SIL-based confocal fluorescence microscope for investigating individual nanostructures

Similar documents
Multiplexed 3D FRET imaging in deep tissue of live embryos Ming Zhao, Xiaoyang Wan, Yu Li, Weibin Zhou and Leilei Peng

Supporting Information for. Electrical control of Förster energy transfer.

Crystallographic Characterization of GaN Nanowires by Raman Spectral Image Mapping

Confocal Microscopy & Imaging Technology. Yan Wu

Supplementary Table 1. Components of an FCS setup (1PE and 2PE)

Supporting Information. Two-Photon Luminescence of Single Colloidal Gold NanoRods: Revealing the Origin of Plasmon Relaxation in Small Nanocrystals

Cellular imaging using Nano- Materials. A Case-Study based approach Arun Murali, Srivats V

Lab 5: Optical trapping and single molecule fluorescence

Sub-micron scale patterning of fluorescent. silver nanoclusters using low-power laser

A Thin Layer Imaging with the Total Internal Reflection Fluorescence Microscopy

Sample region with fluorescent labeled molecules

Introduction. (b) (a)

Active delivery of single DNA molecules into a plasmonic nanopore for. label-free optical sensing

Measurement of surface concentration of fluorophores using. fluorescence fluctuation spectroscopy

Super Resolution Microscopy - Breaking the Diffraction Limit Radiological Research Accelerator Facility

micromachines ISSN X

Directional Surface Plasmon Coupled Emission

Title: Localized surface plasmon resonance of metal nanodot and nanowire arrays studied by far-field and near-field optical microscopy

Special Techniques 1. Mark Scott FILM Facility

Visualisation, Sizing and Counting of Fluorescent and Fluorescently-Labelled Nanoparticles

Building An Ultrafast Photon-Induced Near-field Transmission Electron Microscope

Quantum Dot applications in Fluorescence Imaging for Calibration and Molecular Imaging

A Brief History of Light Microscopy And How It Transformed Biomedical Research

ECE280: Nano-Plasmonics and Its Applications. Week5. Extraordinary Optical Transmission (EOT)

F* techniques: FRAP, FLIP, FRET, FLIM,

PALM/STORM, BALM, STED

Fluorescence Light Microscopy for Cell Biology

Supporting Information. Label-Free Optical Detection of DNA. Translocations Through Plasmonic Nanopores

Department of Chemistry, Center for Photochemical Sciences, Bowling Green State University,

Plasmonics using Metal Nanoparticles. Tammy K. Lee and Parama Pal ECE 580 Nano-Electro-Opto-Bio

Preparation and characterization of Co BaTiO 3 nano-composite films by the pulsed laser deposition

Articles. High-Resolution Fluorescence Near-Field Imaging of Individual Nanoparticles via the Tip-Induced Quenching Technique

Strong electromagnetic confinement near dielectric microspheres to enhance single-molecule fluorescence

Fundamentals of X-ray diffraction and scattering

average diameter = 3 nm, from PlasmaChem) was mixed in NLCs to produce QDembedded

Imaging facilities at WUR

Factors Affecting QE and Dark Current in Alkali Cathodes. John Smedley Brookhaven National Laboratory

Real-Time Multiplex Kinase Phosphorylation Sensors in Living Cells

OCTOPLUS QPLEX FLUORESCENCE IMAGER. for fast & powerful fast 2D Gel image acquisition

Femtosecond micromachining in polymers

Monitoring dynamic photocatalytic activity of single CdS nanoparticles

Supporting information

In situ semi-quantitative assessment of single cell viability by resonance

Confocal Microscopy of Electronic Devices. James Saczuk. Consumer Optical Electronics EE594 02/22/2000

Bragg diffraction using a 100ps 17.5 kev x-ray backlighter and the Bragg Diffraction Imager

Laser Processing and Characterisation of 3D Diamond Detectors

SLS-process monitoring and temperature control

Supplementary Note 1: Estimation of the number of the spectroscopic units inside the single Pdots

