Broadband Tunable Microlasers Based on ... - ACS Publications

Jan 12, 2016 - incorporating a highly polarized organic intramolecular charge-transfer (ICT) compound with a supramolecular host. The spatial confinem...
0 downloads 0 Views 1MB Size
Subscriber access provided by UNIVERSITY OF KENTUCKY

Communication

Broadband Tunable Microlasers Based on Controlled Intramolecular Charge Transfer Process in Organic Supramolecular Microcrystals Haiyun Dong, Yanhui Wei, Wei Zhang, Cong Wei, Chunhuan Zhang, Jiannian Yao, and Yong Sheng Zhao J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.5b11525 • Publication Date (Web): 12 Jan 2016 Downloaded from http://pubs.acs.org on January 13, 2016

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

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

Page 1 of 5

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

Journal of the American Chemical Society

Broadband Tunable Microlasers Based on Controlled Intramolecular Charge Transfer Process in Organic Supramolecular Microcrystals Haiyun Dong, Yanhui Wei, Wei Zhang, Cong Wei, Chunhuan Zhang, Jiannian Yao, and Yong Sheng Zhao* Beijing National Laboratory for Molecular Sciences, Key Laboratory of Photochemistry, Institute of Chemistry, Chinese Academy of Sciences, Beijing 100190, China Supporting Information Placeholder ABSTRACT: Wavelength tunable micro-/nanolasers are indispensable components for various photonic devices. In this paper, we report broadband tunable microlasers built by incorporating a highly polarized organic intramolecular charge transfer (ICT) compound with a supramolecular host. The spatial confinement of the ICT dye generates an optimized energy level system that favors controlled population distribution between the locally excited (LE) state and the twisted intramolecular charge transfer (TICT) state, which is beneficial for significantly broadening the tailorable gain region. As a result, we realized a wide tuning of lasing wavelength in the organic supramolecular microcrystals based on temperature controlled population transfer from the LE to TICT state. The results will provide a useful enlightenment for the rational design of miniaturized lasers with desired performances.

Miniaturized lasers have attracted much interest for delivering intense coherent light signals at micro-/nanoscale, offering many advanced applications ranging from on-chip optical communication to high-throughput sensing.1-3 Wavelength variability is a key requirement of micro-/nanolasers for yielding more compact and more versatile integrated photonic elements. Since the gain region is determined by the bandgap of the material, bandgap engineering with composition tunable gain media has been widely utilized to tailor the lasing wavelength.4,5 This approach permits a wide tuning range6,7 and multi-wavelength laser output,8,9 but requires sophisticated fabrication technology to control the composition. The self-absorption10,11 and Burstein-Moss effect,12 as well as mode selection based on an optical resonator with variable parameters,13,14 were developed to fulfill lasing wavelength tunability in the single-component materials. Nevertheless, most of them either suffer from large footprints and limited gain bandwidth, or preclude the possibility of dynamic and continuous tuning. The temperature dependence of semiconductor bandgap has been proved to be very effective in continuously modulating the output wavelength of a miniaturized laser.11,15 However, most materials require very large temperature variation to obtain substantial modulation due to the small temperature response of their bandgaps. Therefore, the development of gain materials that support wide tuning of gain region within small range near room temperature is crucial for achieving broadband dynamically tunable microlasers for practical photonic integration. The energy levels and excited-state processes in organic materials are usually very sensitive to the molecular environments,16,17

