Open Access
14 November 2023 Fiber photometry-based investigation of brain function and dysfunction
Nicole Byron, Shuzo Sakata
Author Affiliations +
Abstract

Fiber photometry is an optical method to monitor fluorescent signals using a fiber optic cannula. Over the past two decades, together with the development of various genetically encoded biosensors, it has been applied to investigate various types of activity in the central nervous system. This includes not only type-specific neuronal population activity, but also non-neuronal activity and neurotransmitter/neuropeptide signals in awake, freely behaving animals. In this perspective, we summarize the recent development of this technique. After describing common technical pitfalls, we discuss future directions of this powerful approach for investigating brain function and dysfunction.

1.

Introduction

The development of neural interface technologies has been an active research area.1,2 Historically, this field can be dated back to Luigi Galvani’s experiments in the 18th century. While electrophysiology continues to be the gold standard for monitoring individual neuronal activity in living brain tissue at high temporal resolution, optical approaches have unique advantages over electrophysiology. Optical monitoring and manipulation of neuronal activity have been routinely performed in a cell-type-specific fashion by expressing genetically encoded actuators and sensors.38 Out of various optical approaches, fiber photometry offers a simple, but powerful solution to monitor type-specific neuronal population activity in freely behaving animals.

Fiber photometry was first introduced to neuroscience in 2005.9 The advent of genetically encoded calcium indicators (GECIs) allows fiber photometry to monitor cell type-specific population activity from deep brain regions in freely behaving mice.10,11 Fiber photometry has been widely adopted to characterize neural population activity in behaving animals along with opto- and chemo-genetic experiments over the past two decades (Fig. 1).

Fig. 1

Increased interest in fiber photometry over recent years. Total number of hits from PubMed, Web of Science, and Google Scholar when searching either “fiber photometry” or “fibre photometry” across all available years.

NPh_11_S1_S11502_f001.png

Fiber photometry typically involves two major components (Fig. 2): fluorescent indicators and optical devices. The former can be either chemical indicators or genetically encoded sensors. While the pioneering study used calcium-sensitive dyes,9 GCaMPs are the most popular choice [Fig. 2(c)]. Genetically engineered voltage indicators have also been deployed to monitor fast neural oscillations.13,14 Over the past 5 years, the use of genetically encoded sensors for neurotransmitters and neuromodulators has gained popularity.5,15

Fig. 2

General principle of fiber photometry in neuroscience. (a) Principle of fiber photometry. Light from a light source is passed through a fiber optic implant to excite fluorescent indicators. Emitted light passes back through the fiber optic implant and is collected at a photodetector. Signals will be transferred to a computer for data analysis. All systems will be varied and include other optical elements. (b) Commonly used flat multimode fiber (left) and newly developed and established tapered optic fiber12 (right). Images of the tapered optic fibers light profile illustrate the depth resolution capacity. (c) GCaMP fluorescence occurs from a fluorescence protein (circularly permuted enhanced green fluorescent protein, cpEGFP) after calcium binds.

NPh_11_S1_S11502_f002.png

In typical photometry experiments, fiber optics delivers and collects photons in brain tissue [Fig. 2(a)]. While conventional multimode fibers are commonly used, a tapered fiber allows depth-resolved light delivery and collection in freely behaving mice, with reduced tissue damage [Fig. 2(b)].12,16 Applications of micro-photodiodes are also on the horizon.17,18 Thus, by combining genetically encoded sensors and advanced optical technologies, fiber photometry is a highly versatile approach to investigate neural functions.

In this perspective, we summarize two emerging applications: fiber photometry in non-neuronal cells and multi-modal monitoring of neural activity. After commenting on potential technical pitfalls, we will discuss future directions of fiber photometry.

2.

Fiber Photometry-Based Investigation of Non-Neuronal Cell Activity

The nervous system consists of not only neurons, but also glial cells. Astrocytes are the most abundant glial cells in the brain and play diverse roles, including cognitive functions.1921 Despite the importance of astrocytes, monitoring astrocyte activity was impractical for many years as they are not electrically excitable, meaning common electrophysiological methods were not applicable. However, their ability to exhibit rich intracellular calcium signals means they can be interrogated using new optical tools that allow for fluorescent indicators to be expressed by astrocytes. Thus, GECIs have been applied to study astrocyte function in vivo.22

Recently, Tsunematsu and her colleagues23 utilized YC-nano50 to investigate region-specific and state-dependent calcium dynamics in astrocytes across sleep-wake cycles. YC-nano50 is a different category of GECIs since two fluorescent proteins are encoded and fluorescent signals are emitted based on Förster resonance energy transfer (FRET).24,25 Thus, YC-nano50 allows ratiometric photometry by monitoring two fluorescent signals with a single excitation wavelength.