Structure and optical properties of M/ZnO (M=Au, Cu, Pt) nanocomposites

Total Internal Reflection Fluorescence Microscopy

SUPPLEMENTARY INFORMATION

Supplementary Figures

Spectra Chacracterizations of Optical Nanoparticles

Supplementary Information

Photoluminescence Spectroscopy on Chemically Synthesized Nanoparticles

IMMERSION HOLOGRAPHIC RECORDING OF SUBWAVELENGTH GRATINGS IN AMORPHOUS CHALCOGENIDE THIN FILMS

Applicability of Hyperspectral Fluorescence Imaging to Mineral Sorting

Macromolecular environments influence proteins

Using a Scientifica Multiphoton SliceScope for Fluorescence Lifetime Imaging

Microscopy. CS/CME/BioE/Biophys/BMI 279 Nov. 2, 2017 Ron Dror

Micro- and Nano-Technology... for Optics

Fluorescence Microscopy. Terms and concepts to know: 10/11/2011. Visible spectrum (of light) and energy

Confocal Microscopes. Evolution of Imaging

Supplementary Figure 1 Scanning electron micrograph (SEM) of a groove-structured silicon substrate. The micropillars are ca. 10 μm wide, 20 μm high

Gold nanorods as multifunctional probes. in liquid crystalline DNA matrix

Carbon Black At-line Characterization. Using a Portable Raman Spectrometer

Compensation: Fundamental Principles

Performance of the Micro Photon Devices PDM 50CT SPAD detector with PicoQuant TCSPC systems

Cell analysis and bioimaging technology illustrated

Synthesis and Laser Processing of ZnO Nanocrystalline Thin Films. GPEC, UMR 6631 CNRS, Marseille, France.

Design for Manufacturability (DFM) in the Life Sciences

Supporting Information

Supplementary Figures

Optical Observation - Hyperspectral Characterization of Nano-scale Materials In-situ

ENERGY-DISPERSIVE X-RAY FLUORESCENCE ANALYSIS OF MONO- AND POLYCRYSTALS OF SELENIDE SPINELS BY FUNDAMENTAL PARAMETER METHOD

Digital resolution enhancement in surface plasmon microscopy

Widefield Microscopy Bleed-Through

Surface plasmon enhanced emission from dye doped polymer layers

Modification of Glass by FS Laser for Optical/Memory Applications

Fs- Using Ultrafast Lasers to Add New Functionality to Glass

New developments in STED Microscopy

Nano Wear Testing Polymer Coated Glass

Electron Probe Micro-Analysis (EPMA)

Flow Cytometry - The Essentials

In-situ monitoring of optical near-field material processing by electron microscopes

Seminar: Structural characterization of photonic crystals based on synthetic and natural opals. Olga Kavtreva. July 19, 2005

Monitoring and Optimizing the Lipopolysaccharides-plasmid DNA interaction by FLIM-FRET

FLIM Fluorescence Lifetime IMaging

Supporting Information: Gold nanorod plasmonic upconversion microlaser

Simultaneous multi-color, multiphoton fluorophore excitation using dual-color fiber lasers

Bioinstrumentation Light Sources Lasers or LEDs?

Raman Spectroscopy Measurement System and Data Analysis for Characterization of Composite Overwrapped Pressure Vessels (COPVs)

Localization Microscopy

Super-resolution Microscopy

Supporting Information

LASERS, LEDS AND OTHER LIGHTING SOURCES FOR LIFE SCIENCE APPLICATIONS. Wallace Latimer Coherent

Two-Photon Microscopy for Deep Tissue Imaging of Living Specimens

Supplemental Information. Ca 2+ and Myosin Cycle States Work as Allosteric Effectors of Troponin. Activation

Short Length High Gain ASE Fiber Laser at 1.54µm by High Co-doped Erbium and Ytterbium Phosphate Laser Glasses

Direct visualization, sizing and concentration measurement of fluorescently labeled nanoparticles using NTA

Transcription:

