LETTERS PUBLISHED ONLINE: 25 JULY 2010 | DOI: 10.1038/NPHOTON.2010.174

Subnanosecond spectral diffusion measurement using photon correlation G. Sallen1, A. Tribu2, T. Aichele1,3, R. Andre´1, L. Besombes1, C. Bougerol1, M. Richard1, S. Tatarenko1, K. Kheng2 and J.-Ph. Poizat1 * Spectral diffusion is a result of random spectral jumps of a narrow line as a result of a fluctuating environment. It is an important issue in spectroscopy, because the observed spectral broadening prevents access to the intrinsic line properties. However, its characteristic parameters provide local information on the environment of a light emitter embedded in a solid matrix, or moving within a fluid, leading to numerous applications in physics and biology. We present a new experimental technique for measuring spectral diffusion based on photon correlations within a spectral line. Autocorrelation on half of the line and cross-correlation between the two halves give a quantitative value of the spectral diffusion time, with a resolution only limited by the correlation set-up. We have measured spectral diffusion of the photoluminescence of a single light emitter with a time resolution of 90 ps, exceeding by four orders of magnitude the best resolution reported to date. Spectral diffusion (SD) was first studied in spin resonance experiments1 Since then, it has been observed in various light-emitting systems including rare-earth ions2, ruby3, molecules4,5 or semiconductor quantum dots6–10. The SD of a single emitter results from fluctuations in its environment4–10. It is generally due to the Stark

effect, caused by randomly trapped charges in the vicinity of the emitter10. In this work we focus on light-emitting semiconducting nanostructures and, more specifically, quantum dots (QDs). Such nanostructures are very promising in the fields of quantum information or laser physics. Understanding their luminescence linewidth is obviously of primary importance in either of these applications where such an emitter is to be coupled to another emitter or to an optical cavity. Optical coherent control of a single qubit encoded on the spin of a QD can be implemented only in the absence of SD. The SD characteristic time gives the maximum time under which the system can be considered to be SD-free. More generally, reducing the SD requires a good understanding of its origin and therefore an accurate measurement of its temporal behaviour. The usual method with which to gather evidence of SD in a single emitter is to record a time series of spectra and to visualize directly the spectral wandering4,6–9. The time resolution of this method is limited by the ability to acquire a spectrum in a short time. The counting rate from a single emitter can barely exceed 1 × 105 s21 and it is therefore impossible to extract a spectrum on a timescale shorter than 1 × 1025 s. In practice, single-photon-counting charge-coupled 2

d b

BS2

H side

APD Spectro

e

Spectro

2.325

TCSPC

g(2) cx (L,L,τ)

BS1 2.330 Energy (eV)

1 τcx 0

1 τd

c Whole line τd

APD

τcx

t

L side H side t

0

f g(2) cx (L,H,τ)

Sample

L side Int (a.u.)

Laser

g(2) cx (τ)

a

1

0 −20

−10

0

10

20

Delay, τ (ns)

Figure 1 | Spectral diffusion measurements with photon correlations. a, Experimental set-up. BS1 and BS2, beamsplitters; spectro, spectrometer; APD, avalanche photodiode; TCSPC, time-correlated single-photon-counting data acquisition. b, Photoluminescence spectrum of the charged exciton transition of a single QD integrated over 1 s. The two halves of the line are labelled L and H. c, Representation of the photon time distribution in the two halves of the profile. d, Autocorrelation of the whole profile. e, Autocorrelation of one half of the profile showing the bunching due to SD (td ¼ 4 ns) and the narrower single photon antibunching. f, Cross-correlation between the two halves of the profile, displaying the antibunching due to SD with the same characteristic time td ¼ 4 ns as above. All data in d–f were obtained on the same QD with the same excitation power. The solid lines in e,f are fits with the model explained in the text and the Supplementary information.

1

CEA–CNRS–UJF group ‘Nanophysique et Semiconducteurs’, Institut Ne´el, CNRS – Universite´ Joseph Fourier, 38042 Grenoble, France, 2 CEA–CNRS–UJF group ‘Nanophysique et Semiconducteurs’, CEA/INAC/SP2M, 38054 Grenoble, France, 3 Physics Institute, Humboldt University, Berlin, Germany. * e-mail: [email protected] 696

NATURE PHOTONICS | VOL 4 | OCTOBER 2010 | www.nature.com/naturephotonics

© 2010 Macmillan Publishers Limited. All rights reserved.