By expressing YC-nano50 across the brain by crossing two transgenic mouse lines, Tsunematsu and her colleagues23 monitored calcium signals across multiple brain regions and sleep-wake cycles. Previously, astrocyte calcium concentration was thought to increase during wakefulness and decrease during sleep.26 However, they discovered that although cortical and cerebellar astrocytes exhibit state-dependent calcium signals as predicted, astrocytes in the hypothalamus and pons exhibited distinct patterns: calcium signals remain high even during non-rapid eye movement sleep whereas they decrease during rapid eye movement (REM) sleep.

Although this study demonstrated region-specific and state-dependent astrocytic calcium dynamics for the first time, there are at least three limitations in this study. First, pH influences the fluorescence properties of fluorophores differently and pH can change throughout the sleep-wake cycle.27,28 Ratiometric measurements of fluorescent signals across different vigilance states should be interpreted carefully. Indeed, although several experimental variables (e.g., stereotaxic coordinates and astrocyte classes) are not the same, using GCaMPs, a recent study observed state-dependent calcium dynamics in the pons with higher calcium signals in both wakefulness and REM sleep.29 This discrepancy must be resolved in the near future by monitoring astrocyte calcium signals across multiple brainstem nuclei. It may be worth referencing GFP signals to evaluate factors independent of calcium signals. Second, since astrocytes are diverse cell populations,3032 the findings cannot be generalized. It is imperative to characterize calcium signals across different astrocyte types. Finally, photometry provides calcium signals only from cell populations. Because astrocyte calcium signals differ depending on cellular locations,33,34 monitoring calcium signals by optical means with high spatial resolution is required in the future. Despite these limitations and challenges, detailed characterization of astrocyte calcium dynamics across multiple brain regions and astrocyte types will lead to a better understanding of brain functions in health and disease.

3.

Simultaneous In Vivo Electrophysiology and Fiber Photometry

A major drawback of GCaMP-based optical monitoring is its low temporal resolution. Genetically encoded voltage indicators have been combined with fiber photometry,13,14 but their adoption is difficult in a conventional laboratory setting. Therefore, combining electrophysiology can be an alternative.

Patel and her colleagues took a simple approach to utilize an “optrode” (a hybrid of an optical fiber and wire electrode) to complement each other. Here, an electrode monitored fast field potentials while an optic fiber allowed cell-type-specific photometry.35 As an application, they focused on pontine waves (P-waves). P-waves or ponto-geniculo-occipital waves are 100-ms long brain waves and a major electrophysiological marker of REM sleep along with theta oscillations. Despite the prominence of P-waves, their existence in mice was anecdotal for decades. Leveraging their own discovery of P-waves in mice,36 they examined if mesopontine cholinergic neurons are involved in mouse P-waves as implicated in other species. Indeed, they found that cholinergic transients are associated with P-waves.35,37

While this particular study was a simple combination of conventional electrophysiology and fiber photometry, other recent studies have monitored spiking activity.38,39 This type of combination of electrophysiology and photometry creates an opportunity where either approach alone cannot answer a particular scientific question.

More specifically, although electrophysiology allows monitoring neuronal signals at microsecond resolution, cell type classification is typically based on spike waveforms. Thus, it is challenging or impractical to extract cell type-specific activity. On the other hand, although GCaMP-based photometry allows monitoring of cholinergic activity, it cannot detect sub-second neural events like P-waves or other neural oscillations, which can be easily detected electrophysiologically. Thus, the combination of electrophysiological and photometry complements each other and creates the opportunity to investigate cell type-specific activity in the context of a certain brain state.

4.

Technical Considerations

Although fiber photometry is easy to implement in many laboratories, non-specific signals must be taken into consideration to accurately monitor fluorescent signals. Non-specific signals may originate from two sources: movement artifacts and autofluorescence (AF). Since a fiber optic cannula is rigid and fixed to the skull, the relative displacement between the brain tissue and fiber tip results in artificial fluorescent changes. Additionally, the fiber optic is susceptible to subtle bending, which can alter the light output. This can cause artifacts within signal collection. Therefore, it is essential to correct this. Practically, two options can be considered: first, an FRET-based sensor can be employed.23,25 The second and commonly used approach is to use isosbestic illumination, which provides signals independent of indicators.35,40 If such control signals are not measured, careful data interpretation is essential.

