Summary

Real-Time cAMP Dynamics in Live Cells Using the Fluorescent cAMP Difference Detector In Situ

Published: March 22, 2024
doi:

Summary

The protocol presented in this study illustrates the effectiveness of the cAMP Difference Detector In Situ in measuring cAMP through two methods. One method utilizes a 96-well plate reading spectrophotometer with HEK-293 cells. The other method demonstrates individual HASM cells under a fluorescent microscope.

Abstract

cAMP Difference Detector In Situ (cADDis) is a novel biosensor that allows for the continuous measurement of cAMP levels in living cells. The biosensor is created from a circularly permuted fluorescent protein linked to the hinge region of Epac2. This creates a single fluorophore biosensor that displays either increased or decreased fluorescence upon binding of cAMP. The biosensor exists in red and green upward versions, as well as green downward versions, and several red and green versions targeted to subcellular locations. To illustrate the effectiveness of the biosensor, the green downward version, which decreases in fluorescence upon cAMP binding, was used. Two protocols using this sensor are demonstrated: one utilizing a 96-well plate reading spectrophotometer compatible with high-throughput screening and another utilizing single-cell imaging on a fluorescent microscope. On the plate reader, HEK-293 cells cultured in 96-well plates were stimulated with 10 µM forskolin or 10 nM isoproterenol, which induced rapid and large decreases in fluorescence in the green downward version. The biosensor was used to measure cAMP levels in individual human airway smooth muscle (HASM) cells monitored under a fluorescent microscope. The green downward biosensor displayed similar responses to populations of cells when stimulated with forskolin or isoproterenol. This single-cell assay allows visualization of the biosensor location at 20x and 40x magnification. Thus, this cAMP biosensor is sensitive and flexible, allowing real-time measurement of cAMP in both immortalized and primary cells, and with single cells or populations of cells. These attributes make cADDis a valuable tool for studying cAMP signaling dynamics in living cells.

Introduction

Adenosine 3′,5′-cyclic monophosphate, cAMP, plays a central role in cellular communication and the coordination of various physiological processes. cAMP acts as a second messenger, relaying external signals from hormones, neurotransmitters, or other extracellular molecules to initiate a cascade of intracellular events1. Moreover, cAMP is intricately involved in various signaling pathways, including those associated with G-protein-coupled receptors (GPCRs) and adenylyl cyclases. Understanding the role of cAMP in cellular signaling is fundamental to unraveling the complex mechanisms that underlie normal cellular functions and the development of potential therapies for a wide range of medical conditions2.

In the past, various methods have been employed to measure cAMP directly or indirectly. These included radiolabeling of cellular ATP pools followed by column purification, HPLC, radioimmunoassays, and enzyme-linked immunoassays1,2. These legacy assays are limited by the fact that they are end-point measures, requiring a large number of samples to construct time-dependent responses. More recently, Fluorescence Resonance Energy Transfer (FRET) sensors were developed to create assays in living cells, producing real-time, dynamic data and allowing sensors to be targeted in different subcellular locations3. FRET leverages two fluorophores, one fluorescent donor, and one fluorescent acceptor that when in close proximity, the acceptor fluorophore will be excited by the donor fluorescent output. The two fluorophores most used are cyan fluorescent protein (CFP) and yellow fluorescent protein (YFP) since these have compatible excitation and emission properties. In addition to CFP and YFP, the utilization of the green fluorescent protein (GFP) and red fluorescent protein (RFP) is commonly used for FRET biosensors. cAMP FRET biosensors operate by having a donor and acceptor on opposite ends of the Epac2 cAMP binding protein. cAMP binding alters the confirmation of Epac and increases the distance between donor and acceptor fluorophores3,4. This conformational change is detected by a loss of FRET, that is, the excitation of the acceptor fluorophore by energy transferred from the donor fluorophore drops3. While a seemingly simple process, there are an abundance of limitations and issues with the FRET biosensor for cAMP research5. One of which is the selection of fluorescent proteins, for example, GFP, which can dimerize naturally, thus reducing sensitivity6. FRET-based cAMP biosensors have been targeted to specific microdomains7, but there may be limitations owing to the large size of a construct with two fluorophores6. Another significant issue is the low signal-to-noise ratio of FRET signals resulting from the overlap between excitation and emission of the fluorophore, resulting in high sampling frequency and complicating analysis of the results4,5.