NATURE PHOTONICS

LETTERS

DOI: 10.1038/NPHOTON.2010.174

a

b 1.2

1.0

g(2) (L,H,τ)

2.330

2.325

0.8

Energy (eV) 0.6

c

0.4

0.2 2.330

2.325 Energy (eV)

0.0 −10

0 Delay, τ (ns)

10

e

d 1.2

1.0 2.327 g(2) (L,H,τ)

0.8

2.330 Energy (eV)

0.6

f 0.4

0.2 2.327 0.0 −10

0 Delay, τ (ns)

2.330 Energy (eV)

10

Figure 2 | Influence of the width and position of the spectral windows in cross-correlations. a–c, The red trace (a) is the cross-correlation between spectral windows of width 0.3 meV (depicted in b) and the blue trace represents the cross-correlation between spectral windows of width 1.0 meV (c). d, Crosscorrelation between spectral windows with different energy separations. The red trace corresponds to the situation depicted in e and the blue trace that in f. In a and d, the red and blue traces are superimposed and therefore yield the same characteristic time td. Note that the pumping power for a is higher (15 mW) than in d (10 mW). These different experimental conditions are the reason for the different td values for a and d.

devices (CCD) cannot provide more than 1,000 images per second, so the time resolution cannot be better than a few milliseconds. Better SD time resolution has been obtained for inhomogeneously broadened ensembles of semiconducting nanocrystals by measuring a modulation frequency-dependent linewidth in a spectral hole-burning experiment11. With this technique the time resolution is set by the modulation frequency, and Palinginis and colleagues11 have obtained a resolution in the range of 100 ms. Another technique that also relies on resonant absorption has been developed by Zumbusch and colleagues5 for a single emitter. The absorption of a narrow laser line fluctuates because of the spectral fluctuations of the emitter line by means of which the resonant excitation is performed. The emitted fluorescence light therefore undergoes intensity fluctuations that are then measured by photon correlation. Zumbusch reports an ideal time resolution of 200 ns, but in that case the integration time set a practical limit of 1 ms. Achieving a short time resolution with individual emitters has been reported recently using photon-correlation Fourier spectroscopy (PCFS)12,13. This technique makes use of intensity correlations at the two outputs of a Michelson interferometer. The

fringe patterns at the two outputs are complementary for a given wavelength. Spectral jumps can cause a bright fringe to become dark on one output and appear bright on the other. This leads to changes in the output intensities on a characteristic time scale given by the SD time. The theoretical time resolution is determined by the photon correlation set-up as in the method we present here. However PCFS requires interferometric stability, and other researchers have only achieved a time resolution of 20 ms because of drift problems in their set-up14. In this Letter, we present a new and simple photon correlation technique to access characteristic SD times of a single emitter with a subnanosecond resolution. This is, to our knowledge, four orders of magnitude better than the best time resolution achieved to date5,14. Our technique is very robust because it is phase-insensitive and relies on linear optics. It is based on correlations of photons emitted within a spectral window narrower than the SD broadened inhomogeneous line (Fig. 1a–c). Owing to the wandering of the homogeneous line, the emission peak remains within this spectral window for a limited time, leading to photon bunching with a characteristic time td on autocorrelation on one half of the line

NATURE PHOTONICS | VOL 4 | OCTOBER 2010 | www.nature.com/naturephotonics

© 2010 Macmillan Publishers Limited. All rights reserved.

697

LETTERS

NATURE PHOTONICS γH

|1,L

|1,H

a

XX

5

γ

r

γH

|0,L L part

γL

2.214

3

2.216 E (eV)

2.218

2.234 2.236 E (eV)

2 1

H part

Figure 3 | Simplified level scheme used for the rate equation model of SD. The QD emits photons either on the low (L) or high (H) energy side of the emission peak. The QD emission energy diffuses with a rate gH from the L to the H part and gL from the H to the L part. On each energy side, the QD is modelled with two levels containing zero or one exciton, respectively. Note that other levels are included for the fits (see text). The quantities r and g are the incoherent pumping rate and the exciton decay rate, respectively.

