Next Article in Journal
Fluoxetine Decreases the Proliferation and Adipogenic Differentiation of Human Adipose-Derived Stem Cells
Next Article in Special Issue
Validation of FRET Assay for the Screening of Growth Inhibitors of Escherichia coli Reveals Elongasome Assembly Dynamics
Previous Article in Journal
14-3-3γ Regulates Lipopolysaccharide-Induced Inflammatory Responses and Lactation in Dairy Cow Mammary Epithelial Cells by Inhibiting NF-κB and MAPKs and Up-Regulating mTOR Signaling
Previous Article in Special Issue
Homo-FRET Based Biosensors and Their Application to Multiplexed Imaging of Signalling Events in Live Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Genetically Encoded FRET-Sensor Based on Terbium Chelate and Red Fluorescent Protein for Detection of Caspase-3 Activity

1
A. N. Bach Institute of Biochemistry, Russian Academy of Sciences, 119071 Moscow, Russia
2
M. V. Lomonosov Moscow State University, Department of Chemistry, 119991 Moscow, Russia
3
Skryabin Institute of Biochemistry and Physiology of Microorganisms, Russian Academy of Sciences, 142290 Pushchino, Russia
4
Branch of Shemyakin and Ovchinnikov Institute of Bioorganic Chemistry, Russian Academy of Sciences, 142290 Pushchino, Russia
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Int. J. Mol. Sci. 2015, 16(7), 16642-16654; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms160716642
Submission received: 30 April 2015 / Revised: 30 June 2015 / Accepted: 6 July 2015 / Published: 22 July 2015
(This article belongs to the Special Issue Förster Resonance Energy Transfer (FRET) 2015)

Abstract

:
This article describes the genetically encoded caspase-3 FRET-sensor based on the terbium-binding peptide, cleavable linker with caspase-3 recognition site, and red fluorescent protein TagRFP. The engineered construction performs two induction-resonance energy transfer processes: from tryptophan of the terbium-binding peptide to Tb3+ and from sensitized Tb3+ to acceptor—the chromophore of TagRFP. Long-lived terbium-sensitized emission (microseconds), pulse excitation source, and time-resolved detection were utilized to eliminate directly excited TagRFP fluorescence and background cellular autofluorescence, which lasts a fraction of nanosecond, and thus to improve sensitivity of analyses. Furthermore the technique facilitates selective detection of fluorescence, induced by uncleaved acceptor emission. For the first time it was shown that fluorescence resonance energy transfer between sensitized terbium and TagRFP in the engineered construction can be studied via detection of microsecond TagRFP fluorescence intensities. The lifetime and distance distribution between donor and acceptor were calculated using molecular dynamics simulation. Using this data, quantum yield of terbium ions with binding peptide was estimated.

