm 2 s -1 , corresponding to a mobility of about 180 cm 2 V -1 s -1 . A similar direct transition is also observed in mono- layer Bi 2 O 2 Se, suggesting that the states in the Γ valley do not change significantly with the thickness. The temporal dynamics of the excitons in the monolayer is quite dierent from the nanoplates, with a strong saturation eect and fast exciton-exciton annihilation at high densities. Spatially and temporally resolved measurements yielded an exciton diusion coecient of about 20 cm 2 s -1 . Keywords: two-dimensional material; Bi 2 O 2 Se; photocarrier; ultrafast process; diusion; transient absorption 1

Upload: others

Post on 04-Nov-2021

2 views

Category:

Documents


0 download

TRANSCRIPT

Optical Properties and Photocarrier Dynamics of Bi2O2Se Monolayer and

Nanoplates

Shuangyan Liu,1, ∗ Congwei Tan,2, ∗ Dawei He,1

Yongsheng Wang,1, † Hailin Peng,2, ‡ and Hui Zhao3, §

1Key Laboratory of Luminescence and Optical Information,

Ministry of Education, Institute of Optoelectronic Technology,

Beijing Jiaotong University, Beijing 100044, China2Center for Nanochemistry, Beijing National Laboratory for Molecular Sciences (BNLMS),

College of Chemistry and Molecular Engineering,

Peking University, Beijing 100871, China3Department of Physics and Astronomy, The University of Kansas,

Lawrence, Kansas 66045, United States

(Dated: November 11, 2019)

Abstract

We report a comprehensive experimental study on optical properties and photocarrier dynamics in

Bi2O2Se monolayers and nanoplates. Large and uniform Bi2O2Se nanoplates with various thicknesses down

to the monolayer limit were fabricated. In nanoplates, a direct optical transition near 720 nm was identified

by optical transmission, photoluminescence, and transient absorption spectroscopic measurements and was

attributed to the transition between the valence and conduction bands in the Γ valley. Time-resolved dif-

ferential reflection measurements revealed ultrafast carrier thermalization and energy relaxation processes

and a photocarrier recombination lifetime of about 200 ps in nanoplates. Furthermore, by spatially resolv-

ing the differential reflection signal, we obtained a photocarrier diffusion coefficient of about 4.8 cm2 s−1,

corresponding to a mobility of about 180 cm2 V−1 s−1. A similar direct transition is also observed in mono-

layer Bi2O2Se, suggesting that the states in the Γ valley do not change significantly with the thickness. The

temporal dynamics of the excitons in the monolayer is quite different from the nanoplates, with a strong

saturation effect and fast exciton-exciton annihilation at high densities. Spatially and temporally resolved

measurements yielded an exciton diffusion coefficient of about 20 cm2 s−1.

Keywords: two-dimensional material; Bi2O2Se; photocarrier; ultrafast process; diffusion; transient absorption

1

I. INTRODUCTION

The discovery of graphene has created a fast-growing interest in two-dimensional (2D) mate-

rials, such as transition metal dichalcogenides (TMDs) and phosphorene1. These atomically thin

materials possess several unique features that make them attractive to both fundamental research

and applications2–6. They also provide a new route to fabricate multilayer structures by stacking

together several 2D materials7. As such, it is highly desirable to discover new 2D materials with

different properties to expand the material library.

Very recently, Bi2O2Se has emerged as a new type of layered two-dimensional (2D) material

with a unique structure and novel electronic transport properties. Bi2O2Se crystals are formed by

tetragonal (BiO)n layers separated by Se atomic sheets. Instead of the van der Waals interlayer

coupling that is commonly found in layered materials, in Bi2O2Se the layered are coupled by

electrostatic forces. Bulk crystals of Bi2O2Se and Bi2O2S can be used as novel thermoelectric8–12

and photovoltaic materials13,14 with good room-temperature charge mobilities15,16. Very recently,

ultrathin and large-sized Bi2O2Se nanoplates were synthesized by epitaxy17–20. These materials

show room- and low-temperature mobilities as high as 450 and 20,000 cm2V−1S−1, respectively17,

which could be attributed to the suppression of carrier scattering by layer separation of the carriers

and the electron donors21. They also possess thickness-tunable optical bandgaps22 and low dop-

ing concentrations19 - both are attractive features for electronic and optoelectronic applications.

Indeed, several types of devices based on ultrathin Bi2O2Se have been demonstrated, including

transistors20,23,24, infrared photodetectors18,23,25–27, optical switches28, and memristor29. From a

fundamental point of view, its non-van-der-Waals interlayer coupling offers a unique platform to

study effect of interlayer interaction on properties of layered structures.

The fast process on material synthesis and device demonstration invites studies of the physical

properties of Bi2O2Se, which will provide a foundation for future development of this exotic mate-

rial. For example, the high-quality samples have allowed measurement of the electronic structures

of this material30. However, the optical properties of Bi2O2Se, especially the dynamical properties

of photocarriers, have been less studies, despite their importance for most optoelectronic devices.

Here we report steady-state optical and spatiotemporally resolved pump-probe measurements on

optical properties and photocarrier dynamics in both monolayers and nanoplates of Bi2O2Se. We

obtained important parameters describing the carrier dynamics, such as the recombination life-

times and the diffusion coefficients. These results provide basic information that is useful for

2

understanding, designing, and optimizing electronic and optoelectronic devices based on ultrathin

Bi2O2Se.

II. RESULTS AND DISCUSSION

A. Bi2O2Se Nanoplates

Nanoplates of Bi2O2Se, with a lattice structure schematically shown in Figure 1(a), were grown

on mica substrates by chemical vapor deposition (CVD). Figure 1(b) shows an optical microscopic

image of several Bi2O2Se nanoplates with a squared shape and with sizes as large as 30× 30 µm2.