(Fig. 1e), and to photon antibunching with the same characteristic time td for cross-correlation between the two halves of the profile (Fig. 1f ). The resolution is then only limited by the photon correlation set-up (see Methods). The minimum time delay between photons is of the order of tCX ¼ 600 ps and leads to the narrow antibunching dip observed in Fig. 1d–f. We have applied this technique to individual semiconducting quantum dots (QDs) embedded in a nanowire. Just like semiconducting nanocrystals6,10,11,14, these QDs suffer from ultrafast fluctuations caused by the vicinity of surface states, which is in contrast to the usual encapsulated self-assembled QDs. Details on the growth of the CdSe/ZnSe nanowires can be found in a previous work15. Exciton (X), biexciton (XX) and charged exciton (CX) transitions have been identified unambiguously using photon correlation spectroscopy16. The radiative lifetimes of these transitions are respectively tX ¼ 700 ps, tXX ¼ 400 ps and tCX ¼ 600 ps. The luminescence wavelength is 550 nm, with a high count rate of 25,000 counts per second at T ¼ 4 K. This system has demonstrated single-photon operation up to a temperature of 220 K (ref. 17). The microphotoluminescence experimental set-up is described in the Methods. In this work we first focused on the CX line shown in Fig. 1b. A series of spectra taken every 0.15 s over a period of 25 s does not exhibit any visible SD. The lineshape is better fitted with a Gaussian than with a Lorentzian. A Gaussian shape is characteristic of an inhomogeneous broadening mechanism like SD18. In Fig. 1d, the autocorrelation of the whole profile exhibits the characteristic antibunching dip of a single photon source. It is not very pronounced because the timing resolution of the experimental setup (800 ps) is of the same order as the lifetime of the emitter (tCX ¼ 600 ps). The slight bunching is due to hopping between the neutral and charged states of the QD. All these features have been discussed in detail for data obtained with a 90-ps-resolution set-up16. Fig. 1e presents the result of autocorrelation on one half of the emission peak. In addition to the clear antibunching dip (width ≈tCX) characteristic of single photon emission, it exhibits a clear bunching feature that is significantly larger than that obtained with the whole line profile in Fig. 1d. This demonstrates that the homogeneous line remains during a characteristic time td ¼ 4 ns within one-half of the SD-broadened line, as illustrated in Fig. 1c. Figure 1f shows cross-correlation measurements between the low- and high-energy sides of the emission spectrum. Broad antibunching is observed on the same timescale td as the bunching peak of Fig. 1e. This is different from the single photon antibunching and is a clear signature of SD, showing that it takes a characteristic time td for the intrinsic line to move from one half of the spectral profile to the other (Fig. 1c). 698

X

Cascade peak

4

|0,H

0 −2

−4

b

0 Delay, τ (ns)

5

2

4

XX X

4 g(2) x−xx (L,H,τ)

γ

g(2) x−xx (H,H,τ)

γL r

DOI: 10.1038/NPHOTON.2010.174

3

2.214

2.216

2.218

E (eV)

2.234 2.236 E (eV)

2 1 0 −4

−2

0

2

4

Delay, τ (ns)

Figure 4 | Correlated spectral diffusion between two lines. a, Crosscorrelation between the high energy sides (H) of the exciton (X) and the biexciton (XX) (see inset) exhibiting the characteristic biexciton–exciton cascade peak. b, Cross-correlation between the low energy sides (L) of the exciton (X) and the high energy side (H) of the biexciton (XX) lines (see inset). The absence of the characteristic cascade peak shows that two photons originating from different halves are not part of the same cascade. The solid line is a fit based on the model depicted in Fig. 3, where extra levels have been added as in ref. 16.

Autocorrelation and cross-correlation methods yield the same SD time td for the emitter, regardless of the width and position of the non-overlapping spectral windows, provided that the widths and energy difference of the latter are larger than the intrinsic homogeneous linewidth of the wandering line, which is generally the case. This makes this technique very robust. We have checked this non-trivial property experimentally (Fig. 2), and a theoretical analysis is given in the Supplementary Information. In autocorrelation experiments, the homogeneous line does not move for timescales shorter than td. After a time larger than td the memory of the spectral position is lost, so the number of autocorrelation events drops (Fig. 1e). As demonstrated in the Supplementary Information, the width of the bunching peak does not depend on the size of the spectral window; only the height of the peak does. In a similar manner, the cross-correlation function exhibits a lack of events for |t| , td (antibunching dip in Fig. 1f) corresponding to situations where the homogeneous line has not had time to hop to a position within the other spectral window. As shown in Fig. 2a, we have experimentally checked that the diffusion time td extracted from the data does not depend on the widths of the two spectral windows, as derived in the Supplementary Information for the case of spectral windows larger than the homogeneous linewidth. Figure 2d also shows that the energy separation between the two spectral windows in the cross-correlation set-up has no influence. This is due to the fact that, at each spectral jump, the new position is independent of the former and is randomly distributed over the whole inhomogeneous profile with a Gaussian probability. NATURE PHOTONICS | VOL 4 | OCTOBER 2010 | www.nature.com/naturephotonics

© 2010 Macmillan Publishers Limited. All rights reserved.