The second major source of non-specific signals is AF. The source of AF can be from brain tissue, fiber optics and patch cables.41 While a low-AF patch cable is commercially available, regular photobleaching of a patch cable before recording is recommended. Additionally, since the emission spectra of brain tissue AF overlaps with GCaMPs emission spectra,42,43 background fluorescence intensity also decreases as a function of time because of AF photobleaching. Experiments in which several seconds of trials are repeated are less likely to be affected by tissue AF since ΔF/F can be easily obtained. On the other hand, hours-long recordings require not just a high signal-to-noise ratio, but also additional baseline adjustment. Thus, although fiber photometry is a versatile approach, implementation, and signal processing require a basic understanding of the principles of the method.

5.

Future Directions

While fiber photometry has provided an innovative solution to monitor neurobiological signals in deep brain structures, we expect at least four major domains to be developed further in coming years.

The first are sensors. We expect further development of genetically encoded sensors for a wide variety of biological targets.5,15,4447 In particular, the development of neurotransmitters and neuropeptides has been successful.5,15,4852 These emerging tools have expanded not just to various neurotransmitters and neuropeptides, but also to their excitation/emission spectrum.50,52,53 These tools now allow us to monitor multiple neurotransmitters simultaneously in behaving animals.54,55 It would be interesting to expand the repertoire to other extracellular molecules, such as cytokines and metabolites. In particular, tracking pathological molecular events in brain disorders, such as Alzheimer’s disease, will provide an opportunity to develop and optimize treatment strategies. In addition to monitoring extracellular molecules, optical monitoring of intracellular signaling is also an exciting avenue. For example, Sabatini and his colleagues56 elegantly demonstrated cell type-specific dopamine-induced protein kinase A (PKA) signaling in vivo. This approach is based on fluorescence lifetime photometry (FLiP).57,58 Since the fluorescence lifetime of a fluorophore reflects the microenvironment of fluorophores, such as pH changes, ion concentrations, or molecular interactions,5961 the combination of FLiP with various sensors is a promising direction.

Second, we also expect further development of optical devices. Tapered fibers have already demonstrated depth-resolved photometry in freely behaving animals.12 Although the current limitation of this approach is its depth coverage and spatial resolution, it may be interesting to place miniaturized light sources, such as μLED arrays,62 along the fiber. Another direction is to adopt microfabrication technologies to integrate both μLED and photodiodes.17,18 Pioneering work has already demonstrated wireless battery-free photometry systems.18 Ultimately, microfabrication technologies may allow us to monitor fluorescent signals from individual neurons by deploying single-photon avalanche diode arrays.17 In addition, adopting a beam-shaping approach is an attractive possibility.63

Third, as we discussed above, combining other modalities, such as electrophysiology, is critical to compensate for the drawback of photometry technology.35,38,64,65 In addition to acquiring signals in the brain, controlling neural and non-neuronal signals optically or by other means is also a crucial area to explore.

Finally, since an advantage of photometry is its minimal invasiveness to target deep brain tissue, increasing the scalability is critical to achieving brain-wide functional mapping. Although several pioneering studies have tackled this challenge,40,66,67 implants and optical devices can be developed further to obtain volumetric signals across the brain. More specifically, because conventional flat fiber-based approaches only allow monitoring activity at the tip of the fiber, depth-resolved photometry is critical.12 Scaling up of channels not just in a horizontal plane, but also in a vertical plane is required for brain-wide functional mapping.

In conclusion, since 2005, fiber photometry has been deployed for many experiments. In conjunction with the development of novel genetically encoded sensors and other technologies, fiber photometry will keep playing a significant role in a better understanding of brain function and dysfunction.

Disclosures

Authors report no conflict of interest.

Code, Data, and Materials Availability

Data sharing is not applicable.

Acknowledgments

This work was supported by MRC (Grant No. MR/V033964/1) and the European Union’s Horizon 2020 (Grant Nos. H2020-ICT, DEEPER, and 101016787) to S.S.

References

1. 