An atomic force microscopic image of a representative nanoplate is shown in Figure 1(c), which

yields a thickness of about 13 nm. This corresponds to 21 layers, according to the monolayer

thickness of about 0.61 nm17. We also measured the Raman spectrum of this nanoplate with a

532-nm laser beam, as shown in Figure 1(d). The pronounced peak at 160 cm−1 is assigned to the

A1g mode, which originates from the out-of-plane vibration of Bi atoms according to previously

reported analyses12,16,31.

Previous studies have shown that the lowest states in the conduction band and the highest states

in the valence band in Bi2O2Se nanoplates are located in the X and Γ valleys, respectively17,30.

Hence, Bi2O2Se nanoplates are indirect semiconductors. The bandgap is about 0.8 eV17,30. For

optical applications, however, it is desirable to utilize direct transitions that can provide large

absorption coefficients and efficient photoluminescence (PL). To probe direct optical transitions

in Bi2O2Se nanoplates, we performed PL, transmission, and differential reflection spectroscopic

measurements. First, we obtained the PL spectrum of the sample, as shown as the red curve in Fig-

ure 2(a), under the excitation of a 532-nm continuous-wave laser. The PL peak is at about 720 nm

(1.72 eV), with a width of about 70 nm. This width is about twice larger than that of typical TMD

monolayers. We attribute this width to the inhomogeneous broadening due to defects in addition

to the homogeneous broaden during to phonons. Second, the transmission spectrum, Figure 2(b),

shows a reduction of transmission, and thus an increase of absorption, in the range of 680 - 800

nm. This absorption feature is reasonably consistent with the PL line shape. Finally, we measured

the differential reflection spectrum. A 410-nm pump pulse was used to inject photocarriers in the

Bi2O2Se nanoplate. The differential reflection of a probe pulse, defined as ∆R/R0 = (R − R0)/R0,

where R and R0 are the probe reflection with and without the presence of the pump, respectively,

3

was measured by a homemade transient absorption microscope (see Experimental Section)32 . The

peak differential reflection signal, obtained by arranging the probe pulse to arrive at the sample at

about 0.5 ps after the pump pulse, was measured as the probe wavelength was tuned. The results

are plotted in Figure 2(a) as the blue symbols. Despite of the small range of the wavelengths used

due to the limited tuning range of the instrument in this configuration, we observed a wavelength

dependence that agrees well with the PL line shape. We note that the spectrum of the differen-

tial reflection can be treated as that of transient absorption only for extremely thin films, and a

full Fresnel analysis would be required to accurately correlate the two spectra from this 13-nm

sample.33. However, several parameters needed for such an analysis are unknown for this new

material. Nevertheless, the results presented in Figure 2 establish a direct optical transition at

about 1.7 eV. By comparing with the electronic structure from first-principles calculations17, we

can assign it to the transition between the conduction-band and valence-band states in the Γ valley.

We also note that this transition was not observed in previously fabricated samples, which could

be attributed to their higher doping concentrations22.

We next used this direct transition to study photocarrier dynamics in nanoplates. Several

nanoplates were studied and similar results were obtained. Here, we focus on the 13-nm nanoplate

shown in Figure 1. Using a 720-nm probe, we measured time-dependent differential reflection

signal with various energy fluences of the 410-nm pump pulse. As shown in Figure 3(a), the sig-

nal reaches a peak shortly after the pump excitation. The rising part of the signal can be fit by

the integral of a Gaussian function with a full width at half maximum of 0.6 ps, as indicated by

the red curve. This time is slightly longer than the instrument response time of 0.4 ps. Since the

photocarriers are injected with large excess energies by the 410-nm pump pulse and are detected

by the probe that is tuned to the direct bandgap, the rising of the signal could be attributed to the

thermalization and energy relaxation of hot carriers. The peak signal is proportional to the pump

fluence, as shown in Figure 3(c). Since the injected carrier density is proportional to the pump

fluence, this result establishes the linear relation between the differential reflection signal and the

photocarrier density. We note that the highest fluence used in this study, 32.7 µJ cm−2, corresponds

to an injected areal carrier density of about 1.65 ×1012 cm−2 in the first layer of Bi2O2Se (see Ex-

perimental Section) and an average distance of 8 nm between two pairs of photocarriers. After

the peak, the signal decays exponentially. The yellow curves in Figure 3(b) are single exponential

fits. The decay time constants obtained from these fits are plotted in Figure 3(d). The weak depen-

dence of the decay constant on the pump fluence suggests that the carrier-carrier interaction has a

4

200 300 400 500 600 700

Inte

nsity

(a.u

.)

Raman Shift (cm-1)

(b)(a)

(c) (d)

0 2 4 6 8 1016

18

20

22

24

26

28

30

32

34

Hei

ght (

nm)

Distance (m)

13 nm

FIG. 1. (a) Crystalline structure of Bi2O2Se. The purple, golden, and red spheres represent Bi, O, and Se

atoms, respectively. (b) Optical microscopic image of Bi2O2Se nanoplates on a mica substrate. (c) Atomic

force microscopic image and height profile [inset of (d)] of the large nanoplate shown in (b). The thickness

of 13 nm corresponds to 21 layers. (d) Raman spectrum of the nanoplate shown in (b) measured with a

532-nm laser beam.

minor effect on the recombination dynamics. Hence, we conclude that the recombination lifetime

of the photocarriers in Bi2O2Se nanoplates is about 200 ps. This lifetime is longer than that of 2D

TMDs32.