1. Introduction

Nowadays biosensors based on the Förster resonance energy transfer (FRET) phenomenon are widely used to study the enzymatic activity in living cells [1]. Change of the FRET efficiency allows direct monitoring of enzymatic activity, protein-protein interactions, conformational changes in proteins, etc. [2,3,4,5].
In FRET-sensors, energy transfer represents a dynamic type of fluorescence quenching of the donor and simultaneous increase of the acceptor fluorescence, respectively. Therefore, FRET can be detected by (1) a decrease of the donor’s fluorescence intensity; (2) an increase of the acceptor fluorescence intensity; (3) in dual channel mode measurements (at two wavelengths)—by change of the ratio of donor and acceptor fluorescence intensities. All these measurements can be carried out not only in the stationary mode (fluorescence intensity), but also in time-resolved mode, because FRET is also characterized by a decrease of the donor fluorescence lifetime and increase of the appeared acceptor fluorescence lifetime [4,5,6,7]. FRET efficiency measurement using fluorescence lifetime change is the most accurate and reliable method, since the fluorescence lifetime is independent of concentration of fluorophores, photobleaching, changes in the intensity of the excitation light, and light scattering [4,5].
Use of fluorescent complexes of lanthanides, which have a microsecond fluorescence, as donors in FRET-pair, and spectroscopy with time delay allows us to eliminate short-lived background signals associated with the autofluorescence of biomolecules and scattered light [3,5,8]. In addition, measurements with time delay detect only sensitized fluorescence of the acceptor. It solves the problem of simultaneous fluorescence excitation of the donor and acceptor thereby increasing the dynamic range of measurements and the accuracy of the FRET efficiency determination [5].
Lanthanide-based FREТ sensors were used to study interactions between proteins [9] and subunits of oligomeric proteins [10,11,12]; to detect the conformational changes induced by the binding of activators and repressors [13], DNA hybridization in solutions, and cell extracts and in cell culture [14,15,16]; to analyze the enzymatic activity [17,18,19,20,21,22,23], for example the catalytic activity of key enzymes of apoptosis—caspases.
Apoptosis is a genetically regulated process of cell death, which contributes to normal functioning of the organism through the destruction of useless, diseased, and mutated cells which are potentially dangerous and the ones which have finished their life cycle. Dysregulation of apoptosis underlies many neurodegenerative, neoplastic, and autoimmune diseases [24].
In most cases apoptosis is associated with the activation of caspases—a family of proteolytic enzymes that specifically cleave proteins at the N-terminus of aspartic acid residues. Caspase-3 is a main effector caspase uniting mitochondrial and receptor pathways of apoptosis. In addition, it is expressed in all tissues and involved in almost all the currently known cycles of amplification [25]. Therefore, the caspase-3 activity is a useful marker for determining the aggressiveness of pathological processes and evaluation of efficiency of apoptosis-inducing drugs.
Our laboratory has developed a genetically encoded FRET sensor for caspase-3 Tb3+-TBP-19-TagRFP [26,27], which consists of the complex of Tb3+ with terbium-binding peptide YIDTNNDGWYEGDELLA [28] as a donor, flexible linker VDGGSGGDEVDGWGGSGLD [29] with the caspase-3 cleavage site DEVD [30], which was used earlier in the FRET-pair with a non-fluorescent acceptor [31], and red fluorescent protein TagRFP [26,32] as an acceptor. There are two donor-acceptor pairs in the resulting structure, which transfer energy using the fluorescence resonance mechanism. Initially, energy transfers from the tryptophan in the terbium-binding peptide, which sensitizes terbium ion, and then from the terbium ion to the TagRFP chromophore (Figure 1).
Figure 1. Structure of Tb3+-TBP-19-TagRFP sensor. Carbon atoms are shown in green, nitrogen atoms are shown in blue, and sulphur atom is shown in yellow.
Figure 1. Structure of Tb3+-TBP-19-TagRFP sensor. Carbon atoms are shown in green, nitrogen atoms are shown in blue, and sulphur atom is shown in yellow.
Ijms 16 16642 g001
The fluorescence of the sensor is in the orange spectral region that can reduce the contribution of background fluorescence in the total intensity. In addition, unlike previously developed lanthanide sensors, in which cryptand-lanthanide complex was a donor, and allophycocyanin or synthetic fluorophores as acceptor [15,16], the Tb3+-TBP-19-TagRFP sensor is genetically encoded and requires only the additional introduction of terbium ions, which on one hand allows for highly standardized samples for in vitro measurements, and on the other hand in the long term makes its use in experiments on living cells possible.

2. Results and Discussion

2.1. Molecular Dynamics Simulations

We performed 200 ns molecular dynamics simulation of the linker connecting energy donor and acceptor in a water box to study its conformational dynamics. MD simulation demonstrates that the linker is rather flexible resulting in a wide distribution of distances between donor and acceptor (Figure 2a). FRET efficiency E and energy transfer rate constant kT as well as lifetime of donor in the presence of acceptor τDA were calculated on the fly for each frame of MD trajectory (Figure 2b). Lifetime distribution τDA is broad and corresponds to the mean value 0.19 ms that is in good agreement with the experimentally observed value. FRET efficiency calculated as a mean value from all frames is 43% (compared with 44% from experiment).
Figure 2. (a) Distance distribution between donor and acceptor (blue curve) and corresponding cumulative probability (red curve); (b) lifetime distribution τDA (blue curve) and corresponding cumulative probability (red curve).
Figure 2. (a) Distance distribution between donor and acceptor (blue curve) and corresponding cumulative probability (red curve); (b) lifetime distribution τDA (blue curve) and corresponding cumulative probability (red curve).
Ijms 16 16642 g002

2.2. Determination of Oligomeric State of the Protein

Using the dynamic light scattering technique, the effective diffusion radii of the sensor sample and their corresponding molecular masses were obtained. Sample protein TBP-19-TagRFP contains molecules with a radius of 3.5 and 2.6 nm, which in the spherical approximation correspond to the molecular mass of 62 and 33 kDa, respectively.
The oligomeric state of the protein TBP-19-TagRFP was also determined using gel-filtration chromatography. The target protein was collected by absorption at 560 nm. In the chromatogram there were two peaks, first one corresponded to the molecular mass of 60 kDa, the second—to 28 kDa.
Molecular weight of the protein, calculated from the amino acid sequence, is 30.5 kDa. The data obtained in two independent experiments gave the same results—the protein TBP-19-TagRFP is a mixture of dimeric and monomeric forms.