which shows great potential for tailoring the gain region. For example, the intramolecular charge transfer (ICT) process in highly polarized compounds can be effectively modulated by temperature variation, which determines the population distribution on the initially formed locally excited (LE) state and the low-lying twisted intramolecular charge transfer (TICT) state17. This makes ICT molecules promising candidates for dynamically tailoring the gain region. However, in general condensed states, it is difficult to achieve TICT state population because of the aggregation induced geometry restriction on the excited-state rotation and fast nonradiative decay of TICT state.17 Herein, we demonstrate the construction of optimized energy levels with two cooperative gain states using cyclodextrin inclusion of ICT molecules, where the temperature controlled population transfer from LE to TICT state enables to dynamically tune the lasing wavelength over a wide range. The well-defined platelike microstructures of ICT dye@cyclodextrin complex were synthesized with a liquid-phase self-assembly. In the supramolecular crystals, cyclodextrins provide partially restricted space that facilitates the population transfer from LE to TICT state, and lowpolarity environment that effectively reduces the nonradiative decay of TICT state, which help to build an effective level system comprising two gain states for low-threshold lasing. The luminescence of the microplates can be modulated continuously by varying the temperature, which was utilized to build wide wavelength tunable microlasers. More generally, the tunable lasing action offers an insight into the gain processes of organic materials, and provides guidance for the development of miniaturized lasers with specific functionalities. As a kind of typical ICT laser dyes, stilbazolium derivatives exhibit extraordinarily wide luminescence band that is identified as dual luminescence with a large spectral overlap due to the small energy difference.17 The two-upper-level structure is very suitable for the instantaneous tuning of the gain region. To date, most of stilbazolium based stimulated emissions were obtained in solution. In polar solvent, rapid twist relaxation from LE to TICT state occurs,17 because no potential barrier exists between the two states (Scheme 1A). The TICT state is characterized by quick nonradiative decay to ground state, resulting in high lasing threshold. Conformational restriction might raise the barrier between LE state and TICT state, which would help to decrease the threshold pump density. In condensed state, the barrier height is much higher than the thermal energy at room temperature (Scheme 1B), because of the great lattice constraint on intramolecular rotations, which makes the excited-state population com-

ACS Paragon Plus Environment

Journal of the American Chemical Society pletely trapped on the LE state. Nevertheless, stimulated emission fails to take place in stilbazolium crystals owing to the aggregation-caused quenching. Recently, it has been demonstrated that immobilizing stilbazolium dyes in the porous spaces of MOFs can minimize the quenching effect.18 In order to make full use of the two upper states, a partially restricted space is required to produce a suitable potential barrier that favors tailorable population distribution between the LE and TICT states (Scheme 1C). Here cyclodextrin inclusion strategy was applied, because the cyclodextrin can not only provide proper spatial confinement, but effectively reduce the nonradiative decay of TICT state.17 Scheme 1. Potential energy diagrams and corresponding excitedstate processes of stilbazolium compounds, (A) in solution without potential barrier, (B) in solid state with a high potential barrier, and (C) in space-confined system with a proper potential barrier between LE and TICT states.

A LE

B TICT

LE

that the CMP has regularly shaped morphology with smooth surfaces and flat end facets. This would minimize the optical scattering loss and efficiently reflect the guided emission, which is beneficial for the achievement of strong microcavity effects. These outstanding optical properties hold a promise that the CMPs would be a candidate active media for the lasing action. A β-CD

Inclusion

LE

C

19800

β-CD

13200

6600

C TICT

DASPI@β-CD

DASPI

B

0

54000

TICT

Intensity (a.u.)

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

Page 2 of 5

DASPI@β-CD

36000

18000

D

0

19500

G

Monomer

G

Aggregate

G

Confined system

Trans-4-[4-(Dimethylamino)styryl]-1-methylpyridinium iodide (DASPI, Figure 1A) was selected as the model compound due to its typical ICT property (Figure S1). β-cyclodextrin (β-CD, Figure 1A) with cavity dimension similar to the DASPI molecular size was chosen to obtain partial restriction on the rotation of the DASPI excited state. Upon the introduction of β-CD, the absorption spectra of DASPI undergo a red shift along with the enhancement of absorption intensity, leading to an isobestic point (Figure S2A), which suggests the formation of DASPI@β-CD complex. The reciprocal of the change in absorbance (1/ΔA) by the addition of β-CD varied linearly with the inverse of β-CD concentration, which reveals that the molecular ratio of DASPI to β-CD in the complex is 1:1 (Figure S2B). The 1H NMR spectra (Figure S3) of DASPI, β-CD and the complex clearly show that the DASPI molecules are successfully included in the β-CD cavities to form the DASPI@β-CD inclusion complexes and the DASPI molecules enter the β-CD cavities preferentially from the secondary hydroxyl group side with insertion of the dimethylaminostyryl group first (Figure 1A). The DASPI@β-CD inclusion complex microplates (CMPs) were fabricated with a liquid-phase self-assembly method (Supporting Information).19 The distinct differences in color and morphology between the CMPs and pure DASPI and β-CD microcrystals (Figure S4) imply the formation of DASPI@β-CD inclusion complex microcrystals. Figure 1B displays the corresponding X-ray diffraction (XRD) patterns of the CMPs and pure DASPI and β-CD microcrystals, where their comparison confirms that the as-prepared microplates are supramolecular crystals rather than simply the mixtures of DASPI and β-CD. This is further proved by the second-harmonic generation measurements (Figure S5).20 Under blue light excitation, the CMPs exhibit strong orange photoluminescence (PL, Figure 1C), which is in sharp contrast to the red PL of DASPI microcrystals (Figure S6A). The CMPs show only monomer fluorescence (Figure S6B), which indicates that the cyclodextrin inclusion effectively restricts the intermolecular interaction. Accordingly, the CMPs possess excellent optical properties with fluorescence quantum yield of 10.2 %, much higher than that of the DASPI single crystal (1.4 %, Figure S7), which is essential for the optical gain and amplification. The scanning electron microscopy (SEM) image (Figure 1D) shows