To study the transport properties of photocarriers in Bi2O2Se nanoplates, we performed spa-

tially and temporally resolved pump-probe measurements with pump and probe laser spot sizes of

about 2 µm. The probe spot was scanned across the pump spot by tilting a mirror in the probe arm

and thus changing the incident angle of the probe to the objective lens. At each probe position,

the differential reflection signal was measured as a function of the probe delay. Figure 4(a) shows

the differential reflection signal as a function of the probe delay and the probe position, which is

defined as the position of the probe spot center with respect to the pump spot center. To analyze

5

640 660 680 700 720 740 760 780 800 8200

20

40

60

0.00.51.01.52.02.53.0

Pea

k R

/R0 (

10-3

) 0

50

100

150

PL (c

ount

s pe

r sec

ond)

T (%

)

Wavelength (nm)

(a)

(b)

FIG. 2. (a) Peak differential reflection of the 13-nm nanoplate as a function of the probe wavelength (blue

symbols, left axis) measured with a 410-nm pump and the photoluminescence spectrum of the same sam-

ple (red curve, right axis) under the 532-nm laser excitation. (b) The transmission spectrum of the same

nanoplate.

the spatiotemporal dynamics, we plot the spatial profiles at all the probe delays and fit them with

Gaussian functions. Figure 4(b) shows a few examples of the profiles and the fits. The obtained

widths of the profiles, defined as the full widths at half maximum, are squared and plotted as a

function of the probe delay in Figure 4(c). We note that the probe laser spot size does not change

with the delay, as confirmed by using the imaging system. Furthermore, since the probe delay was

scanned at each probe position, any drift of the spot during the measurement would have resulted

in asymmetric spatial profiles, instead of an artifact of broadening.

The expansion of the profile observed in Figure 4(c) is clear evidence of in-plane diffusion of

photocarriers. The tightly focused pump spot injects photocarriers with a narrow spatial distribu-

tion. The density gradient drives the diffusion of the photocarriers away from the center of the

distribution. By solving the diffusion-recombination equation, it is straightforward to show that,

for an initial Gaussian distribution, the profile remains Gaussian during the whole process with the

width increasing as34

w2(t) = w20 + 16ln(2)Dt, (1)

where w0 is the initial width, which is determined by the convolution of the pump and probe spots,

and D, the diffusion coefficient. The linear increase of the squared width is consistent with the

6

-1 0 1 2 3 4

0

1

2

3

4

5

R

/R0 (

10-3)

Probe Delay (ps)0 200 400 600 800

0

1

2

3

4

5

Pump Fluence(J cm-2)

0.8 2.0 3.3 4.9 8.210.212.316.320.424.532.7

R

/R0 (

10-3)

Probe Delay (ps)

0 5 10 15 20 25 30 35

0

1

2

3

4

5

6

Peak

R

/R0 (

10-3)

Pump Fluence (J cm-2)0 5 10 15 20 25 30 35

0

50

100

150

200

250

Dec

ay T

ime

(ps)

Pump Fluence (J cm-2)

(a) (b)

(c) (d)

FIG. 3. (a) Differential reflection signal of the 13-nm nanoplate as a function of the probe delay measured

with a 720-nm probe and 410-nm pump pulses with various fluences. (b) The same as (a) but with a larger

range of probe delays and with exponential fits (curves). (c) Peak differential reflection signal as a function

of the pump fluence. The red line is a linear fit. (d) The decay time constant deduced from the fits shown in

(b) as a function of the pump fluence.

data shown in Figure 4(c) . A linear fit, shown as the red line, gives a diffusion coefficient of

4.8 ± 0.5 cm2 s−1. The slight deviation from the linear expansion at later probe delays could be

attributed to the effect of defects on photocarrier transport. From the diffusion coefficient and

the carrier lifetime of τ = 200 ps, we can deduce a diffusion length of L =√

Dτ = 310 nm.

Furthermore, by using the Einstein’s relation, D/kBT = µ/e, where kB, e, and T = 300 K are the

Boltzmann constant, the elementary charge, and the carrier temperature, respectively, we obtained

a photocarrier mobility of about 180 cm2 V−1 s−1.

In general, photocarriers in semiconductors can exist in the forms of free electron-hole pair

and exciton - the quasiparticle formed by an electron and a hole tightly bound by their Coulomb

attraction. For a thermalized carrier system, the distribution between these two forms is determined

7

-4 -2 0 2 40

1

2

3

4

5

Probe Delay 1 ps21 ps41 ps62 ps

R

/R0

(10-3

)

Probe Position (m)-4 -2 0 2 4

60

50

40

30

20

10

0

Probe Position (m)

Prob

e D

elay

(ps)

-0.1

5.3

R/R0 (10-3)

0 10 20 30 40 50 606.1

6.2

6.3

6.4

6.5

w2 (

m2 )

Probe Delay (ps)

D = 4.8 0.5 cm2 s-1

(a) (b) (c)

FIG. 4. (a) Differential reflection signal of the 13-nm Bi2O2Se nanoplate as a function of both the probe

delay and the probe position. (b) Examples of the spatial profiles of the differential reflection signal at probe

delays as labeled in the figure. The red curves are Gaussian fits. (c) The squared width of the spatial profiles

obtained by Gaussian fits as a function of the probe delay. The linear fit, shown as the red line, gives a

diffusion coefficient of about 4.8 cm2 s−1.

by the ratio between the exciton binding energy and the thermal energy of the lattice. So far, there

have been no reports on the exciton binding energy or excitonic effects in Bi2O2Se nanoplates.

Hence, it is unclear whether the observed dynamics is dominated by the excitons or the electron-

hole pairs. If the excitonic effect is not strong in this material, the measured diffusion coefficient

describes the ambipolar diffusion of the free electron-hole pairs.

B. Bi2O2Se Monolayer

We studied monolayer Bi2O2Se samples using the similar experimental approaches. A Bi2O2Se

monolayer on a mica substrate is shown in Figure 5(a). Unlike the nanoplates that all have regular

squared shapes, the monolayer flake has an irregular shape. However, the flake is very large,

about 100 µm in length, and uniform. Atomic force microscopic measurements, shown in Figure

5(b), resulted in a thickness of 0.82 nm, confirming its monolayer nature18,20. Figure 5(c) shows

a Raman spectrum of the monolayer sample. According to theory, the Raman active modes in

monolayer Bi2O2Se include two double-degenerated Eg modes and two non-degenerated A1g and

B1g modes35. We observed five distinct Raman peaks in Figure 5(c). The peak at 160 cm−1 is

assigned to the active mode of A1g, similar to the nanoplate shown in Figure 1. The peak at 312

cm−1 could be attributed to the B1g-like mode due to the out of-plane vibration of the O atoms.