2.3. Fluorimetric Titration of the TBP-19-TagRFP Protein Using Tb(NO3)3 Solution

To measure the sensor signal, fluorimetric titration of terbium binding sites was performed. The time delay technique in such experiments (Figure 3) permits selective measurement of the fluorescence of the sensor without the influence of background autofluorescence and scattered light. In the limit of a short acceptor lifetime, gated measurement after a delay time of 100 µs the fluorescence of directly excited acceptor is negligible, while the acceptor emission resulting from energy transfer displays almost the same lifetime as the quenched donor [7].
Figure 3. The principle of fluorescence spectroscopy with a time delay. (1) is the light flash; (2) is the fluorescence of the donor in the presence of acceptor; (3) is the sensitized fluorescence of acceptor; (4) is the fluorescence of the donor without the acceptor.
Figure 3. The principle of fluorescence spectroscopy with a time delay. (1) is the light flash; (2) is the fluorescence of the donor in the presence of acceptor; (3) is the sensitized fluorescence of acceptor; (4) is the fluorescence of the donor without the acceptor.
Ijms 16 16642 g003
Registration of spectra was performed with a time delay of 100 μs, allowing to exclude the contribution of the direct excitation of the acceptor to the total intensity, and to measure only the fluorescence of the acceptor due to the energy transfer. In addition, the time delay cut off the short-lived background fluorescence [26].
However, at high concentration of the sensor its fluorescence intensity is lower than expected because of the reabsorption effect, which leads to distortion of the short-wave part of the TagRFP emission spectrum, where absorption of the sample is particularly high (Figure 4):
If the fluorescence path is considered from the center of the cuvette, the corrected fluorescence intensity is determined by the ratio (1):
I corr = I × 10 D 2  
where Icorr is the corrected fluorescence intensity, I is measured fluorescence intensity, D is the optical density of the solution.
The results of such correction are shown in the Figure 5. The intensity of corrected spectra (Figure 5b) is higher than initial one (Figure 5a).
Figure 4. Overlay of absorption and fluorescence spectra of the Tb3+-TBP-19-TagRFP sensor. The absorption spectrum is shown by the blue line, the original spectrum of the fluorescence is shown by the black line, and the corrected fluorescence spectrum is shown by the red line.
Figure 4. Overlay of absorption and fluorescence spectra of the Tb3+-TBP-19-TagRFP sensor. The absorption spectrum is shown by the blue line, the original spectrum of the fluorescence is shown by the black line, and the corrected fluorescence spectrum is shown by the red line.
Ijms 16 16642 g004
Figure 5. Fluorescence spectra of the Tb3+-TBP-19-TagRFP sensor before (a) and after (b) the correction for reabsorption. Measurements were performed using 280 nm excitation wavelength and a time delay of 100 μs, gate 1 ms in the phosphorescence mode. Spectra are arranged one above the other with the terbium concentration increase from 1 to 50 µM.
Figure 5. Fluorescence spectra of the Tb3+-TBP-19-TagRFP sensor before (a) and after (b) the correction for reabsorption. Measurements were performed using 280 nm excitation wavelength and a time delay of 100 μs, gate 1 ms in the phosphorescence mode. Spectra are arranged one above the other with the terbium concentration increase from 1 to 50 µM.
Ijms 16 16642 g005aIjms 16 16642 g005b
In order to calculate the TagRFP fluorescence without the contribution of terbium fluorescence, it was necessary to carry out the titration of terbium-binding peptide linked to the TagRFP protein with photobleached chromophore using Tb(NO3)3 solution as described in [26]. The results of that titration are shown in Figure 6.
Figure 6. Fluorescence spectra of the Tb3+-TBP-19-TagRFP sensor with photobleached TagRFP chromophore. Measurements were performed using 280 nm excitation wavelength and a time delay of 100 μs, gate 1 ms in the phosphorescence mode. Spectra are arranged one above the other with the terbium concentration increase from 1 to 50 µM.
Figure 6. Fluorescence spectra of the Tb3+-TBP-19-TagRFP sensor with photobleached TagRFP chromophore. Measurements were performed using 280 nm excitation wavelength and a time delay of 100 μs, gate 1 ms in the phosphorescence mode. Spectra are arranged one above the other with the terbium concentration increase from 1 to 50 µM.
Ijms 16 16642 g006
The fluorescence intensity of TagRFP (Figure 7a) was calculated by subtraction of intensity of the photobleached sample (Figure 6), which lacks the signal from the TagRFP chromophore, from the intensity of the initial non-bleached sample (Figure 5).
Figure 7. (a) Spectrum of the sensitized fluorescence of TagRFP in the Tb3+-TBP-19-TagRFP sensor; (b) The dependence of sensitized fluorescence intensity of TagRFP, which was calculated by subtraction of intensity of the photobleached sample from the intensity of the non-bleached sample corrected for reabsorption, on the Tb concentration at 606 nm.
Figure 7. (a) Spectrum of the sensitized fluorescence of TagRFP in the Tb3+-TBP-19-TagRFP sensor; (b) The dependence of sensitized fluorescence intensity of TagRFP, which was calculated by subtraction of intensity of the photobleached sample from the intensity of the non-bleached sample corrected for reabsorption, on the Tb concentration at 606 nm.
Ijms 16 16642 g007
As a result, we showed that there is an energy transfer in the Tb3+-TBP-19-TagRFP sensor from the terbium ion to the TagRFP chromophore. For the first time we obtained the spectrum of the sensitized TagRFP fluorescence with time delay for the previously described sensor [26,27]. Using the titration curve of TBP-19-TagRFP at 606 nm (Figure 7b), which was obtained by subtraction of intensity of the photobleached sample from the intensity of the corrected for reabsorption non-bleached sample, the dissociation constant of Tb3+ and TBP-19-TagRFP complex was calculated. We performed nonlinear curve fit of the titration curve in Origin 9.0.0 software using Equation (2):
I = I max [Tb] K d + [Tb]
where I is measured fluorescence intensity, Imax (which was the fitting parameter) is the maximum fluorescence intensity at the saturating concentration, [Tb] is the Tb3+ concentration and Kd is the dissociation constant.
The saturating concentration was found as the intersection of tangents to the linear parts of the fitting curve (Figure 7b) and is equal to 24.2 µM. Imax is 95 ± 16 and the calculated Kd value is 17 ± 7 µM.