Cent. Eur. J. Phys. 9(2) 2011 293-299 DOI: 10.2478/s11534-010-0098-5 Central European Journal of Physics SIL-based confocal fluorescence microscope for investigating individual nanostructures Research Article Bartosz Krajnik 1, Tim Schulte 2, Dawid Piątkowski 1, Nikodem Czechowski 1, Eckhard Hofmann 2, Sebastian Mackowski 1 1 Institute of Physics, Nicolaus Copernicus University, Grudziadzka 5, Torun, Poland 2 Department of Biology and Biotechnology, Ruhr-University Bochum, Bochum, Germany Received 22 July 2010; accepted 29 September 2010 Abstract: We developed a fluorescence confocal microscope equipped with a hemispherical solid immersion lens (SIL) and apply it to study the optical properties of light-harvesting complexes. We demonstrate that the collection efficiency of the SIL-equipped microscope is significantly improved, as is the spatial resolution, which reaches 600 nm. This experimental setup is suitable for detailed studies of physical phenomena in hybrid nanostructures. In particular, we compare the results of fluorescence intensity measurements for a light-harvesting peridinin-chlorophyll-protein (PCP) complex with and without the SIL. PACS (2008): 33.50.-j, 87.14.E-, 87.64.kv, 87.64.M- Keywords: confocal microscopy light-harvesting complex solid immersion lens fluorescence imaging Versita Sp. z o.o. 1. Introduction The spatial resolution of all optical systems is limited by diffraction, which results from the wave nature of light. Resolution of the conventional optical microscope is approximately λ 0, where λ 2NA 0 is the free space wavelength and NA is the numerical aperture of the objective defined by the angular semiaperature in object space θ a, and the refractive index n, through the relation NA = n sin θ a. For a microscope objective with NA = 0.45 and wavelength λ 0 = 670 nm, the resolution in air with n = 1 is approxi- E-mail: bakrag@phys.uni.torun.pl E-mail: mackowski@fizyka.umk.pl (Corresponding author) mately 750 nm. Another important drawback of using standard collection optics, in particular microscope objectives with large working distances, is very low collection efficiency, which prevents high-accuracy studies of individual nanostructures. There are several ways to overcome this limitation. One possibility is to fill the space between an objective and a sample with high-refraction-index liquid, as commonly applied in oil-immersion microscopy. The oil used for that purpose is characterized with a refractive index of 1.53 in order to match the refractive index of microscope coverslips. However despite its many advantages, oil immersion microscopy has certain limitations: first, it cannot be used for measurements of samples sensitive to the contact with immersion oil; second, oil-immersion microscopy is not suitable for low-temperature experiments that could provide valuable information about biological 293

SIL-based confocal fluorescence microscope for investigating individual nanostructures systems [1]. In recent years SIL-based microscopy has been used to investigate the optical properties of semiconductor quantum wells [2] and quantum dots [3 5]. It has been shown that the numerical aperture of 2 can be achieved when the refraction index of the SIL is very high (GaP 3.5) and closely matches that of the studied sample. One of the critical issues related to this technique concerns the presence of the air gap between the surface of the SIL and the epitaxially grown sample [3]. In this work we apply a solid immersion lens-based confocal fluorescence microscope to study pigment-protein complexes embedded in a polymer matrix. We investigate the optical properties of a peridinin-chlorophyll-protein (PCP) light-harvesting complex. This is a water-soluble light-harvesting antenna from the dinoflagellate Amphidinium carterae [6]. Our experiments obtained for PCP complexes show that application of the SIL significantly improves the optical collection efficiency. We achieve an optical resolution for the setup of about 600 nm, which is almost half of the bare objective. Enhanced collection efficiency results in much higher contrast in measured fluorescence maps. The results provide solid evidence that such an experimental setup can be applied for studying individual nanostructures. excellent matching of the peridinin absorption and typical energies of plasmon excitations in metallic nanoparticles, render PCP a model system for studying the plasmonic interactions in light-harvesting biomolecules [10, 11]. 2. Materials and methods In our experiment, we use PCP reconstituted with Chl a synthesized according to the protocol by Miller et al. [7]. The structure of the reconstituted PCP has been determined with better than < 1.5 Å resolution using X-ray crystallography [8] and is shown in Fig. 1a. Each protein monomer of PCP surrounds a pigment cluster of four peridinin molecules, one Chl a and one lipid molecule. The Chl a molecules in the PCP complex are separated by less than 2 nm and peridinin molecules are in van der Waals contact with the Chl a molecules. The absorption spectrum of the PCP complex (Fig. 1b) shows that peridinin is the primary absorbing pigment responsible for an absorption band ranging from 350 to 550 nm. On the other hand, the Chl a molecules absorb around 668 nm (Q y band) and 440 nm (Soret band). This light-harvesting complex is characterized with effcient energy transfer from the peridinins to the Chl a molecules with an efficiency higher than 90% [9]. The fluorescence of PCP originates from the Q y transition of Chl a and is located at 673 nm (Fig. 1c) [9]. Relatively high quantum yield and stability of the PCP complexes make them suitable for singlemolecule studies [9]. Large energy separation between peridinin absorption and Chl fluorescence together with Figure 1. (a) Structure of a homodimer of refolded PCP (RFPCP): four peridins (orange), one chlorophyll a (green) and one lipid molecule (blue). PDB entry 3IIS. (b) Absorption and (c) fluorescence spectrum of PCP complexes measured at room temperature (after [9]). The excitation wavelength for the fluorescence was 485 nm. 294