8

100 200 300 400 500 600 700 800

Inte

nsity

(a.u

.)

Raman Shift (cm-1)660 680 700 720 740 760 780 8000

50

100

150

200

Wavelength (nm)

PL (c

ount

s pe

r sec

ond)

0

2

4

6

Peak

R

/R0 (

10-4)

(c) (d)

(b)(b)

0 1 2 3 4 5 6-0.4

-0.2

0.0

0.2

0.4

0.6

Hei

ght (

nm)

Distance (m)

0.82 nm

FIG. 5. (a) Optical microscopic image of a Bi2O2Se monolayer on a mica substrate. (b) Atomic force

microscopic image and height profile [inset in (d)] showing a thickness of 0.82 nm, corresponding to a

monolayer. (c) Raman spectrum of the monolayer measured with a 532-nm laser beam. (d) Peak differential

reflection of the monolayer as a function of the probe wavelength (blue symbols, right axis) measured with

a 410-nm pump pulse and the photoluminescence spectrum (red curve, left axis) under a 532-nm laser

excitation.

However, there is a discrepancy of 42 cm−1 with the theoretical value (354.3 cm−1)35. We note that

so far, B1g-like mode of this material has not been reported. Besides, the peak at 514 cm−1 could be

assigned to the Eg-like mode, although the wavenumber is higher than the theoretical value (433.3

cm−1)35.

To identify a direct optical transition in monolayer Bi2O2Se, we performed PL measurements

with a 532-nm excitation laser. The PL spectrum, shown as the red curve in Figure 5(d), is similar

to the 13-nm nanoplate, with a similar width and a small blue shift of only a few nanometers. This

observation suggests that the electronic bandstructure in the Γ valley is not strongly affected by

the thickness. The blue symbols in Figure 5(d) show the peak differential reflection signal as a

function of the probe wavelength, measured with a 410-nm pump. The result also shows a peak at

9

about 720 nm. Due to the small absorbance of the monolayer compared to the thicker nanoplates,

we could not obtain a reliable transmission spectrum. Nevertheless, the spectra shown in Figure 5

establish a direct transition at about 720 nm, which is similar to the nanoplates.

0 100 200 300 400-1

0

1

2

3

4

5

6

7

8

R

/R0

(10-4

)

Probe Delay (ps)0 2 4 6 8 10 12

-1

0

1

2

3

4

5

6

7

8

Pump Fluence (J cm-2) 0.9 1.8 3.2 5.5

R

/R0

(10-4

)

Probe Delay (ps)1 2 3 4 5 6

2

3

4

5

6

7

Peak

R

/R0

(10-4

)

Pump Fluence (J cm-2)

(a) (b) (c)

FIG. 6. (a) Differential reflection signal of the Bi2O2Se monolayer as a function of the probe delay measured

with a 716-nm probe and a 410-nm pump with various pump fluences as labeled. (b) The same as (a) but

with a larger range of delays. (c) Peak differential reflection signal as a function of the pump fluence. The

red line is a fit with a saturation fluence of 5.6 µJ cm−2 (see text).

We next used a 716-nm probe and a 410-nm pump to study photocarrier dynamics in monolayer

Bi2O2Se. Figure 6(a) and 6(b) shows the differential reflection signal (over a short and long time

ranges, respectively) measured with different values of the pump fluence. One distinct difference

from the nanoplate results is that, there is a strong fluence dependence of the dynamics. At rela-

tively high fluences, such as 5.5 µJ cm−2 (corresponding to an areal carrier density of 2.7 × 1011

cm−2 or an average photocarrier distance of 19 nm), there is a fast-decay component of the dynam-

ics, which is absent when the density is low, as shown in Figure 6(a). Furthermore, the peak signal

is a nonlinear function of the pump fluence, with a significant saturation effect, as shown in Figure

6(c). The behavior could be described by a saturation model36 that is widely used for nonlinear

absorption, ∆R/R0 ∝ F/(F + Fsat). By a fit, shown as the red curve in Figure 6(c), we obtained a

saturation fluence of Fsat = 5.6 ± 0.6 µJ cm−2. This corresponds to a saturation density of about

2.8 × 1011 cm−2. This value is comparable to those obtained from other 2D semiconductors, such

as WS237,38. This effect can be attributed to the phase-space filling effect of the photocarriers, and

could be utilized in developing saturable absorbers. At lower densities, the fast decay channel is

absent and the decay can be described as a density-independent exponential decay process. The

orange curve in Figure 6(b) shows an example of single-exponential fits to the data in this regime,

producing a decay time constant of about 100±20 ps. This constant is assigned as the photocarrier

10

lifetime in monolayer Bi2O2Se.

It has been well established through recent studies that in 2D semiconductors, the exciton bind-

ing energies are orders-of-magnitude larger than their 3D counterparts due to the reduced dielec-

tric screening. For example, in 2D TMD semiconductors, excitons are formed from photo-excited

electron-hole pairs on a sub-picosecond time scale39,40 and are stable at room temperature4,41. The

dominance of excitons are considered a general feature of 2D semiconductors, since the reduction

of the dielectric screening is a geometric, rather than a material, effect. Hence, although there

have been no reports on excitonic effects in monolayer Bi2O2Se, it is safe to expect that at least in

this 2D limit, excitons are the dominant form of photocarriers. Thus, the behaviors summarized in

Figure 6 reflects excitonic dynamics in monolayer Bi2O2Se.