3. Experimental Section

3.1. Molecular Dynamics Simulation

We applied the NAMD program [33] to perform molecular dynamics simulations. We manually constructed oligopeptide representing linker between donor and acceptor in the inline conformation. The structure was protonated according to the basic principles of valences. Linker VDGGSGGDEVDGWGGSGLD was solvated in a rectangular water box 69 × 53 × 48 Å3 and neutralized by adding 0.2 M NaCl. Protein was described with CHARMM27 force field and water molecules with TIP3P model [34,35]. All long range electrostatic interactions were computed by using the particle mesh Ewald method [36]. Constant temperature MD simulations were performed for the NPT ensemble at 300 K using the Langevin thermostat. Following a 2000 step energy minimization, the MD simulations were carried out with a 2 fs integration time step and the total length of 200 ns. Rigid bonds approximation was imposed for the whole system. The VMD program was used for the visualization of the results [37].

3.2. Calculation of FRET Parameters

All FRET parameters, i.e., efficiency E, rate constant of energy transfer kT, and lifetime of donor in the presence of acceptor τDA were calculated according to the conventional formulae of FRET theory [7] on the fly for each MD frame. The procedure was as follows. The distance between donor and acceptor was extracted from each frame and kT was calculated as:
k T ( r ) = Q D κ 2 τ D r 6 ( 9000 ( l n 10 ) 128 π 5 N n 4 ) J ( λ )
where r, J(λ), N and n are distance between donor and acceptor, overlap integral, Avogadro’s number, and refractive index, respectively. In assumption of a freely rotational donor and acceptor, we used orientation factor κ2 equal to 2/3. FRET efficiency was calculated for each frame from the corresponding kT value as:
E = k T τ D 1 + k T
The lifetime of donor in the presence of acceptor τDA was calculated as:
τ D A = τ D ( 1 E )
Quantum yield of donor QD value was estimated from the experimental lifetime of the donor without acceptor τD = 0.33 ms [26] and radiative lifetime of Tb3+ τD,R = 3.5 ms [38] as 0.094. Overlap integral was calculated from the experimental data and equals to 3.22591 × 1015 М−1·cm−1·nm4.