Bartosz Krajnik, Tim Schulte, Dawid Piątkowski, Nikodem Czechowski, Eckhard Hofmann, Sebastian Mackowski In Fig. 2 we show the scheme of our home-built confocal fluorescence microscope based on the Olympus infinitycorrected microscope objective LMPlan 50x, characterized with a numerical aperture of 0.5 and a working distance of 6 mm. The resulting laser spot size is about 1 µm for the excitation laser of 485 nm. In order to improve the collection efficiency and spatial resolution we use a 3 mm diameter hemispherical solid immersion lens made of S-LAH-71 glass with a refraction index n = 1.85. The SIL is pressed against the sample by spring mount in order to reduce the air gap between the sample and the lens. The sample is placed on a XYZ piezoelectric stage (Physik Instrumente) with 1 nm nominal resolution of a single step, which enables us to raster-scan the sample surface in order to collect fluorescence maps. The maps are formed by combining fluorescence intensity measurements with the motion of the XY translation stage. For fluorescence excitation, we use one of four diode-pumped solid-state lasers with wavelengths of 405, 485, 532 and 640 nm. Typical optical power of the laser sources is about 7 mw, but in the case of actual measurements it needs to be strongly reduced in order to prevent photobleaching of the molecules. We used excitation powers of about 40 and 4 µw. Gaussian beams of the lasers are achieved by using a spatial filter. The fluorescence is detected in a back-scattering geometry and focused on a confocal pinhole (150 µm) in order to reduce stray light coming out of the focal plane. The emission of PCP complexes is extracted with HQ 650LP (Chroma) dichroic mirror and HQ 670/10 (Chroma) bandpass filter. Figure 2. Experimental setup of the SIL-based confocal fluorescence microscopy. Our experimental configuration allows for measuring fluorescence intensity, spectra and lifetimes. The spectrum, dispersed using the Amici prism is measured with a CCD camera (Andor idus DV 420A-BV). The spectral resolution of the system is about 2 nm. Fluorescence intensity maps are collected with an avalanche photodiode (PerkinElmer SPCM-AQRH-14) with dark count rate of about 80 cps. Fluorescence lifetimes are measured using a time-correlated single photon counting module (Becker & Hickl) equipped with fast avalanche photodiode (idquantique id100-50) triggered by a laser pulse. The time resolution of the TCSPC setup is about 30 ps. First, in order to observe fluorescence maps for various concentrations of a fluorophore, a series of samples was prepared. PCP with concentration of 0.49 mg/ml was dissolved in a 2% water solution of PVA in a 1 to 10 ratio (stock solution). Next, we prepared 100-fold, 500-fold, and 2500-fold diluted PCP solutions. To make the samples, 10 µl of each solution was dropped on a coverslip and spin-coated at 2500 rpm for 60 seconds. In this way, a thin, homogeneous layer was obtained. Embedding the PCP complexes in a polyvinyl alcohol (PVA) matrix has been shown to improve the photostability of the PCP complexes [12] and should also minimize the impact of the air 295