J. Rivnay et al., “Next-generation probes, particles, and proteins for neural interfacing,” Sci. Adv., 3 (6), e1601649 https://doi.org/10.1126/sciadv.1601649 STAMCV 1468-6996 (2017). Google Scholar

2. 

J. A. Frank, M. J. Antonini and P. Anikeeva, “Next-generation interfaces for studying neural function,” Nat. Biotechnol., 37 (9), 1013 –1023 https://doi.org/10.1038/s41587-019-0198-8 NABIF9 1087-0156 (2019). Google Scholar

3. 

V. Emiliani et al., “Optogenetics for light control of biological systems,” Nat. Rev. Methods Prim., 2 (1), 55 https://doi.org/10.1038/s43586-022-00136-4 (2022). Google Scholar

4. 

M. Z. Lin and M. J. Schnitzer, “Genetically encoded indicators of neuronal activity,” Nat. Neurosci., 19 (9), 1142 –1153 https://doi.org/10.1038/nn.4359 NANEFN 1097-6256 (2016). Google Scholar

5. 

Z. Wu, D. Lin and Y. Li, “Pushing the frontiers: tools for monitoring neurotransmitters and neuromodulators,” Nat. Rev. Neurosci., 23 (5), 257 –274 https://doi.org/10.1038/s41583-022-00577-6 NRNAAN 1471-003X (2022). Google Scholar

6. 

O. Yizhar et al., “Optogenetics in neural systems,” Neuron, 71 (1), 9 –34 https://doi.org/10.1016/j.neuron.2011.06.004 NERNET 0896-6273 (2011). Google Scholar

7. 

E. S. Boyden et al., “Millisecond-timescale, genetically targeted optical control of neural activity,” Nat. Neurosci., 8 (9), 1263 –1268 https://doi.org/10.1038/nn1525 NANEFN 1097-6256 (2005). Google Scholar

8. 

J. Nakai, M. Ohkura and K. Imoto, “A high signal-to-noise Ca2+ probe composed of a single green fluorescent protein,” Nat. Biotechnol., 19 (2), 137 –141 https://doi.org/10.1038/84397 NABIF9 1087-0156 (2001). Google Scholar

9. 

H. Adelsberger, O. Garaschuk and A. Konnerth, “Cortical calcium waves in resting newborn mice,” Nat. Neurosci., 8 (8), 988 –990 https://doi.org/10.1038/nn1502 NANEFN 1097-6256 (2005). Google Scholar

10. 

L. A. Gunaydin et al., “Natural neural projection dynamics underlying social behavior,” Cell, 157 (7), 1535 –1551 https://doi.org/10.1016/j.cell.2014.05.017 CELLB5 0092-8674 (2014). Google Scholar

11. 

H. Lutcke et al., “Optical recording of neuronal activity with a genetically-encoded calcium indicator in anesthetized and freely moving mice,” Front Neural Circuits, 4 9 https://doi.org/10.3389/fncir.2010.00009 (2010). Google Scholar

12. 

F. Pisano et al., “Depth-resolved fiber photometry with a single tapered optical fiber implant,” Nat. Methods, 16 (11), 1185 –1192 https://doi.org/10.1038/s41592-019-0581-x 1548-7091 (2019). Google Scholar

13. 

M. Kannan et al., “Fast, in vivo voltage imaging using a red fluorescent indicator,” Nat. Methods, 15 (12), 1108 –1116 https://doi.org/10.1038/s41592-018-0188-7 1548-7091 (2018). Google Scholar

14. 

J. D. Marshall et al., “Cell-type-specific optical recording of membrane voltage dynamics in freely moving mice,” Cell, 167 (6), 1650 –1662.e15 https://doi.org/10.1016/j.cell.2016.11.021 CELLB5 0092-8674 (2016). Google Scholar

15. 

M. A. Labouesse and T. Patriarchi, “A versatile GPCR toolkit to track in vivo neuromodulation: not a one-size-fits-all sensor,” Neuropsychopharmacology, 46 (12), 2043 –2047 https://doi.org/10.1038/s41386-021-00982-y NEROEW 0893-133X (2021). Google Scholar

16. 

F. Pisanello et al., “Multipoint-emitting optical fibers for spatially addressable in vivo optogenetics,” Neuron, 82 (6), 1245 –1254 https://doi.org/10.1016/j.neuron.2014.04.041 NERNET 0896-6273 (2014). Google Scholar

17. 