3.3. Determination of Oligomeric State of the Protein

Determination of oligomeric state was carried out using the hydrodynamic radii data obtained on DynaPro Titan (Wyatt Technology, Goleta, CA, USA) and Dynamics 6.6.7.9 software (Wyatt Technology, Goleta, CA, USA). The sample protein was in 20 mM·Tris-HCl, pH 7.5 buffer at 25 °C, the intensity of laser irradiation was set to 100%. For the interpretation of the results we used a Globular protein model.
The oligomeric state of the sensor was also determined using the molecular mass values obtained by gel-filtration chromatography. The separation of components was performed on the Superdex 200 10/300 GL column using ÄKTA Purifier (GE Healthcare, Little Chalfont, UK). The elution buffer was 20 mM Tris-HCl, pH 7.5. Registration of absorption was performed at two wavelengths—280 and 560 nm. The protein was collected by absorption at 560 nm.

3.4. Spectrophotometric Analysis

Absorption spectra were measured using Cary 300 (Varian, Palo Alto, CA, USA) spectrophotometer and 3 mm quartz microcuvette. Sample was in 20 mM Tris-HCl, pH 7.5 buffer.

3.5. Spectrofluorimetric Analysis

All samples were in 20 mm Tris-HCl, pH 7.5 buffer, measurements were performed in a 3 mm quartz microcuvette. 23.4 µM sample protein was titrated using the terbium nitrate (III) in 0.01 M HCl (AppliChem, Darmstadt, Germany) to a final concentration of 50 μM. Cary Eclipse spectrofluorometer (Varian, Palo Alto, CA, USA) was used for fluorescence measurements. Measurements were performed using 280 nm excitation wavelength and a time delay of 100 μs, gate 1 ms in the phosphorescence mode, excitation and emission slits were set to 20 and 5 nm, respectively. TagRFP photobleaching was carried out using a 532 nm laser with 105 mW power. Sample irradiation was continued until TagRFP fluorescence reached zero.

4. Conclusions

For the first time we have detected sensitized TagRFP fluorescence using microsecond time gated technique. This approach allows to solve the problem of direct acceptor excitation due to spectral cross-talk between donor and acceptor. In Ref. [26] FRET efficiency of 44% was detected by change of donor (terbium ion) fluorescence lifetime from 0.33 ms for photobleached form of Tb3+-TBP-19-TagRFP to 0.18 ms for initial form. This method also allows the exclusion of signal cross-talk, but fluorescence lifetime of terbium ion can vary because of changes in its coordination sphere [39], decreasing the accuracy of the results. Direct detection of sensitized TagRFP fluorescence using time gated approach allows us to get rid of both problems mentioned above, because we only measure the signal that is connected to FRET, while the background fluorescence and signal from directly excited TagRFP were already gone to zero during the delay time. We calculated the Kd value of the Tb3+ and TBP-19-TagRFP complex and found it to be 17 ± 7 µM.
The molecular dynamics modeling allowed us to calculate the distribution of possible distances between the donor and acceptor in Tb3+-TBP-19-TagRFP sensor. Calculated FRET efficiency is equal to 43% which in good agreement with the experimental value of 44%.

Acknowledgments

We thank Alexander Nemukhin for advice and discussion of molecular dynamics simulation.
This work was supported by Russian Foundation for Basic Research, grants No. 13-04-40305-Н and 11-02-01532-а.