Most recently, the novel biosensor (cAMP Difference Detector In Situ), cADDis has solved these and other limitations when it comes to studying the regulation of cAMP signaling8. One important improvement is the dependence on a single fluorophore. This allows for a rapid and efficient signal with a wider dynamic range and a high signal-to-noise ratio. As a result, there can be more accuracy as there is a less broad scope of wavelengths to comb through8. Like FRET probes, the biosensor has been targeted to subcellular locations, allowing research into the compartmentation theory and exploration in lipid rafts and non-rafts, and other subcellular domains9. Perhaps most important is the suitability of a single fluorophore biosensor for high-throughput screening, which has improved sensitivity and reproducibility over FRET-based biosensors. The biosensor is packaged into a BacMam vector for easy transduction of a wide array of cell types and precise control over protein expression.

Expression control via the BacMam vector can be particularly useful in assays using GPCR orthologs from different species to facilitate the interpretation of data from animal studies. Furthermore, control over receptor expression is critical for measuring the different degrees of drug efficacy (e.g., inverse agonists and partial agonists), and low levels of receptor expression are useful to mimic the low levels found in animal tissue. BacMam is a baculovirus vector that has been modified to transduce mammalian cells such as primary cell cultures and HEK-293 lines10. Dominant selectable markers allow for BacMam to provide more stability over traditional plasmid infections11. Such selective promoters allow for more efficient gene delivery and expression. In addition, adding trichostatin A (a histone deacetylase inhibitor) enhances the reporter protein levels11. Expression levels can be controlled via the titer of the BacMam virus used and should be optimized for each cell type. In the case of this biosensor, a red or green fluorescent protein is linked to the Epac at the N- and C-termini. When cAMP binds, a conformational change in the biosensor moves amino acids adjacent to the fluorescent protein. Such a shift moves the absorbance from the anionic state to the neutral state at 400 nm, thus decreasing the fluorescence.

There are 90-120 GPCRs expressed in a single cell that respond to a wide variety of neurohumoral signals12. Therefore, it can be hypothesized that at least several dozen GPCRs per cell can stimulate or inhibit cAMP through Gs or Gi coupling, respectively. While there has been progress in monitoring this second messenger in real-time, such as FRET, more efficient methods are needed. The methodology for monitoring the synthesis and degradation of cAMP signals using cADDis in real-time is presented here. The change in fluorescence can be monitored in real-time using a fluorescence plate reader for high throughput assays or using a fluorescent microscope for single-cell assays. These methods are useful for a wide array of biological questions regarding GPCR signaling via cAMP.

Protocol

The details of all the reagents and equipment used for the study are listed in the Table of Materials. 1. Plate reading spectrophotometer high-throughput assay Seeding HEK-293 cells with the cADDis BacMam vector in a 96-well plate (Day 1) Split and infect cells in a viral hood. Warm HEK media (Table 1) and 0.25% trypsin-EDTA in a 37 °C water bath. Disinfect the hood and the materials by wi…

Representative Results

The present study validated the cytosolic biosensor in both plate reader and microscope assays. Once cells expressed the biosensor, they were stimulated with either 10 µM forskolin (a direct activator of adenylyl cyclase), 10 nM isoproterenol (an agonist at ß1AR and ß2AR), or vehicle (Figure 1). The subsequent changes in fluorescence, indicative of cAMP production, were captured every 30 s. The data was transformed as the chang…

Discussion

Accurate and sensitive measurement of cAMP is crucial for understanding its role in various cellular processes and for studying the activity of cAMP-dependent signaling pathways. There are several methods commonly employed to measure cAMP levels, including ELISA, radioimmunoassay, FRET biosensors, and the GloSensor cAMP assay14,15,16,17,18. Each cAMP assay has…

Divulgaciones

The authors have nothing to disclose.