To further understand the dynamics in the high-density regime, we include a bi-molecular

exciton-exciton annihilation process. This process has been generally observed in systems

with strong interactions between excitons, such as organic crystals42–44, semiconducting carbon

nanotubes45,46, and 2D semiconductors47–51. With this process, the rate equation of the exciton

density can be written asdNdt

= −1τ

N −12γN2, (2)

where τ and γ are the exciton recombination lifetime and the exciton-exciton annihilation rate,

respectively. If the first item on the right-hand side of Eq. 2 is much smaller than the second item,

we can ignore it and find a simple solution of

N0

N(t)− 1 = γN0t, (3)

where N0 is the initially injected exciton density. This approximation is valid when γNτ/2 � 1.

To better compare the experimental results with this model, we first converted the measured

differential reflection signal to the exciton density by using the nonlinear relation represented as

the red curve in Figure 6(c). A few examples are shown in Figure 7(a). Then, the time evolution of

the quantities N0/N−1 was calculated from these curves and plotted in Figure 7(b). At early times,

this quantity indeed increases linearly, validating the approximation of ignoring the first item on

the right-hand side of Eq. 2. By linear fits, shown as the red lines, we deduce the slope, γN0, for

each value of N0, and plot it as a function of N0 in (c). By a linear fit according to the model, we

obtained an exciton-exciton annihilation rate of γ of about 0.21 cm2 s−1. Interestingly, this value

is similar to previously reported exciton-exciton annihilation rate of other 2D semiconductors,

such as MoSe247, MoS2

49, and WS251, suggesting that the enhanced exciton-exciton interaction

11

(a) (b) (c)

048

12

048

04

04

0 20 40 60 80 100 120

0

4

0

4

N0/N

-1

Probe Delay (ps)0.0 0.5 1.0 1.5 2.0 2.5

0

1

2

3

4

5

6

7

8

Slop

e (1

010 s

-1)

Injected Carrier Density (1011 cm-2)

0123

012

012

012

01

0 50 100 150 200 250 300 350 40001

5.5

Pump Fluence (J cm-2)

Probe Delay (ps)

4.5

4.1

3.6

3.2

N (1

011 c

m-2)

2.7

FIG. 7. (a) Time evolution of the exciton density with various pump fluences calculated from the time-

dependent differential reflection signals by using the relation shown as the red curve in Figure 6(c). (b)

Time evolution of the quantity, N0/N − 1, calculated from (a). (c) The slope obtained from linear fits shown

as the red lines in (b) as a function of N0. The red line is a linear fit.

in such 2D structures could be due to the geometric effect. We note that the bimolecular carrier

recombination was also observed in bulk Bi2O2Se crystals with very high carrier densities (two

orders of magnitude higher than used in this study).14.

-4 -2 0 2 480

70

60

50

40

30

20

10

0

Probe Position (m)

Prob

e D

elay

(ps)

0.0

10.9R/R0 (10-4)

-6 -4 -2 0 2 4 6

0

2

4

6

8

10

12

Probe Delay

1 ps 21 ps 41 ps 84 ps

Probe Position (m)

R

/R0 (

10-4)

0 20 40 60 80

10

11

12

13

w2 (

m2 )

Probe Delay (ps)

Equation y = a + b*xWeight InstrumentalResidual Sum of Squares

24.95656

Pearson's r 0.58727Adj. R-Square 0.28533

Value Standard Error

?$OP:A=1Intercept 10.49421 0.58446Slope 0.02444 0.01016

D = 20 10 cm2 s-1

(a) (b) (c)

FIG. 8. (a) Differential reflection signal of the monolayer Bi2O2Se sample as a function of both the probe

delay and the probe position. (b) Examples of the spatial profiles of the differential reflection signal at

several probe delays as labeled in the figure. The red curves are Gaussian fits. (c) The squared width of the

spatial profiles obtained by Gaussian fits as a function of the probe delay. The linear fit shown as the red

line gives a diffusion coefficient of about 20 cm2 s−1.

Finally, we performed spatially and temporally resolved differential reflection measurements

to study the exciton transport properties in monolayer Bi2O2Se. The results are summarized in

Figure 8. The procedure of the measurement is the same as that used for the nanoplates. However,

slightly larger laser spots of 2.6 µm were used in order to inject a lower carrier density while

12

maintaining the same optical power at the detector. The differential reflection signal as a function

of the probe delay and position is shown in Figure 8(a). Figure 8(b) shows a few examples of

the spatial profiles of the signal and the corresponding Gaussian fits. The widths of the profiles

deduced from the fits are plotted in Figure 8(c). The larger uncertainties in the width, compared to

the nanoplate measurement, are due to the lower signal and thus poor signal-to-noise ratio. Unlike

the nanoplate measurement where the signal is proportional to the carrier density and the carrier

recombination is independent of the density, here the nonlinear relation between the signal and

the exciton density as well as the exciton-exciton annihilation complicates the interpretation of the

results. For example, the increase of the width at early probe delays could have a contribution from

the faster decay of the exciton density near the spot center, where the exciton density is higher.

However, at later probe delays, the density has dropped to low values and these factors become

less important. By fitting the widths after 40 ps, we deduced an exciton diffusion coefficient of

about 20 ± 10 cm2 s−1. This value is a few times large than the nanoplates, suggesting better

transport performance of monolayer Bi2O2Se. With the same procedures used in analyzing the

nanoplate, we obtained a diffusion length of 450 nm and an exciton mobility of about 770 cm2 V−1

s−1. This value is comparable to the room-temperature charge carrier mobilities in 2D Bi2O2Se

obtained from transport measurements17,27,52 in the range of 300 - 450 cm2 V−1 s−1 . However, it

should be noted that the exciton mobility and the charge carrier mobilities are different physical

quantities that are related to various scattering mechanisms differently53.