DASPI

13000

6500

0

10 10

20 2θ (degree) 20

30 30

Figure 1. (A) Schematic illustration for the inclusion complex formation of DASPI and β-CD. (B) XRD patterns of the CMPs, and pure DASPI and β-CD microcrystals. (C) PL microscope image of ensemble CMPs excited with the blue band (450-490 nm) of a mercury lamp. Scale bar is 20 μm. (D) SEM image of a typical CMP with smooth faces. Scale bar is 5 μm. Optically pumped lasing measurements were carried out on a home-built far-field microphotoluminescence system (Figure S8). A pulsed laser beam (465 nm, 150 fs, 1 kHz), with the beam waist adjusted to be larger than the length of each microplate, was used as a nearly uniform pump source. Figure 2A summarizes the PL spectra of an isolated CMP as a function of pump fluence. At low pump energy with fluence less than 300 nJ/cm2, the PL spectra are dominated by broad spontaneous emissions. When the pump fluence exceeds a threshold, strong laser emission develops as a set of sharp peaks. The full width at half maximum (FWHM) dramatically decreases down to ∼0.9 nm above the threshold. The plot of lasing intensity versus pump fluence exhibits an S-shaped nonlinear behavior characteristic of the transition from spontaneous emission via amplified spontaneous emission, to full lasing oscillation. The optical images of the CMP were recorded below and above the lasing threshold (Figure 2A, inset).Below the threshold, the CMP only exhibits weak PL, while two impressive yellow spots emerge on the two facets along the length of the microplate above threshold. This indicates that the microplate constitutes a Fabry– Pérot (FP) optical resonator with the two facets functioning as reflecting mirrors (Figure S9).5,21 To study the microcavity effects, the lasing spectra of CMPs with different lengths were measured (Figure 2B). For the FP-type resonance, the mode spacing is given by the equation Δλ = λ2/2Lng, where L is the length of the microplate resonator, λ is the light wavelength, and ng is the group refractive index. The group refractive index ng at 586 nm was calculated to be ∼5.56 (Figure S10), which is high enough to induce the tight confinement of the FP-type modes in the CMPs.22 Figure 2C plots the mode spacing Δλ at 586 nm against 1/L of the

ACS Paragon Plus Environment

200

520 520

B

540 540

560 560

L = 15.4 μm

2.0

Intensity (a.u.)

1.5

1.5

1

1000

1.0

1.0

600

610

C

600 700 Wavelength (nm)

640

620

640

660 660

3.0 2.0

0.08 1/L (μm-1)

102 100

670 nm 610 nm 550 nm

101 10

2 2

4 6 8 10 12 Decay time (ns) 4

6

kL-T ≈ kL

8

10

12

kL-T » kL

LE

TICT

TICT Cooling

TICT Heating

kL

G

G

Figure 3. (A) Fluorescence spectra of DASPI solutions with various concentrations of β-CD. The concentration of DASPI is kept constant at 0.01 mM. (B) PL decay profiles of the CMPs at different wavelengths, which were best fitted with two exponential components. (C) Temperature controlled ICT process in the dualupper-level lasing system.

λ = 586 nm

0.04

1000

LE

G

C 4.0

103

800

kL-T « kL

LE

kL

1.0 580 590 600 Wavelength (nm)

TICT 500

800

L = 22.0 μm 570

15.0 5.0 2.0 0.5 0.2 0.0

200 400 800 Pump fluence (nJ/cm2) 400

B

cβ-CD (mM)

LE

800

2.0

200

600

L = 8.65 μm

600

10000

580 600 620 Wavelength (nm) 580

400

10

Δλ (nm)

Intensity (a.u.)