L. C. Moreaux et al., “Integrated neurophotonics: toward dense volumetric interrogation of brain circuit activity-at depth and in real time,” Neuron, 108 (1), 66 –92 https://doi.org/10.1016/j.neuron.2020.09.043 NERNET 0896-6273 (2020). Google Scholar

18. 

A. Burton et al., “Wireless, battery-free subdermally implantable photometry systems for chronic recording of neural dynamics,” Proc. Natl. Acad. Sci. U. S. A., 117 (6), 2835 –2845 https://doi.org/10.1073/pnas.1920073117 (2020). Google Scholar

19. 

J. Nagai et al., “Specific and behaviorally consequential astrocyte G(q) GPCR signaling attenuation in vivo with ibetaARK,” Neuron, 109 (14), 2256 –2274.e9 https://doi.org/10.1016/j.neuron.2021.05.023 NERNET 0896-6273 (2021). Google Scholar

20. 

P. Kofuji and A. Araque, “Astrocytes and behavior,” Annu. Rev. Neurosci., 44 49 –67 https://doi.org/10.1146/annurev-neuro-101920-112225 ARNSD5 0147-006X (2021). Google Scholar

21. 

A. Doron et al., “Hippocampal astrocytes encode reward location,” Nature, 609 (7928), 772 –778 https://doi.org/10.1038/s41586-022-05146-6 (2022). Google Scholar

22. 

X. Yu, J. Nagai and B. S. Khakh, “Improved tools to study astrocytes,” Nat. Rev. Neurosci., 21 (3), 121 –138 https://doi.org/10.1038/s41583-020-0264-8 NRNAAN 1471-003X (2020). Google Scholar

23. 

T. Tsunematsu et al., “Region-specific and state-dependent astrocyte Ca2+ dynamics during the sleep-wake cycle in mice,” J. Neurosci., 41 (25), 5440 –5452 https://doi.org/10.1523/JNEUROSCI.2912-20.2021 JNRSDS 0270-6474 (2021). Google Scholar

24. 

K. Horikawa et al., “Spontaneous network activity visualized by ultrasensitive Ca2+ indicators, yellow Cameleon-Nano,” Nat. Methods, 7 (9), 729 –732 https://doi.org/10.1038/nmeth.1488 1548-7091 (2010). Google Scholar

25. 

K. Kanemaru et al., “In vivo visualization of subtle, transient, and local activity of astrocytes using an ultrasensitive Ca2+ indicator,” Cell Rep., 8 (1), 311 –318 https://doi.org/10.1016/j.celrep.2014.05.056 (2014). Google Scholar

26. 

M. G. Frank, “The role of Glia in sleep regulation and function,” Sleep-Wake Neurobiology and Pharmacology, 83 –96 Springer International Publishing, Cham (2019). Google Scholar

27. 

D. B. Betolngar et al., “pH sensitivity of FRET reporters based on cyan and yellow fluorescent proteins,” Anal. Bioanal. Chem., 407 (14), 4183 –4193 https://doi.org/10.1007/s00216-015-8636-z ABCNBP 1618-2642 (2015). Google Scholar

28. 

Y. Ikoma et al., “Properties of REM sleep alterations with epilepsy,” Brain, 146 (6), 2431 –2442 https://doi.org/10.1093/brain/awac499 BRAIAK 0006-8950 (2023). Google Scholar

29. 

W. Peng et al., “Adenosine-independent regulation of the sleep-wake cycle by astrocyte activity,” Cell Discov., 9 (1), 16 https://doi.org/10.1038/s41421-022-00498-9 (2023). Google Scholar

30. 

R. de Ceglia et al., “Specialized astrocytes mediate glutamatergic gliotransmission in the CNS,” Nature, 622 (7981), 120 –129 https://doi.org/10.1038/s41586-023-06502-w (2023). Google Scholar

31. 

F. Endo et al., “Molecular basis of astrocyte diversity and morphology across the CNS in health and disease,” Science, 378 (6619), eadc9020 https://doi.org/10.1126/science.adc9020 SCIEAS 0036-8075 (2022). Google Scholar

32. 

A. Zeisel et al., “Molecular architecture of the mouse nervous system,” Cell, 174 (4), 999 –1014.e22 https://doi.org/10.1016/j.cell.2018.06.021 CELLB5 0092-8674 (2018). Google Scholar

33. 