Author Contributions

Alexander S. Goryashchenko and Anna A. Bochkova—experimental work; Maria G. Khrenova—molecular dynamics simulation and calculation of FRET parameters; Tatiana V. Ivashina and Leonid M. Vinokurov—expression and purification of TBP-19-TagRFP protein; Alexander P. Savitsky—leadership of the work process.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Forster, T. Delocalized excitation and excitation transfer. In Modern Quantum Chemistry; Sinanoglu, O., Ed.; Academic Press: New York, NY, USA, 1965; pp. 93–138. [Google Scholar]
  2. Li, I.T.; Pham, E.; Truong, K. Protein biosensors based on the principle of fluorescence resonance energy transfer for monitoring cellular dynamics. Biotechnol. Lett. 2006, 28, 1971–1982. [Google Scholar] [CrossRef] [PubMed]
  3. Selvin, P.R. The renaissance of fluorescence resonance energy transfer. Nat. Struct. Biol. 2000, 7, 730–734. [Google Scholar] [CrossRef] [PubMed]
  4. Suhling, K.; French, P.M.W.; Phillips, D. Time-resolved fluorescence microscopy. Photochem. Photobiol. Sci. 2005, 4, 13–22. [Google Scholar] [CrossRef] [PubMed]
  5. Zherdeva, V.V.; Savitsky, A.P. Using lanthanide-based energy transfer for in vitro and in vivo studies of biological processes. Biochemistry 2012, 77, 1553–1574. [Google Scholar] [CrossRef] [PubMed]
  6. Vogel, S.S.; Thaler, C.; Koushik, S.V. Fanciful Fret. Sci. STKE 2006, 331, re2. [Google Scholar] [CrossRef] [PubMed]
  7. Maliwal, B.P.; Gryczynski, Z.; Lacowicz, J.R. Long-wavelength long-lifetime luminophores. Anal. Chem. 2001, 73, 4277–4285. [Google Scholar] [CrossRef] [PubMed]
  8. Soini, E.; Hemmila, I. Fluoroimmunoassay-present status and key problems. Clin. Chem. 1979, 25, 353–361. [Google Scholar] [PubMed]
  9. Leifert, W.R.; Bailey, K.; Cooper, T.H.; Aloia, A.L.; Glatz, R.V.; McMurchie, E.J. Measurement of heterotrimeric G-protein and regulators of G-protein signaling interactions by time-resolved fluorescence resonance energy transfer. Anal. Biochem. 2006, 355, 201–212. [Google Scholar] [CrossRef] [PubMed]
  10. Guo, W.; Urizar, E.; Kralikova, M.; Mobarec, J.C.; Shi, L.; Filizola, M.; Javitch, J.A. Dopamine D2 receptors form higher order oligomers at physiological expression levels. EMBO J. 2008, 27, 2293–2304. [Google Scholar] [CrossRef] [PubMed]
  11. Wang, J.; Norcross, M. Dimerization of chemokine receptors in living cells: Key to receptor function and novel targets for therapy. Drug Discov. Today 2008, 13, 625–632. [Google Scholar] [CrossRef] [PubMed]
  12. Ramsay, D.; Carr, I.C.; Pediani, J.; Lopez-Gimenez, J.F.; Thurlow, R.; Fidock, M.; Milligan, G. High-affinity interactions between human α1A-adrenoceptor C-terminal splice variants produce homo- and heterodimers but do not generate the α1L adrenoceptor. Mol. Pharmacol. 2004, 66, 228–239. [Google Scholar] [CrossRef] [PubMed]
  13. Kupcho, K.R.; Stafslien, D.K.; DeRosier, T.; Hallis, T.M.; Ozers, M.S.; Vogel, K.W. Simultaneous monitoring of discrete binding events using dual-acceptor terbium-based LRET. J. Am. Chem. Soc. 2007, 129, 13372–13373. [Google Scholar] [CrossRef] [PubMed]
  14. Sueda, S.; Yuan, J.; Matsumoto, K. A homogeneous DNA hybridization system by using a new luminescence terbium chelate. Bioconj. Chem. 2002, 13, 200–205. [Google Scholar] [CrossRef]
  15. Mathis, G. Probing molecular interactions with homogeneous techniques based on rare earth cryptates and fluorescence energy transfer. Clin. Chem. 1995, 41, 1391–1397. [Google Scholar] [PubMed]
  16. Laitala, V.; Ylikoski, A.; Raussi, H.M.; Ollikka, P.; Hemmila, I. Time-resolved detection probe for homogeneous nucleic acid analyses in one-step format. Anal. Biochem. 2007, 361, 126–131. [Google Scholar] [CrossRef] [PubMed]
  17. Riddle, S.M.; Vedvik, K.L.; Hanson, G.T.; Vogel, K.W. Time-resolved fluorescence resonance energy transfer kinase assays using physiological protein substrates: Applications of terbium-fluorescein and terbium-green fluorescent protein fluorescence resonance energy transfer pairs. Anal. Biochem. 2006, 356, 108–116. [Google Scholar] [CrossRef] [PubMed]
  18. Vuojola, J.; Lamminmäki, U.; Soukka, T. Resonance energy transfer from lanthanide chelates to overlapping and nonoverlapping fluorescent protein acceptors. Anal. Chem. 2009, 81, 5033–5038. [Google Scholar] [CrossRef] [PubMed]
  19. Robers, M.B.; Machleidt, T.; Carlson, C.B.; Bi, K. Cellular LanthaScreen and beta-lactamase reporter assays for high-throughput screening of JAK2 inhibitors. Assay Drug Dev. Technol. 2008, 6, 519–529. [Google Scholar] [CrossRef] [PubMed]
  20. Carlson, C.B.; Robers, M.B.; Vogel, K.W.; Machleidt, T. Development of LanthaScreen cellular assays for key components within the PI3K/AKT/mTOR pathway. J. Biomol. Screen. 2009, 14, 121–132. [Google Scholar] [CrossRef] [PubMed]