Pump fluence (nJ/cm2) 500 440 390 340 300 240 170

FWHM (nm)

A

A

Intensity (counts)

microplates, showing clearly a linear relationship, which indicates that the PL modulation is due to the FP-type cavity resonance. Furthermore, the simulation of electric field intensity distribution inside the microplate confirms the existence of the FP-mode along the length direction of the CMP (Figure S11).

Intensity (a.u.)

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

Journal of the American Chemical Society

PL Intensity (a.u.)

Page 3 of 5

0.12

Figure 2. (A) PL spectra of an individual CMP pumped with different laser energies. Left inset: Optical images of the CMP below and above lasing threshold. Scale bar is 10 μm. Right inset: Power dependent profiles of the PL peak intensity and FWHM. (B) Modulated lasing spectra CMPs with different lengths. (C) The plot and fitted curve of the mode spacing measured at 586 nm versus reciprocal microplate length. Steady and transient spectra were measured to probe the lasing mechanism in the dual-upper-level structure in the CMPs. Figure 3A presents the β-CD concentration dependent luminescence spectra of DASPI. Upon the addition of β-CD, the fluorescence spectra of DASPI experience a blue shift accompanied by a strong enhancement of emission intensity with luminescence efficiency increasing from 0.6 to 1.32 % (Figure S12A). This indicates that the steric restriction of the interior of the β-CD cavity introduces an energy barrier between the two upper levels, which restrains the transition from the LE to the TICT state, and therefore strongly enhances the LE emission at short wavelength.23 The PL FWHM increases first and then decreases with the increase of the β-CD concentration (Figure S12B), which clearly illustrates the evolution from TICT emission to LE emission.17 These results reveal that the supramolecular confinement enables the population on LE state in the DASPI@β-CD complex. The CMPs show a smooth PL spectrum without vibronic fine structure, implying the ICT character of the luminescence. The luminescence decays of the supramolecular microcrystals were measured at different wavelengths (Figure 3B). The PL decay profiles were best fitted with two exponential components listed in Table S1, which means the luminescence of the supramolecular microcrystals is originated from two excited states, namely the LE and TICT states, respectively. This reveals that the constructed energy barrier allows a part of the DASPI molecules to relax from the LE to TICT state at room temperature. In addition, the relatively low polarity of the β-CD cavity remarkably reduces nonradiative transition of TICT state.17 All of these characterizations demonstrate that the supramolecular confinement endows the DASPI with optimized energy levels that supports the simultaneous population and radiative transition of the LE and TICT states. The laser generation in the CMPs is achieved through the cooperative light amplification from the two upper states.

The brand new lasing mechanism shows great potential for building widely tunable lasers, because the luminescence wavelength of ICT compounds is determined by the dynamic equilibrium between the LE and TICT states.24 Figure 3C depicts the temperature-dependent excited-state processes of the CMPs, where kL is radiative rate constant of LE state and kL-T represents the rate constant of the twisted transition from the LE to TICT state. The initially formed LE state is preferentially occupied with decreasing temperature; thereby a blue shift of the luminescence is expected. Upon heating, the intramolecular thermal motion that crosses the energy barrier will be enhanced, which would increase the population on the low-lying TICT state at the expense of that on LE state, resulting in a bathochromic luminescence. Accordingly, the luminescence wavelength and gain region of the CMPs can be controlled by changing temperature. Exactly, this is what we have observed from the experiments. Figure 4A reveals that with temperature increasing from -60 to 120 oC, the PL images of a typical CMP exhibit a gradual color change from yellow to red, and the PL spectra gradually shift to lower energy. Figure 4B plots the corresponding luminescence peak as a function of temperature. At low temperature range, the fluorescence wavelength shows almost no change with temperature variation, indicating that the twist relaxation from the LE to TICT state is completely inhibited (kL-T is far less than kL), and fluorescence originates from pure LE state. Elevating temperature from 0 to 90 oC leads to a sharp increase of emission wavelength, which can be attributed to the increase of TICT emission with the loss of LE emission. With the further increase of temperature above 90 oC, the spectral shift levels off and reaches a wavelength limit at 622 nm. This suggests that the population on LE state totally transfers to TICT state (kL-T becomes much larger than kL), and the fluorescence is mainly from the transition of TICT state. The temperature-dependent PL spectra prove the feasibility of tuning the gain region in the CMPs, which can be utilized to build wide wavelength tunable lasers. As shown in Figure 4C, with the temperature increasing from -60 to 120 oC, the lasing wavelength is tuned continuously from 570 nm to 620 nm, which undoubtedly verifies the cooperative light amplification process from the two upper states. The temperature sweep was conducted at a fixed pump power (2200 nJ/cm2) higher than the threshold for the entire temperature range (Figure S13). The lasing FWHM is broadened at high temperature, which can be attributed to the reduction of