SIL-based confocal fluorescence microscope for investigating individual nanostructures gap between the SIL surface and the sample. Next, we prepared a solution that was diluted by 12500 times compared to the stock solution. The purpose was to obtain a sample with a concentration comparable to the single molecule limit for this system [13]. These highly diluted samples were also spin-coated in a PVA matrix on glass coverslips. As a reference, samples containing only the PVA water solution were prepared. 3. Results and discussion In Fig. 3 we show fluorescence intensity maps obtained with the bare microscope objective for the 100-fold (3a), 500-fold (3b), and 2500-fold (3c) diluted samples. The size of each map is 20 20 µm and the step size is 250 nm. The PCP complexes were excited with a wavelength of 485 nm and the laser power at the objective was about 45 µw, while the integration time was 0.01 s. The fluorescence maps are quite homogeneous, which suggests a uniform distribution of the PCP complexes in a polymer matrix for such concentrated solutions. We analyzed the maps quantitatively by counting the occurrence of each fluorescence intensity value within the map (histograms shown in Fig. 3). The average fluorescence intensity decreases with decreasing PCP concentration from about 8040 through 2060 to 460 cps for the most diluted sample. The normalized widths of the histograms peaks are: 0.11, 0.27, and 0.47, respectively. The increase in the peak width is perhaps the result of reduced signal-to-noise ratio due to a decrease of the fluorophore concentration. As can be seen from Fig. 3, the histograms are not symmetrical and low intensity tails are present. The origin of these tails is not clear at this point: one possibility could be the photobleaching of complexes during the experiment; on the other hand, such tails could also due to small variations in PCP concentration across the sample. Nevertheless, the results shown in Fig. 3 indicate that the samples prepared in the PVA matrix indeed contain PCP complexes. This conclusion is important for carrying out experiments for highly diluted samples. Figure 3. Upper row: Fluorescence intensity maps measured with the bare micorscope objective for PCP samples with concentrations diluted (a) 100-fold (b) 500-fold, and (c) 2500-fold from the stock solution. Size of the maps is 20 20 µm with a pixel size of 250 250 nm. Bottom row: Histograms displaying fluorescence intensities extracted from the respective maps. In order to test the performance of the bare objective we have measured fluorescence maps of a 12500-fold diluted (4 10 6 mg/ml) PCP sample. Such a low concentration is comparable to the single molecule limit described pre- 296

Bartosz Krajnik, Tim Schulte, Dawid Piątkowski, Nikodem Czechowski, Eckhard Hofmann, Sebastian Mackowski viously for PCP in a PVA matrix [9]. The fluorescence map of the sample is shown in Fig. 4a, and we compare it to the map obtained for our reference sample, the pure PVA solution (Fig. 4b). The latter was prepared in exactly the same way to facilitate straightforward comparison. In order to reduce the effect of photobleaching, particularly important for low concentration samples, the laser power was reduced to 4 µw. At the same time the integration time was increased to 0.1 s. As in the case of the concentrated samples (Fig. 3) we use histograms of fluorescence intensity to quantify the observations. Figure 4. Left column: (a) Fluorescence intensity maps measured with bare objective for highly diluted PCP sample (12500-fold diluted from the stock solution) compared to that of the reference PVA-only sample (b). Size of the maps is 20 20 µm with pixel size of 250 250 nm. Right column: Corresponding histograms of signal intensity obtained for the same two samples as in (a). The reference sample (2% PVA water solution) intensity histogram shown in Fig. 4b, is well defined by a Gaussian distribution with an average intensity of 1800 cps and a width of about 340 cps. In contrast, the PCP fluorescence intensity map (Fig. 4a) shows many peaks with high fluorescence intensity, superimposed on a Gaussian background distribution. The peaks correspond to PCP aggregates present in the sample, recognized as dark spots in the image, a few pixels large (marked by the circles). Low optical resolution makes estimation of the sizes of these structures extremely difficult. The final step to demonstrate the advantages of using SIL was to repeat the experiment for the highly diluted PCP sample in a configuration where the microscope objective is coupled with the SIL. A fluorescence map obtained for the highly diluted PCP sample (12500-fold from the stock solution) is shown in Fig. 5a together with an image collected for a reference sample comprised of PVA 297