L. Bojarskaite et al., “Astrocytic Ca2+ signaling is reduced during sleep and is involved in the regulation of slow wave sleep,” Nat. Commun., 11 (1), 3240 https://doi.org/10.1038/s41467-020-17062-2 NCAOBW 2041-1723 (2020). Google Scholar

34. 

A. M. Ingiosi et al., “A role for astroglial calcium in mammalian sleep and sleep regulation,” Curr. Biol., 30 (22), 4373 –4383.e7 https://doi.org/10.1016/j.cub.2020.08.052 CUBLE2 0960-9822 (2020). Google Scholar

35. 

A. A. Patel et al., “Simultaneous electrophysiology and fiber photometry in freely behaving mice,” Front. Neurosci., 14 148 https://doi.org/10.3389/fnins.2020.00148 1662-453X (2020). Google Scholar

36. 

T. Tsunematsu et al., “State-dependent brainstem ensemble dynamics and their interactions with hippocampus across sleep states,” Elife, 9 e52244 https://doi.org/10.7554/eLife.52244 (2020). Google Scholar

37. 

T. Tsunematsu et al., “Pontine waves accompanied by short hippocampal sharp wave-ripples during non-rapid eye movement sleep,” Sleep, 46 (9), zsad193 https://doi.org/10.1093/sleep/zsad193 SLEED6 0161-8105 (2023). Google Scholar

38. 

B. Spagnolo et al., “Tapered fibertrodes for optoelectrical neural interfacing in small brain volumes with reduced artefacts,” Nat. Mater., 21 (7), 826 –835 https://doi.org/10.1038/s41563-022-01272-8 NMAACR 1476-1122 (2022). Google Scholar

39. 

A. A. Legaria et al., “Fiber photometry in striatum reflects primarily nonsomatic changes in calcium,” Nat. Neurosci., 25 (9), 1124 –1128 https://doi.org/10.1038/s41593-022-01152-z NANEFN 1097-6256 (2022). Google Scholar

40. 

C. K. Kim et al., “Simultaneous fast measurement of circuit dynamics at multiple sites across the mammalian brain,” Nat. Methods, 13 (4), 325 –328 https://doi.org/10.1038/nmeth.3770 1548-7091 (2016). Google Scholar

41. 

A. Formozov, A. Dieter and J. S. Wiegert, “A flexible and versatile system for multi-color fiber photometry and optogenetic manipulation,” Cell Rep. Methods, 3 (3), 100418 https://doi.org/10.1016/j.crmeth.2023.100418 (2023). Google Scholar

42. 

K. C. Reinert et al., “Flavoprotein autofluorescence imaging in the cerebellar cortex in vivo,” J. Neurosci. Res., 85 (15), 3221 –3232 https://doi.org/10.1002/jnr.21348 JNREDK 0360-4012 (2007). Google Scholar

43. 

B. Chance, “Optical method,” Annu. Rev. Biophys. Biophys. Chem., 20 1 –28 https://doi.org/10.1146/annurev.bb.20.060191.000245 ARBCEY 0883-9182 (1991). Google Scholar

44. 

Y. Zhang et al., “Fast and sensitive GCaMP calcium indicators for imaging neural populations,” Nature, 615 (7954), 884 –891 https://doi.org/10.1038/s41586-023-05828-9 (2023). Google Scholar

45. 

H. Tian et al., “Video-based pooled screening yields improved far-red genetically encoded voltage indicators,” Nat. Methods, 20 1082 –1094 https://doi.org/10.1038/s41592-022-01743-5 1548-7091 (2023). Google Scholar

46. 

J. Platisa et al., “High-speed low-light in vivo two-photon voltage imaging of large neuronal populations,” Nat Methods, 20 1095 –1103 https://doi.org/10.1038/s41592-023-01820-3 (2023). Google Scholar

47. 

A. S. Abdelfattah et al., “Sensitivity optimization of a rhodopsin-based fluorescent voltage indicator,” Neuron, 111 (10), 1547 –1563.e9 https://doi.org/10.1016/j.neuron.2023.03.009 NERNET 0896-6273 (2023). Google Scholar

48. 

F. Sun et al., “A genetically encoded fluorescent sensor enables rapid and specific detection of dopamine in flies, fish, and mice,” Cell, 174 (2), 481 –496.e19 https://doi.org/10.1016/j.cell.2018.06.042 CELLB5 0092-8674 (2018). Google Scholar

49. 