ACS Paragon Plus Environment

Journal of the American Chemical Society resonator quality factor due to the weak light confinement resulting from the low refractive index of the CMP at long wavelength and the heating induced cavity loss. The wavelength-temperature relationship (Figure 4D) indicates that the lasing central wavelength can be continuously and precisely modulated over a range as broad as 50 nm, which is so far the widest real-time tunable laser range in visible region to the best of our knowledge. The lasing wavelength shows a very large temperature response coefficient, ∼0.43 nm/oC from 0 to 100 oC. In addition, this shift is fully reversible, and the lasing wavelength returns to its initial value by decreasing the temperature back to -60 oC. PL Intensity (a.u.)

6

4

B

120 oC

Wavelength (nm)

A

70 oC 55 oC 30 oC

2

10 oC -60 oC

output wavelength of the supramolecular microlaser. The results demonstrate a unique lasing mechanism in organic materials for the construction of novel miniaturized lasers.

ASSOCIATED CONTENT Supporting Information Experimental details and additional data. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR INFORMATION Corresponding Author

620 620

LE+TICT

600

[email protected]

600

LE

TICT

Notes The authors declare no competing financial interests.

580 580

0

500 500

6

4

600 700 Wavelength (nm) 600

700

800

85 oC 60 oC 45 oC

2

25 oC oC

550

0 60 120 Temperature (oC) 0

60

120

600 Wavelength (nm) 600

650 650

ACKNOWLEDGMENT This work was supported by the National Natural Science Foundation of China (21125315, 21533013, 21221002), the Ministry of Science and Technology of China (2012YQ120060), and the Strategic Priority Research Program of the Chinese Academy of Sciences (XDB12020300).

615 600 585

REFERENCES

570

0

550

-60

D

120 oC

-60

-60

800

Wavelength (nm)

C

PL Intensity (a.u.)

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

Page 4 of 5

-60

0 60 120 Temperature (oC)

Figure 4. (A) Temperature-dependent fluorescence spectra of the CMPs. Inset: The real-color PL images of a single CMP at increasing temperatures. (B) Maximum emission wavelength versus temperature showing three distinct emission regions: pure LE state emission, LE+TICT emission, and TICT state emission. (C) Temperature dependent lasing spectra of a single CMP. (D) Lasing central wavelength versus temperature, spanning a range of 50 nm continuously. At low temperature, with the inhibition of TICT process, the population inversion occurs only between the initially populated LE level and the lower levels (vibronic sublevels of the ground state), which results in the short wavelength lasing output. Upon elevating the temperature, a part of molecules undergo a heat induced twist relaxation to TICT states. Therefore, the population inversion was achieved simultaneously between the two upper levels and the lower levels, which can jointly amplify the light with wavelength located in gain range overlap of the two states. With further increase of temperature, the population inversion density between the TICT state and the ground states increases at the expense of that of the LE state, causing a remarkable red shift of lasing wavelength. When all the LE states are transformed to TICT states before the radiative transition, the lasing wavelength reaches a maximum. The large gain region overlap of the two excited states enables to tune lasing central wavelength continuously. The tunable lasing action, in turn, further confirmed the coexistence and cooperative gain process of the two upper states. In summary, we reported that broadband tunable microlasers can be realized based on controlled intramolecular charge transfer process in inclusion complex crystals. By incorporating ICT molecule into β-CD, an optimized lasing level system with two cooperative gain states was obtained because the β-CD host provided partially restricted space that facilitates the population transfer from the LE to TICT state. The low-polarity environment of the β-CD cavity effectively reduced the nonradiative decay of TICT state; and therefore, the band of the optical gain can be shifted on the basis of temperature controlled population distribution between LE and TICT states, which enabled a broad tuning of the