SIL-based confocal fluorescence microscope for investigating individual nanostructures only (Fig. 5b). The excitation power of the laser remained unchanged as compared to the data shown in Fig. 4. Since introduction of the SIL changes the overall numerical aperture of the collection optics, the actual area measured during the scan is smaller than the movement of the XY translation stage. From the calibration measurements (not shown) of back-scattered light from the semiconductor sample with 10 µm periodic squares, we have determined that the SIL reduces the image by a factor of 1.7 in a linear dimension. This value closely matches the refractive index of the SIL material (1.85), indicating very good optical design of the setup and in particular, negligible impact of the air gap between the sample and the PCP complexes. Consequently, the scan size of 20 20 µm with step size of 250 nm translates into a map of 12 12 µm. We observe that distribution of the fluorescence intensity obtained for the reference sample (Fig. 5b) has similar shape and average intensity (about 1800 cps) as the results obtained without the SIL (Fig. 4b). Although somewhat surprising, this might indicate that we measure only stray light not associated with any unspecific emission from the reference samples. Otherwise, the intensity should increase by approximately 3-fold due to increased excitation power density with the application of the SIL. In contrast, the fluorescence map measured for samples containing PCP complexes contains many extremely bright spots with average intensities reaching 5000 cps. The intensities of two such spots A and B marked in Fig. 5a are 1700 and 4220 cps, respectively. Moreover, the contrast between the areas associated with emitting PCP complexes and regions where there is no fluorescing molecules is remarkably high. Direct comparison between the fluorescence maps obtained for highly diluted PCP samples measured with the SIL and with the objective only clearly point to substantial improvement of the data quality in the case of the former configuration. Figure 5. Left column: (a) Fluorescence intensity maps measured with the objective coupled to the SIL for a highly diluted PCP sample (12500- fold diluted from the stock solution) compared to that of the reference PVA-only sample (b). Size of the maps is 12 12 µm with pixel size of 150 150 nm. Right column: Corresponding histograms of signal intensity obtained for the same two samples as in (a). 298

Bartosz Krajnik, Tim Schulte, Dawid Piątkowski, Nikodem Czechowski, Eckhard Hofmann, Sebastian Mackowski In order to quantify the upper limit of the spatial resolution in the SIL-based confocal microscope we show two exemplary cross sections (Fig. 6) obtained for spots A and B marked in Fig. 5a. The time trajectory of the fluorescence intensity measured for this spot features exponential behavior, which suggests that the number of fluorescing molecules is greater than ten [13]. For points A and B, we obtain widths of 640±10 and 770±20 nm, respectively. While the larger spot (B) is most probably due to an aggregate of PCP complexes, the emission attributed to the spot A, which is far less intense, could potentially be due to a smaller number of molecules. We note that these numbers do not reflect the actual size of the emitter but are associated with the point spread function of the emitter which is in turn determined by the spatial resolution of the collection optics. Therefore, the sizes obtained in Fig. 6 provide an upper limit for our optical resolution. Future work will focus on studying the PCP complexes without the SIL-based confocal microscopy setup at the single molecule level. This should provide insight into the photophysics of these complexes as well as allow full characterization of the optical parameters of the setup. 4. Conclusions We demonstrate a configuration of fluorescence confocal microscope equipped with a solid immersion lens and its application for acquisition of high-resolution fluorescence maps of light-harvesting complexes. Introduction of the SIL significantly increases the spatial resolution ( 600 nm) and light collection efficiency, allowing for imaging of molecule aggregates with low fluorescence intensity. The results presented in this work strongly suggest that this experimental setup is capable for studying phenomena occurring in hybrid nanostructured systems on a single molecule level. Acknowledgements Financial support from the WELCOME program Hybrid nanostructures as a stepping-stone towards efficient artificial photosynthesis awarded by the Foundation for Polish Science is gratefully acknowledged. References Figure 6. Cross-sections along lines A and B shown in Fig. 5. The numbers correspond to the widths of the intensity profiles. [1] A.M. van Oijen, M. Ketelaars, J. Köhler, T.J. Aartsma, J. Schmidt, Science 285, 400 (1999) [2] Q. Wu, G.D. Feke, R.D. Grober, L.P. Ghislain, Appl. Phys. Lett. 75, 4064 (1999) [3] S. Moehl, H. Zhao, B. Dal Don, S. Wachter, H. Kalt, J. Appl. Phys. 93, 6265 (2003) [4] M. Baba, T. Sasaki, M. Yoshita, H. Akiyama, J. Appl. Phys. 85, 6923 (1999) [5] K.P. Hewaparakrama et al., Appl. Phys. Lett. 85, 5463 (2004) [6] E. Hofmann et al., Science 272, 1788 (1996) [7] D.J. Miller et al., Photosynth. Res. 86, 229 (2005) [8] T. Schulte et al., P. Natl. Acad. Sci. USA 106, 20764 (2009) [9] F.J. Kleima et al., Biophys. J. 78, 344 (2000) [10] S. Mackowski et al., Nano Lett. 8, 558 (2008) [11] S. Mackowski, J. Phys. Condens. Matter 22, 193102 (2010) [12] S. Wörmke et al., J. Fluoresc. 18, 611 (2008) [13] S. Wörmke et al., BBA-Bioenergetics 1767, 956 (2007) 299