III. CONCLUSION

We have presented a comprehensive experimental study on optical properties and photocarrier

dynamics in the newly emerged 2D semiconductor, Bi2O2Se. Large scale and uniform samples of

Bi2O2Se, including nanoplates and monolayers, were fabricated by epitaxy. For the nanoplates,

transmission, photoluminescence, and transient absorption spectroscopic measurements allowed

identification a direction optical transition near 720 nm, which is assigned to the transition in the Γ

valley. Time-resolved differential reflection measurements revealed ultrafast carrier thermalization

and energy relaxation processes, and a 200-ps photocarrier recombination lifetime. Furthermore,

by measuring the spatiotemporal dynamics of the photocarriers, we deduced a photocarrier diffu-

sion coefficient of about 4.8 cm2 s−1. For the monolayer Bi2O2Se, a similar direct transition was

observed, suggesting that the direct transition in the Γ valley does not depend strongly on the thick-

13

ness. The temporal dynamics of the excitons in monolayers shows a strong saturation effect with

a saturation density of about 2.8 ×1011 cm−2 and significant exciton-exciton annihilation effect.

Spatially and temporally resolved measurements revealed an exciton diffusion coefficient as high

as 20 cm2 s−1. The differences in the carrier dynamics in monolayer and nanoplate samples could

be attributed to their different electronic structures and the reduced dielectric screening in mono-

layers. However, further theoretical studies are necessary to fully understand such differences.

The results of this study provide basic information about the optical properties and, in particular,

dynamical properties of photocarriers in Bi2O2Se nanoplates and monolayers. The information is

useful for understanding electronic and optical properties of this new material, evaluating its suit-

ability for various optoelectronic applications, and designing optoelectronic devices with optimal

performances. For example, the strong optical absorption saturation and exciton-exciton interac-

tion in monolayer Bi2O2Se make it a promising nonlinear optical material. Its direction optical

transition in near infrared and good excitonic transport performance are also attractive for infrared

photodetection applications.

IV. EXPERIMENTAL SECTION

A. CVD Synthesis of Bi2O2Se Nanoplates on Mica

The 2D Bi2O2Se nanoplates were synthesized by using a home-made CVD system. In detail,

the tube furnace is equipped with a 12-inch-long and 30-mm-diameter quartz tube. Bi2O3 powders

(Alfa Aesar, 99.995 %) and Bi2Se3 bulks (Alfa Aesar, 99.995 %) were adopted as precursors and

placed in two quartz boat. Bi2O3 and Bi2Se3 were located in the central zone and in the upstream

with a distance of 6 cm, respectively. Then, the freshly cleaved mica substrates were placed

downstream from the hot center with a distance of 9 to 12 cm. The typical flow rate of 200 sccm

Ar was introduced as the carrier gas to keep the chamber pressure of 400 Torr with the assistance

of metal value. The temperature of central zone was heated to 615 ◦C and held for 20 min. After

that, the furnace was cooled down naturally to room temperature.

B. Steady-State Optical Measurements

Absorption spectroscopy was performed with a homemade setup. A broadband and incoherent

light beam is focused to a spot size of about 2 - 3 µm by an objective lens. The transmitted

14

or reflected light from the sample is collimated and directed to a spectrometer equipped with a

thermoelectrically cooled charge-coupled-device camera. Absorption spectra of the samples are

obtained by comparing the spectra acquired from the sample and from the substrate, respectively,

under the same conditions. The PL spectroscopy was carried out with the same setup but with a

532-nm continuous-wave laser as the excitation source.

C. Transient Absorption

Photocarrier dynamics was studied by a homemade transient absorption setup with high spatial

and temporal resolution32. An 820-nm and 100-fs pulse from a 80-MHz Ti:sapphire oscillator

is divided into two parts by a beamsplitter. One part pumps an optical parametric oscillator to

generate a visible output, serving as the probe pulse. The other part is focused to a beta barium

borate crystal to generate its second harmonic at 410 nm, used as the pump pulse. The pump

and probe pulses are combined by a beamsplitter and focused onto the sample by a microscope

objective lens. The reflected probe is collimated by the same objective lens and sent to a silicon

photodiode, which outputs a voltage signal that is proportional to the average power of the probe

reflected by the sample and received by the photodiode. A mechanical chopper is install in the

pump arm to modulate the pump power reaching the sample at about 2 KHz. Thus, the output of

the photodiode alternates between two voltages corresponding to the probe reflections with and

without the presence of the pump, respectively. By using a lock-in amplifier that is referenced

to the modulation frequency, we can measure the differential reflection, ∆R/R0. To reveal the

spatiotemporal dynamics of the photocarriers, the differential reflection is measured as a function

of the probe delay (i.e. the time lag of the probe pulse with respect to the pump pulse) by changing

the distance that the probe pulse propagates with a linear stage, and as a function of the probe spot

position with respect to the pump spot by changing the incident angle of the probe beam to the

objective lens, which is achieved by tilting a mirror in the probe arm. During the measurements,

the samples were kept at room temperature and under ambient condition.

To estimate the injected carrier density for a certain value of the pulse fluence of the 410-nm

pump, we measured the absorbance (A) of the t = 13 nm nanoplate for a normal incident 410-nm

pulse. This was achieved by comparing the reflected and transmitted powers with respect to the

incident power. We found a value of about 0.17, which corresponds to an absorption coefficient of

α = 2.303 A/t of 3 × 107 m−1. With this value, we found that a pump pulse with a peak fluence of

15

1 µJ cm−2 injects a peak volume carrier density of about 6.2 × 1017 cm−3 at the center of the spot

and at the front surface of the sample. This corresponds to an areal carrier density of 5.0 × 1010

cm−2 in the first layer of Bi2O2Se. Assuming the same absorption coefficient at 410 nm for the

monolayer, the fluence of 1 µJ cm−2 corresponds to an injected areal carrier density of the same

value.

V. ACKNOWLEDGMENTS

We are grateful for the financial support of the National Key R&D Program of China (2016

YFA0202302), the National Natural Science Foundation of China (61527817, 61875236, 61905010,

61905007), and National Science Foundation of USA (DMR-1505852).