NATURE PHOTONICS

LETTERS

DOI: 10.1038/NPHOTON.2010.174

The diffusion time td is extracted by fitting experimental data, and its value is more accurate if td is of the same order or larger than the lifetime of the emitter. The principle of the model is depicted in Fig. 3 for the simplified case of two complementary spectral windows. Each spectral window is modelled as a two-level system. Each level is connected to its counterpart with a ‘jump’ rate gi (i ¼ L, H). As shown in the Supplementary Information, the SD rate is given by gd ¼ Sigi and is extracted directly from the time width of the correlation measurement. In practice, the fitting model allows for non-complementary spectral windows, and uses a multilevel system to account for the neutral (X) and charged (CX) excitons, dark exciton and neutral biexciton (see Supplementary Information and ref. 16). An extra feature of this photon correlation technique is that it allows the investigation of correlated SD between different lines coming either from the same emitter or from different emitters. This possibility gives information on the energy shift of different lines caused by a common change in their environment. We have performed cross-correlation between one half of the exciton (X) line and one half of the biexciton (XX) line of the same quantum dot. This experiment was performed with fast avalanche photodiodes (APDs), leading to a correlation timing resolution of 90 ps (see Methods). The results are shown in Fig. 4 for the H half of the X line correlated either with the H half of the XX line (Fig. 4a) or with its L half (Fig. 4b). The characteristic cross-correlation peak of the biexciton–exciton cascade19 can be seen clearly in Fig. 4a for the same halves of the profile, but it is missing in Fig. 4b. This shows clearly that the sign of the energy shifts of the X and XX lines due to the fluctuating environment is the same, as already observed by Besombes and colleagues9 at slower timescales. To summarize, we have presented a simple and robust method to measure SD of single emitters with a resolution of 90 ps. As shown experimentally, this technique enables fluctuations of the nanoenvironment of a single emitter to be probed with a time resolution that is better by four orders of magnitude than that of existing achievements.

Methods Experimental set-up. The set-up was based on a standard microphotoluminescence experiment operating at a temperature of T ¼ 4 K. The sample was excited by a continuous-wave diode laser emitting at 405 nm via a beamsplitter BS1 (Fig. 1) with a transmission of 70%. The luminescence was split by a 50/50 beamsplitter (BS2) and each beam sent to a monochromator (resolution dE ¼ 0.2 meV or dl ¼ 0.05 nm), the output slit of which was imaged on an APD. The width of the output slit could be varied, allowing us to choose the spectral window within the SD inhomogeneously broadened line. The voltage pulses of each APD were sent to a time-correlated single-photon counter (TCSPC) to build a histogram of the time delays between photons. This allowed us to perform either autocorrelation when the two monochromators were tuned to the same wavelength, or cross-correlation otherwise. Other than the results of Fig. 4, the work presented here was obtained with high-quantum-efficiency APDs (h ¼ 60% at 550 nm). A high detection efficiency is important when performing correlation experiments because the integration time is proportional to h 2. The price to pay is the slower timing resolution. With these APDs, the measured timing resolution of the whole set-up is 800 ps (full-width at half-maximum). This rather slow time resolution is not a limitation in our case because the SD times that we are investigating are in the range of 10 ns. The results shown in Fig. 4 have been obtained with fast APDs to allow observation of the temporally narrow structures. In that case the measured time resolution of the set-up was 90 ps, and the APD quantum efficiency was h ¼ 30%.