Acknowledgements

This study was supported by the National Heart, Lung, and Blood Institute (NHLBI) (HL169522).

Materials

96-well plate (clear) Fisherbrand 21-377-203
35 mm dish Greiner Bio-One 627870 Cell culture dishes with glass bottom
96-well plate  Corning 3904 Black with clear flat bottom
Antibiotic-Antimycotic (100x) Gibco 15240062 For HEK and HASM media
BZ-X fluorescence microscope Keyence
Calcium chloride (IM) Quality Biological Inc E506 For HASM media
Centrifuge tube (15 mL) Thermo Scientific 339651
DMEM (1x) Gibco 11965092 HEK media
DPBS with Mg2+ and Ca2+  Gibco 14040-133
DPBS without Mg2+ and Ca2+ Corning 14040-133
Fetal Bovine Serum (FBS) R&D systems S11195 For HEK and HASM media
Forskolin Millipore 344270 Drug
Green Down cADDis cAMP Assay Kit Montana Molecular #D0200G Reagent
Ham's F-12K  Gibco 21127022 For HASM media
HEPES (1M) Gibco 15630080 For HASM media
Isoproterenol Sigma I6504 Drug
L-glutamine 200 mM (100x) Gibco 25030-081 For HASM media
Microcentrifuge tube (2 mL) Eppendorf 22363352
Primocin Invitrogen ant-pm-1 Antibiotic for HASM media
RNAse away Thermo Scientific 700511 Reagent
Sodium hydroxide solution Sigma S2770 For HASM media
Spectrmax M5 plate reader Molecular Devices
Trichostatin A TCI America T2477 Reagent
Trypsin EDTA Gibco 25200-056 Reagent