∗ These two authors contributed equally

[email protected]

[email protected]

§ [email protected]

1 Z. Lin, A. McCreary, N. Briggs, S. Subramanian, K. H. Zhang, Y. F. Sun, X. F. Li, N. J. Borys, H. T.

Yuan, S. K. Fullerton-Shirey, A. Chernikov, H. Zhao, S. McDonnell, A. M. Lindenberg, K. Xiao, B. J.

LeRoy, M. Drndic, J. C. M. Hwang, J. Park, M. Chhowalla, R. E. Schaak, A. Javey, M. C. Hersam,

J. Robinson, and M. Terrones, 2D Mater. 3, 042001 (2016).

2 A. Splendiani, L. Sun, Y. Zhang, T. Li, J. Kim, C. Y. Chim, G. Galli, and F. Wang, Nano Lett. 10, 1271

(2010).

3 K. F. Mak, C. Lee, J. Hone, J. Shan, and T. F. Heinz, Phys. Rev. Lett. 105, 136805 (2010).

4 K. He, N. Kumar, L. Zhao, Z. Wang, K. F. Mak, H. Zhao, and J. Shan, Phys. Rev. Lett. 113, 026803

(2014).

5 N. Kumar, S. Najmaei, Q. Cui, F. Ceballos, P. M. Ajayan, J. Lou, and H. Zhao, Phys. Rev. B 87, 161403

(2013).

6 A. Autere, C. R. Ryder, A. Saynatjoki, L. Karvonen, B. Amirsolaimani, R. A. Norwood, N. Peygham-

barian, K. Kieu, H. Lipsanen, M. C. Hersam, and Z. P. Sun, J. Phys. Chem. Lett. 8, 1343 (2017).

7 A. K. Geim and I. V. Grigorieva, Nature 499, 419 (2013).

16

8 B. Zhan, Y. C. Liu, X. Tan, J. L. Lan, Y. H. Lin, and C. W. Nan, J. Am. Ceramic Soc. 98, 2465 (2015).

9 T. V. Quang and M. Kim, J. Appl. Phys. 120, 195105 (2016).

10 X. Tan, J. L. Lan, G. K. Ren, Y. C. Liu, Y. H. Lin, and C. W. Nan, J. Am. Ceramic Soc. 100, 1494

(2017).

11 J. B. Yu and Q. Sun, Appl. Phys. Lett. 112, 053901 (2018).

12 T. Cheng, C. W. Tan, S. Q. Zhang, T. Tu, H. L. Peng, and Z. R. Liu, J. Phys. Chem. C 122, 19970

(2018).

13 X. W. Zhang, B. Wang, X. H. Niu, Y. H. Li, Y. F. Chen, and J. L. Wang, Mater. Horizons 5, 1058 (2018).

14 C. H. Zhu, T. Tong, Y. J. Liu, Y. F. Meng, Z. H. Nie, X. F. Wang, Y. B. Xu, Y. Shi, R. Zhang, and F. Q.

Wang, Appl. Phys. Lett. 113, 061104 (2018).

15 C. Drasar, P. Ruleova, L. Benes, and P. Lostak, J. Electro. Mater. 41, 2317 (2012).

16 T. Tong, M. H. Zhang, Y. Q. Chen, Y. Li, L. M. Chen, J. R. Zhang, F. Q. Song, X. F. Wang, W. Q. Zou,

Y. B. Xu, and R. Zhang, Appl. Phys. Lett. 113, 072106 (2018).

17 J. X. Wu, H. T. Yuan, M. M. Meng, C. Chen, Y. Sun, Z. Y. Chen, W. H. Dang, C. W. Tan, Y. J. Liu, J. B.

Yin, Y. B. Zhou, S. Y. Huang, H. Q. Xu, Y. Cui, H. Y. Hwang, Z. F. Liu, Y. L. Chen, B. H. Yan, and

H. L. Peng, Nat. Nanotechnol. 12, 530 (2017).

18 U. Khan, Y. T. Luo, L. Tang, C. J. Teng, J. M. Liu, B. L. Liu, and H. M. Cheng, Adv. Funct. Mater. 29,

1807979 (2019).

19 J. X. Wu, C. G. Qiu, H. X. Fu, S. L. Chen, C. C. Zhang, Z. P. Dou, C. W. Tan, T. Tu, T. R. Li, Y. C.

Zhang, Z. Y. Zhang, L. M. Peng, P. Gao, B. H. Yan, and H. L. Peng, Nano Lett. 19, 197 (2019).

20 C. W. Tan, M. Tang, J. X. Wu, Y. N. Liu, T. R. Li, Y. Liang, B. Deng, Z. J. Tan, T. Tu, Y. C. Zhang,

C. Liu, J. H. Chen, Y. Wang, and H. L. Peng, Nano Letters 19, 2148 (2019).

21 H. X. Fu, J. X. Wu, H. L. Peng, and B. H. Yan, Phys. Rev. B 95, 241203 (2018).

22 J. X. Wu, C. W. Tan, Z. J. Tan, Y. J. Liu, J. B. Yin, W. H. Dang, M. Z. Wang, and H. L. Peng, Nano Lett.

17, 3021 (2017).

23 J. X. Wu, Y. J. Liu, Z. J. Tan, C. W. Tan, J. B. Yin, T. R. Li, T. Tu, and H. L. Peng, Adv. Mater. 29,

1704060 (2017).

24 R. G. Quhe, J. C. Liu, J. X. Wu, J. Yang, Y. Y. Wang, Q. H. Li, T. R. Li, Y. Guo, J. B. Yang, H. L. Peng,

M. Lei, and J. Lu, Nanoscale 11, 532 (2019).

25 J. Li, Z. X. Wang, Y. Wen, J. W. Chu, L. Yin, R. Q. Cheng, L. Lei, P. He, C. Jiang, L. P. Feng, and J. He,

Adv. Funct. Mater. 28, 1706437 (2018).