(1) Yan, R.; Gargas, D.; Yang, P. Nat. Photonics 2009, 3, 569. (2) Duan, X.; Huang, Y.; Agarwal, R.; Lieber, C. M. Nature 2003, 421, 241. (3) Zhang, C.; Zou, C.-L.; Zhao, Y.; Dong, C.-H.; Wei, C.; Wang, H.; Liu, Y.; Guo, G.-C.; Yao, J.; Zhao, Y. S. Sci. Adv. 2015, 1, e1500257. (4) Liu, Z.; Yin, L.; Ning, H.; Yang, Z.; Tong, L.; Ning, C.-Z. Nano Lett. 2013, 13, 4945. (5) Xing, J.; Liu, X. F.; Zhang, Q.; Ha, S. T.; Yuan, Y. W.; Shen, C.; Sum, T. C.; Xiong, Q. Nano Lett. 2015, 15, 4571. (6) Pan, A.; Zhou, W.; Leong, E. S. P.; Liu, R.; Chin, A. H.; Zou, B.; Ning, C. Z. Nano Lett. 2009, 9, 784. (7) Zhu, H.; Fu, Y.; Meng, F.; Wu, X.; Gong, Z.; Ding, Q.; Gustafsson, M. V.; Trinh, M. T.; Jin, S.; Zhu, X. Y. Nat. Mater. 2015, 14, 636. (8) Fan, F.; Liu, Z.; Yin, L.; Nichols, P. L.; Ning, H.; Turkdogan, S.; Ning, C. Z. Semicond. Sci. Technol. 2013, 28, 065005. (9) Fan, F.; Turkdogan, S.; Liu, Z.; Shelhammer, D.; Ning, C. Z. Nat. Nanotechnol. 2015, 10, 796. (10) Li, J.; Meng, C.; Liu, Y.; Wu, X.; Lu, Y.; Ye, Y.; Dai, L.; Tong, L.; Liu, X.; Yang, Q. Adv. Mater. 2013, 25, 833. (11) Liu, X.; Zhang, Q.; Xiong, Q.; Sum, T. C. Nano Lett. 2013, 13, 1080. (12) Liu, X.; Zhang, Q.; Yip, J. N.; Xiong, Q.; Sum, T. C. Nano Lett. 2013, 13, 5336. (13) Camposeo, A.; Del Carro, P.; Persano, L.; Pisignano, D. Adv. Mater. 2012, 24, OP221. (14) Ta, V. D.; Chen, R.; Sun, H. D. Sci. Rep. 2013, 3, 1362. (15) Agarwal, R.; Barrelet, C. J.; Lieber, C. M. Nano Lett. 2005, 5, 917. (16) Kwon, J. E.; Park, S. Y. Adv. Mater. 2011, 23, 3615. (17) Grabowski, Z. R.; Rotkiewicz, K.; Rettig, W. Chem. Rev. 2003, 103, 3899. (18) Yu, J.; Cui, Y.; Xu, H.; Yang, Y.; Wang, Z.; Chen, B.; Qian, G. Nat. Commun. 2013, 4, 2719. (19) Zhang, C.; Yan, Y.; Zhao, Y. S.; Yao, J. Acc. Chem. Res. 2014, 47, 3448. (20) Gu, J.; Yan, Y.; Zhang, C.; Yao, J.; Zhao, Y. S. J. Mater. Chem. C 2014, 2, 3199. (21) Takazawa, K.; Inoue, J.-i.; Mitsuishi, K.; Takamasu, T. Phys. Rev. Lett. 2010, 105, 067401. (22) Zhang, W.; Yan, Y.; Gu, J.; Yao, J.; Zhao, Y. S. Angew. Chem. Int. Ed. 2015, 54, 7125. (23) Abdel-Mottaleb, M. S. A. Laser Chem. 1984, 4, 305. (24) Feng, J.; Tian, K.; Hu, D.; Wang, S.; Li, S.; Zeng, Y.; Li, Y.; Yang, G. Angew. Chem. Int. Ed. 2011, 50, 8072.

ACS Paragon Plus Environment

Page 5 of 5

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

Journal of the American Chemical Society

Tunable Lasing

TICT

LE

550 600 Wavelength (nm)

ACS Paragon Plus Environment

650