  21. Carlson, C.B.; Mashock, M.J.; Bi, K. BacMam-enabled LanthaScreen cellular assays for PI3K/Akt pathway compound profiling in disease-relevant cell backgrounds. J. Biomol. Screen. 2010, 15, 327–334. [Google Scholar] [CrossRef] [PubMed]
  22. Enkvist, E.; Vaasa, A.; Kasari, M.; Kriisa, M.; Ivan, T.; Ligi, K.; Raidaru, G.; Uri, A. Protein-induced long lifetime luminescence of nonmetal probes. ACS Chem. Biol. 2011, 6, 1052–1062. [Google Scholar] [CrossRef] [PubMed]
  23. Kasari, M.; Ligi, K.; Williams, J.A.G.; Vaasa, A.; Enkvist, E.; Viht, K.; Pålsson, L.-O.; Uri, A. Responsive microsecond-lifetime photoluminescent probes for analysis of protein kinases and their inhibitors. Biochim. Biophys. Acta 2013, 1834, 1330–1335. [Google Scholar] [CrossRef] [PubMed]
  24. Poon, I.K.H.; Lucas, C.D.; Rossi, A.G.; Ravichandran, K.S. Apoptotic cell clearance: Basic biology and therapeutic potential. Nat. Rev. Immunol. 2014, 14, 166–180. [Google Scholar] [CrossRef] [PubMed]
  25. McIlwain, D.R.; Berger, T.; Mak, T.W. Caspase functions in cell death and disease. Cold Spring Harb. Perspect. Biol. 2013, 5, a008656. [Google Scholar] [CrossRef] [PubMed]
  26. Arslanbaeva, L.R.; Zherdeva, V.V.; Ivashina, T.V.; Vinokurov, L.M.; Morozov, V.B.; Olenin, A.N.; Savitskii, A.P. Induction-resonance energy transfer between the terbium-binding peptide and the red fluorescent proteins DsRed2 and TagRFP. Biophysics 2011, 56, 381–386. [Google Scholar] [CrossRef]
  27. Arslanbaeva, L.R.; Zherdeva, V.V.; Ivashina, T.V.; Vinokurov, L.M.; Rusanov, A.L.; Savitskiĭ, A.P. Genetically encoded FRET-pair on the basis of terbium-binding peptide and red fluorescent protein. Appl. Biochem. Microbiol. 2010, 46, 154–158. [Google Scholar] [CrossRef]
  28. Nitz, M.; Franz, K.J.; Maglathlin, R.L.; Imperiali, B. A powerful combinatorial screen to identify high-affinity terbium(III)-binding peptides. Chembiochem 2003, 4, 272–276. [Google Scholar] [CrossRef] [PubMed]
  29. Xu, X.; Gerard, A.L.; Huang, B.C.; Anderson, D.C.; Payan, D.G.; Luo, Y. Detection of programmed cell death using fluorescence energy transfer. Nucleic Acids Res. 1998, 26, 2034–2035. [Google Scholar] [CrossRef] [PubMed]
  30. Thornberry, N.A.; Rano, T.A.; Peterson, E.P.; Rasper, D.M.; Timkey, T.; Garcia-Calvo, M.; Houtzager, V.M.; Nordstrom, P.A.; Roy, S.; Vaillancourt, J.P.; et al. A combinatorial approach defines specificities of members of the caspase family and granzyme B. J. Biol. Chem. 2007, 272, 17907–17911. [Google Scholar] [CrossRef]
  31. Savitsky, A.P.; Rusanov, A.L.; Zherdeva, V.V.; Gorodnicheva, T.V.; Khrenova, M.G.; Nemukhin, A.V. FLIM-FRET imaging of caspase-3 activity in live cells using pair of red fluorescent proteins. Theranostics 2012, 2, 215–226. [Google Scholar] [CrossRef] [PubMed]
  32. Merzlyak, E.M.; Goedhart, J.; Shcherbo, D.; Bulina, M.E.; Shcheglov, A.S.; Fradkov, A.F.; Gaintzeva, A.; Lukyanov, K.A.; Lukyanov, S.; Gadella, T.W.; et al. Bright monomeric red fluorescent protein with an extended fluorescence lifetime. Nat. Methods 2007, 4, 555–557. [Google Scholar] [CrossRef] [PubMed]
  33. Phillips, J.C.; Braun, R.; Wang, W.; Gumbart, J.; Tajkhorshid, E.; Villa, E.; Chipot, C.; Skeel, R.D.; Kale, L.; Schulten, K. Scalable molecular dynamics with NAMD. J. Comput. Chem. 2005, 26, 1781–1802. [Google Scholar] [CrossRef] [PubMed]
  34. MacKerell, A.D.; Bashford, D.; Bellott, M.; Dunbrack, R.L.; Evanseck, J.D.; Field, M.J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S.; et al. All-atom empirical potential for molecular modeling and dynamics studies of proteins. J. Phys. Chem. B 1998, 102, 3586–3616. [Google Scholar] [CrossRef] [PubMed]
  35. Mackerell, A.D.; Feig, M.; Brooks, C.L. Extending the treatment of backbone energetics in protein force fields: Limitations of gas-phase quantum mechanics in reproducing protein conformational distributions in molecular dynamics simulations. J. Comput. Chem. 2004, 25, 1400–1415. [Google Scholar] [CrossRef] [PubMed]
  36. Darden, T.A.; York, D.M.; Pedersen, L.G. Particle mesh ewald: An log(N) method for ewald sums in large systems. J. Chem. Phys. 1993, 98, 10089–10092. [Google Scholar] [CrossRef]
  37. Humphrey, W.; Dalke, A.; Schulten, K. VMD—Visual molecular dynamics. J. Mol. Graphics 1996, 14, 33–38. [Google Scholar] [CrossRef]
  38. Sardar, D.K.; Nash, K.L.; Yow, R.M.; Gruber, J.B.; Valiev, U.V.; Kokanyan, E.P. Absorption intensities and emission cross sections of Tb3+ (4f8) in TbAlO3. J. Appl. Phys. 2006, 100, 083108. [Google Scholar] [CrossRef]
  39. Zaitoun, M.A.; Al-Tarawneh, S. Effect of varying lanthanide local coordination sphere on luminescence properties illustrated by selected inorganic and organic rare earth complexes synthesized in sol–gel host glasses. J. Lumin. 2011, 131, 1795–1801. [Google Scholar] [CrossRef]