17

26 Q. D. Fu, C. Zhu, X. X. Zhao, X. L. Wang, A. Chaturvedi, C. Zhu, X. W. Wang, Q. S. Zeng, J. D. Zhou,

F. C. Liu, B. K. Tay, H. Zhang, S. J. Pennycook, and Z. Liu, Adv. Mater. 31, 1804945 (2019).

27 J. B. Yin, Z. J. Tan, H. Hong, J. X. Wu, H. T. Yuan, Y. J. Liu, C. Chen, C. W. Tan, F. R. Yao, T. R. Li,

Y. L. Chen, Z. F. Liu, K. H. Liu, and H. L. Peng, Nat. Commun. 9, 3311 (2018).

28 X. L. Tian, H. Y. Luo, R. F. Wei, C. H. Zhu, Q. Y. Guo, D. D. Yang, F. Q. Wang, J. F. Li, and J. R. Qiu,

Adv. Mater. 30, 1801021 (2018).

29 Z. Y. Zhang, T. R. Li, Y. J. Wu, Y. J. Jia, C. W. Tan, X. T. Xu, G. R. Wang, J. Lv, W. Zhang, Y. H. He,

J. Pei, C. Ma, G. Q. Li, H. Z. Xu, L. P. Shi, H. L. Peng, and H. L. Li, Adv. Mater. 31, 1805769 (2019).

30 C. Chen, M. X. Wang, J. X. Wu, H. X. Fu, H. F. Yang, Z. Tian, T. Tu, H. Peng, Y. Sun, X. Xu, J. Jiang,

N. B. M. Schroter, Y. W. Li, D. Pei, S. Liu, S. A. Ekahana, H. T. Yuan, J. M. Xue, G. Li, J. F. Jia, Z. K.

Liu, B. H. Yan, H. L. Peng, and Y. L. Chen, Sci. Adv. 4, eaat8355 (2018).

31 A. L. J. Pereira, D. Santamaria-Perez, J. Ruiz-Fuertes, F. J. Manjon, V. P. Cuenca-Gotor, R. Vilaplana,

O. Gomis, C. Popescu, A. Munoz, P. Rodriguez-Hernandez, A. Segura, L. Gracia, A. Beltran, P. Ruleova,

C. Drasar, and J. A. Sans, J. Phys. Chem. C 122, 8853 (2018).

32 F. Ceballos and H. Zhao, Adv. Funct. Mater. 27, 1604509 (2017).

33 D. J. Morrow, D. D. Kohler, Y. Zhao, S. Jin, and J. C. Wright, (2019), arXiv:1909.06445

[physics.optics].

34 D. A. Neamen, Semiconductor Physics and Devices (McGraw-Hill, Boston, 2002).

35 Y.-D. Xu, C. Wang, Y.-Y. Lv, Y. B. Chen, S.-H. Yao, and J. Zhou, RSC Adv. 9, 18042 (2019).

36 R. W. Boyd, Nonlinear Optics, 3rd ed. (Academy Press, San Diego, USA, 2008).

37 Q. Cui, Y. Li, J. Chang, H. Zhao, and C. Xu, Laser Photon. Rev. 13, 1800225 (2019).

38 S. Q. Zhao, D. W. He, J. Q. He, X. W. Zhang, L. X. Yi, Y. S. Wang, and H. Zhao, Nanoscale 10, 9538

(2018).

39 F. Ceballos, Q. Cui, M. Z. Bellus, and H. Zhao, Nanoscale 8, 11681 (2016).

40 P. Steinleitner, P. Merkl, P. Nagler, J. Mornhinweg, C. Schuller, T. Korn, A. Chernikov, and R. Huber,

Nano Lett. 17, 1455 (2017).

41 A. Chernikov, T. C. Berkelbach, H. M. Hill, A. Rigosi, Y. L. Li, O. B. Aslan, D. R. Reichman, M. S.

Hybertsen, and T. F. Heinz, Phys. Rev. Lett. 113, 076802 (2014).

42 M. D. McGehee and A. J. Heeger, Adv. Mater. 12, 1655 (2000).

43 A. Kohler, J. S. Wilson, and R. H. Friend, Adv. Mater. 14, 701 (2002).

44 A. Suna, Phys. Rev. B 1, 1716 (1970).

18

45 L. Luer, S. Hoseinkhani, D. Polli, J. Crochet, T. Hertel, and G. Lanzani, Nat. Phys. 5, 54 (2009).

46 Y. Z. Ma, L. Valkunas, S. L. Dexheimer, S. M. Bachilo, and G. R. Fleming, Phys. Rev. Lett. 94, 157402

(2005).

47 N. Kumar, Q. Cui, F. Ceballos, D. He, Y. Wang, and H. Zhao, Phys. Rev. B 89, 125427 (2014).

48 S. Mouri, Y. Miyauchi, M. Toh, W. Zhao, G. Eda, and K. Matsuda, Phys. Rev. B 90, 155449 (2014).

49 D. Sun, Y. Rao, G. A. Reider, G. Chen, Y. You, L. Brezin, A. R. Harutyunyan, and T. F. Heinz, Nano

Lett. 14, 5625 (2014).

50 S. Kar, Y. Su, R. R. Nair, and A. K. Sood, ACS Nano 9, 12004 (2015).

51 L. Yuan and L. B. Huang, Nanoscale 7, 7402 (2015).

52 Y.-Y. Lv, L. Xu, S.-T. Dong, Y.-C. Luo, Y.-Y. Zhang, Y. B. Chen, S.-H. Yao, J. Zhou, Y. Cui, S.-T. Zhang,

M.-H. Lu, and Y.-F. Chen, Phys. Rev. B 99, 195143 (2019).

53 Q. Cui, F. Ceballos, N. Kumar, and H. Zhao, ACS Nano 8, 2970 (2014).

19