Received 28 January 2010; accepted 21 June 2010; published online 25 July 2010

References 1. Klauder, J. R. & Anderson, P. W. Spectral diffusion decay in spin resonance experiments. Phys. Rev. 125, 912–932 (1962). 2. Flach, R., Hamilton, D. S., Selzer, P. M. & Yen, W. M. Time-resolved fluorescence line-narrowing studies in LaF3: Prþ 3 . Phys. Rev. Lett. 35, 1034–1037 (1975). 3. Szabo, A. & Kaarli, R. Optical hole burning and spectral diffusion in ruby. Phys. Rev. B 44, 12307–12313 (1991). 4. Ambrose, W. P. & Moerner, W. E. Fluorescence spectroscopy and spectral diffusion of single impurity molecules in a crystal. Nature 349, 225–227 (1991). 5. Zumbusch, A., Fleury, L., Brown, R., Bernard, J. & Orrit, M. Probing individual two-level systems in a polymer by correlations of single molecule fluorescence. Phys. Rev. Lett. 70, 3584–3587 (1993). 6. Empedocles, S. A., Norris, D. J. & Bawendi, M. G. Photoluminescence spectroscopy of single CdSe nanocrystallite quantum dots. Phys. Rev. Lett. 77, 3873–3876 (1996). 7. Robinson, H. D. & Goldberg, B. B. Light-induced spectral diffusion in single self-assembled quantum dots. Phys. Rev. B 61, R5086–R5089 (2000). 8. Tu¨rck, V. et al. Effect of ramdom field fluctuations on excitonic transitions of individual CdSe quantum dots. Phys. Rev. B 61, 9944–9947 (2000). 9. Besombes, L., Kheng, K., Marsal, L. & Mariette, H. Few-particle effects in single CdTe quantum dots. Phys. Rev. B 65, 121314 (2002). 10. Empedocles, S. A. & Bawendi, M. G. Quantum-confined Stark effect in single CdSe nanocrystallite quantum dots. Science 278, 2114–2117 (1997). 11. Palinginis, P., Tavenner, S., Lonergan, M. & Wang, H. Spectral hole burning and zero phonon linewidth in semiconductor nanocrystals. Phys. Rev. B 67, 201307 (2003). 12. Brokmann, X., Bawendi, M. G., Coolen, L. & Hermier, J. P. Photon-correlation Fourier spectroscopy. Opt. Express 14, 6333–6341 (2006). 13. Coolen, L., Brokmann, X. & Hermier, J. P. Modeling coherence measurements on a spectrally diffusing single-photon emitter. Phys. Rev. A 76, 033824 (2007). 14. Coolen, L., Brokmann, X., Spinicelli, P. & Hermier, J. P. Emission characterization of a single CdSe–ZnS nanocrystal with high temporal and spectral resolution by photon-correlation Fourier spectroscopy. Phys. Rev. Lett. 100, 027403 (2008). 15. Aichele, T. et al. Defect-free ZnSe nanowire and nanoneedle nanostructures. Appl. Phys. Lett. 93, 143106 (2008). 16. Sallen, G. et al. Exciton dynamics of a single quantum dot embedded in a nanowire. Phys. Rev. B 80, 085310 (2009). 17. Tribu, A. et al. A high-temperature single-photon source from nanowire quantum dots. Nano Lett. 8, 4326–4329 (2008). 18. Berthelot, A. et al. Unconventional motional narrowing in the optical spectrum of a semiconductor quantum dot. Nature Phys. 2, 759–764 (2006). 19. Moreau, E. et al. Quantum cascade of photons in semiconductor quantum dots. Phys. Rev. Lett. 87, 183601 (2001).

Acknowledgements The authors acknowledge the very efficient technical support of F. Donatini and careful reading of the manuscript by Le Si Dang and G. Nogues. T.A. acknowledges support from the Deutscher Akademischer Austauschdienst (DAAD). Part of this work was supported by the European project QAP (contract no. 15848).

Author contributions G.S. conducted the optical experiments and analysed the data. A.T., T.A., R.A., S.T. and K.K. carried out fabrication and processing of the samples, and C.B. performed their structural analysis. G.S., L.B., M.R. and J.P.P. contributed to the genesis of the idea and to the discussion of the results. J.P.P. supervised the optical experiments and wrote the paper.

Additional information The authors declare no competing financial interests. Supplementary information accompanies this paper at www.nature.com/naturephotonics. Reprints and permission information is available online at http://npg.nature.com/reprintsandpermissions/. Correspondence and requests for materials should be addressed to J.P.P.

NATURE PHOTONICS | VOL 4 | OCTOBER 2010 | www.nature.com/naturephotonics

© 2010 Macmillan Publishers Limited. All rights reserved.

699

Subnanosecond spectral diffusion measurement using ...

Jul 25, 2010 - The authors acknowledge the very efficient technical support of F. Donatini and careful reading of the manuscript by Le Si Dang and G. Nogues ...