Share and Cite

MDPI and ACS Style

Goryashchenko, A.S.; Khrenova, M.G.; Bochkova, A.A.; Ivashina, T.V.; Vinokurov, L.M.; Savitsky, A.P. Genetically Encoded FRET-Sensor Based on Terbium Chelate and Red Fluorescent Protein for Detection of Caspase-3 Activity. Int. J. Mol. Sci. 2015, 16, 16642-16654. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms160716642

AMA Style

Goryashchenko AS, Khrenova MG, Bochkova AA, Ivashina TV, Vinokurov LM, Savitsky AP. Genetically Encoded FRET-Sensor Based on Terbium Chelate and Red Fluorescent Protein for Detection of Caspase-3 Activity. International Journal of Molecular Sciences. 2015; 16(7):16642-16654. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms160716642

Chicago/Turabian Style

Goryashchenko, Alexander S., Maria G. Khrenova, Anna A. Bochkova, Tatiana V. Ivashina, Leonid M. Vinokurov, and Alexander P. Savitsky. 2015. "Genetically Encoded FRET-Sensor Based on Terbium Chelate and Red Fluorescent Protein for Detection of Caspase-3 Activity" International Journal of Molecular Sciences 16, no. 7: 16642-16654. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms160716642

Article Metrics

Back to TopTop