Referencias

  1. Krishna, G., Weiss, B., Brodie, B. B. A simple, sensitive method for the assay of adenyl cyclase. J Pharmacol Exp Ther. 163 (2), 379-385 (1968).
  2. Post, S. R., Ostrom, R. S., Insel, P. A. Biochemical methods for detection and measurement of cyclic amp and adenylyl cyclase activity. Methods Mol Biol. 126, 363-374 (2000).
  3. Li, I. T., Pham, E., Truong, K. Protein biosensors based on the principle of fluorescence resonance energy transfer for monitoring cellular dynamics. Biotechnol Lett. 28 (24), 1971-1982 (2006).
  4. Deal, J., Pleshinger, D. J., Johnson, S. C., Leavesley, S. J., Rich, T. C. Milestones in the development and implementation of fret-based sensors of intracellular signals: A biological perspective of the history of fret. Cell Signal. 75, 109769 (2020).
  5. Leavesley, S. J., Rich, T. C. Overcoming limitations of fret measurements. Cytometry A. 89 (4), 325-327 (2016).
  6. Zacharias, D. A., Violin, J. D., Newton, A. C., Tsien, R. Y. Partitioning of lipid-modified monomeric GFPS into membrane microdomains of live cells. Science. 296 (5569), 913-916 (2002).
  7. Agarwal, S. R., et al. Compartmentalized cAMP signaling associated with lipid raft and non-raft membrane domains in adult ventricular myocytes. Front Pharmacol. 9, 332 (2018).
  8. Tewson, P. H., Martinka, S., Shaner, N. C., Hughes, T. E., Quinn, A. M. New dag and cAMP sensors optimized for live-cell assays in automated laboratories. J Biomol Screen. 21 (3), 298-305 (2016).
  9. Tewson, P., et al. Assay for detecting Galphai-mediated decreases in cAMP in living cells. SLAS Discov. 23 (9), 898-906 (2018).
  10. Nunez, F. J., et al. Glucocorticoids rapidly activate cAMP production via g(alphas) to initiate non-genomic signaling that contributes to one-third of their canonical genomic effects. FASEB J. 34 (2), 2882-2895 (2020).
  11. Condreay, J. P., Witherspoon, S. M., Clay, W. C., Kost, T. A. Transient and stable gene expression in mammalian cells transduced with a recombinant baculovirus vector. Proc Natl Acad Sci U S A. 96 (1), 127-132 (1999).
  12. Insel, P. A., et al. G protein-coupled receptor (GPCR) expression in native cells: "Novel" endogpcrs as physiologic regulators and therapeutic targets. Mol Pharmacol. 88 (1), 181-187 (2015).
  13. Nunez, F. J., et al. Agonist-specific desensitization of PGE(2)-stimulated cAMP signaling due to upregulated phosphodiesterase expression in human lung fibroblasts. Naunyn Schmiedebergs Arch Pharmacol. 393 (2), 843-856 (2020).
  14. Ojiaku, C. A., et al. Transforming growth factor-beta1 decreases beta(2)-agonist-induced relaxation in human airway smooth muscle. Am J Respir Cell Mol Biol. 61 (2), 209-218 (2019).
  15. Zuo, H., et al. Cigarette smoke up-regulates pde3 and pde4 to decrease cAMP in airway cells. Br J Pharmacol. 175 (14), 2988-3006 (2018).
  16. Cullum, S. A., Veprintsev, D. B., Hill, S. J. Kinetic analysis of endogenous beta(2) -adrenoceptor-mediated cAMP glosensor responses in hek293 cells. Br J Pharmacol. 180 (2), 1304-1315 (2023).
  17. Liu, X., Sun, S. Q., Ostrom, R. S. Fibrotic lung fibroblasts show blunted inhibition by cAMP due to deficient cAMP response element-binding protein phosphorylation. J Pharmacol Exp Ther. 315 (2), 678-687 (2005).
  18. Bogard, A. S., Birg, A. V., Ostrom, R. S. Non-raft adenylyl cyclase 2 defines a cAMP signaling compartment that selectively regulates il-6 expression in airway smooth muscle cells: Differential regulation of gene expression by ac isoforms. Naunyn Schmiedebergs Arch Pharmacol. 387 (4), 329-339 (2014).
  19. Schmidt, M., Cattani-Cavalieri, I., Nunez, F. J., Ostrom, R. S. Phosphodiesterase isoforms and cAMP compartments in the development of new therapies for obstructive pulmonary diseases. Curr Opin Pharmacol. 51, 34-42 (2020).
  20. Ostrom, K. F., et al. Physiological roles of mammalian transmembrane adenylyl cyclase isoforms. Physiol Rev. 102 (2), 815-857 (2022).
  21. Auld, D. S., et al. Characterization of chemical libraries for luciferase inhibitory activity. J Med Chem. 51 (8), 2372-2386 (2008).
  22. De Logu, F., et al. Schwann cell endosome CGRP signals elicit periorbital mechanical allodynia in mice. Nat Commun. 13 (1), 646 (2022).
  23. Wray, N. H., Schappi, J. M., Singh, H., Senese, N. B., Rasenick, M. M. NMDAR-independent, cAMP-dependent antidepressant actions of ketamine. Mol Psychiatry. 24 (12), 1833-1843 (2019).
  24. Thornquist, S. C., Pitsch, M. J., Auth, C. S., Crickmore, M. A. Biochemical evidence accumulates across neurons to drive a network-level eruption. Mol Cell. 81 (4), 675-690 (2021).
  25. Zhang, S. X., et al. Hypothalamic dopamine neurons motivate mating through persistent cAMP signalling. Nature. 597 (7875), 245-249 (2021).
  26. Lutas, A., Fernando, K., Zhang, S. X., Sambangi, A., Andermann, M. L. History-dependent dopamine release increases cAMP levels in most basal amygdala glutamatergic neurons to control learning. Cell Rep. 38 (4), 110297 (2022).
  27. Zhang, C., et al. Area postrema cell types that mediate nausea-associated behaviors. Neuron. 109 (3), 461-472 (2021).
check_url/es/66451?article_type=t

Play Video

Citar este artículo
Cattani-Cavalieri, I., Margolis, J., Anicolaesei, C., Nuñez, F. J., Ostrom, R. S. Real-Time cAMP Dynamics in Live Cells Using the Fluorescent cAMP Difference Detector In Situ. J. Vis. Exp. (205), e66451, doi:10.3791/66451 (2024).

View Video