409KB Sizes 4 Downloads 293 Views

Recommend Documents

Subnanosecond spectral diffusion measurement using ...
Jul 25, 2010 - resolution only limited by the correlation set-up. We have measured spectral ... time under which the system can be considered to be SD-free. More generally ... devices (CCD) cannot provide more than 1,000 images per second, so the tim

Diffusion Maps, Spectral Clustering and Eigenfunctions ...
spectral clustering and dimensionality reduction algorithms that use the ... data by the first few eigenvectors, denoted as the diffusion map, is optimal under a ...

Quantitative impedance measurement using atomic ...
Sep 15, 2004 - Building 530, Room 226, Stanford, California 94305-3030 ... example, obtaining quantitative kinetic data from the recently developed atomic force microscopy ... faces. The technique has been applied to visualize electronic.

Periodic Measurement of Advertising Effectiveness Using Multiple ...
pooled to create a single aggregate measurement .... plete their research, make a decision, and then visit a store .... data from each test period with the data from.

Content-Based Video Summarization using Spectral ...
application is developed to qualitatively test the results of video summarization for ... have a resource-constrained development environment. Due to the limited ...

subnanosecond-pulse generator with variable ...
Dec 9, 1974 - problem, if we assume v = 0, i.e. Q2°° and Q3 = 0; ^2°° .... 200 ps. The vertical unit is 5 V. 22. (6) with RE2 from eqn. 5 and (Fig. 2A). 1. 1. R,. 1.

Objective Measurement of Wool Fibre Fineness Using 2 ...
throughout the wool clips. In this study, we employ the concept of texture analysis based on ... and will have more of them per unit fibre length; when the sheep is ...

Noise measurement from magnitude MRI using ...
Aug 3, 2010 - (Some figures in this article are in colour only in the electronic version). 1. ... variance gives a measure of the quality of the MR data. ..... Cocosco C A et al 1997 Brainweb: online interface to a 3D MRI simulated brain database ...

Measurement of brain glutamate using TE‐averaged PRESS at 3T
head coil, PROBE/SVQ™ (automated PRESS), and a mod- ... E-mail: [email protected] ... transform (FFT), and automatic zero-order phase correc- tion.

Spectral unmixing versus spectral angle mapper for ...
to assess the classification performance for identifying and mapping 'desert like' .... The derived spectral angle maps form a new data cube with the number of bands equal .... Interactive visualization and analysis of imaging spectrometer data.

Spectral karyotyping
spectrum at all image points. Here we describe the principle of spectral imaging, define ... lens (the system can also be attached to any other optics such as a telescope or a .... taken for an infinite interferogram, and the zero filling is an optio

measurement of digitally modulated signals by using ...
long distance telephone call was made on 10 August 1876 by Bell from the ... System. 2. System Model. The basic structure of designed communication system ...

Angle measurement using Bevel protractor and Sine ... -
The protractor dial is slotted to hold a blade, which can be rotated with the dial to the required angle and also independently adjusted to any desired length.

Spectral unmixing versus spectral angle mapper for ...
process of desertification and land degradation as a whole. Hyperspectral remote ..... Mapping target signatures via partial unmixing of ... integration of image processing, digital elevation data and field knowledge. (application to Nepal). JAG.

Spectral Clustering - Semantic Scholar
Jan 23, 2009 - 5. 3 Strengths and weaknesses. 6. 3.1 Spherical, well separated clusters . ..... Step into the extracted folder “xvdm spectral” by typing.

Measurement - GitHub
Measurement. ○ We are now entering the wide field era. ○ Single gains no longer adequate. ○ Gains are direction dependant ...

Subnanosecond time-resolved x-ray measurements ...
(Received 31 October 2008; accepted 9 December 2008; published online 29 December 2008). We have .... thickness: 0.9 mm) was mounted on the window of a photo- multiplier ... storage ring by using a fast amplifier, a constant-fraction dis-.

NONLINEAR SPECTRAL TRANSFORMATIONS ... - Semantic Scholar
noisy speech, these two operations lead to some degrada- tion in recognition performance for clean speech. In this paper, we try to alleviate this problem, first by introducing the energy information back into the PAC based features, and second by st

Parallel Spectral Clustering
Key words: Parallel spectral clustering, distributed computing. 1 Introduction. Clustering is one of the most important subroutine in tasks of machine learning.

SPECTRAL ISOMORPHISMS BETWEEN ...
Then we define the norming constants νn,n ≥ 0, by νn = yn. 2 a (resp. νn = .... We now define R(q,β,ν0,λ0)=(q1,α1,a1,b1) for (q,β,ν0,λ0) ∈ ˆE (resp. R(q, a′,b′ ...