T. Patriarchi et al., “Ultrafast neuronal imaging of dopamine dynamics with designed genetically encoded sensors,” Science, 360 (6396), eaat4422 https://doi.org/10.1126/science.aat4422 SCIEAS 0036-8075 (2018). Google Scholar

50. 

Z. Kagiampaki et al., “Sensitive multicolor indicators for monitoring norepinephrine in vivo,” Nat. Methods, 20 (9), 1426 –1436 https://doi.org/10.1038/s41592-023-01959-z 1548-7091 (2023). Google Scholar

51. 

L. Duffet et al., “A genetically encoded sensor for in vivo imaging of orexin neuropeptides,” Nat. Methods, 19 (2), 231 –241 https://doi.org/10.1038/s41592-021-01390-2 1548-7091 (2022). Google Scholar

52. 

F. Sun et al., “Next-generation GRAB sensors for monitoring dopaminergic activity in vivo,” Nat. Methods, 17 (11), 1156 –1166 https://doi.org/10.1038/s41592-020-00981-9 1548-7091 (2020). Google Scholar

53. 

T. Patriarchi et al., “An expanded palette of dopamine sensors for multiplex imaging in vivo,” Nat. Methods, 17 (11), 1147 –1155 https://doi.org/10.1038/s41592-020-0936-3 1548-7091 (2020). Google Scholar

54. 

A. C. Krok et al., “Intrinsic dopamine and acetylcholine dynamics in the striatum of mice,” Nature, 621 (7979), 543 –549 https://doi.org/10.1038/s41586-023-05995-9 (2023). Google Scholar

55. 

L. Chantranupong et al., “Dopamine and glutamate regulate striatal acetylcholine in decision-making,” Nature, 621 (7979), 577 –585 https://doi.org/10.1038/s41586-023-06492-9 (2023). Google Scholar

56. 

S. J. Lee et al., “Cell-type-specific asynchronous modulation of PKA by dopamine in learning,” Nature, 590 (7846), 451 –456 https://doi.org/10.1038/s41586-020-03050-5 (2021). Google Scholar

57. 

S. J. Lee et al., “Monitoring behaviorally induced biochemical changes using fluorescence lifetime photometry,” Front. Neurosci., 13 766 https://doi.org/10.3389/fnins.2019.00766 1662-453X (2019). Google Scholar

58. 

G. Cui et al., “Concurrent activation of striatal direct and indirect pathways during action initiation,” Nature, 494 (7436), 238 –242 https://doi.org/10.1038/nature11846 (2013). Google Scholar

59. 

R. Datta et al., “Fluorescence lifetime imaging microscopy: fundamentals and advances in instrumentation, analysis, and applications,” J. Biomed. Opt., 25 (7), 071203 https://doi.org/10.1117/1.JBO.25.7.071203 JBOPFO 1083-3668 (2020). Google Scholar

60. 

K. Suhling et al., “Fluorescence lifetime imaging (FLIM): basic concepts and some recent developments,” Med. Photonics, 27 3 –40 https://doi.org/10.1016/j.medpho.2014.12.001 (2015). Google Scholar

61. 

R. Yasuda, “Imaging spatiotemporal dynamics of neuronal signaling using fluorescence resonance energy transfer and fluorescence lifetime imaging microscopy,” Curr. Opin. Neurobiol., 16 (5), 551 –561 https://doi.org/10.1016/j.conb.2006.08.012 COPUEN 0959-4388 (2006). Google Scholar

62. 

R. Scharf et al., “Depth-specific optogenetic control in vivo with a scalable, high-density muLED neural probe,” Sci. Rep., 6 28381 https://doi.org/10.1038/srep28381 SRCEC3 2045-2322 (2016). Google Scholar

63. 

M. Stiburek et al., “110  μm thin endo-microscope for deep-brain in vivo observations of neuronal connectivity, activity and blood flow dynamics,” Nat. Commun., 14 (1), 1897 https://doi.org/10.1038/s41467-023-36889-z NCAOBW 2041-1723 (2023). Google Scholar

64. 

C. Liu et al., “A wireless, implantable optoelectrochemical probe for optogenetic stimulation and dopamine detection,” Microsyst. Nanoeng., 6 64 https://doi.org/10.1038/s41378-020-0176-9 (2020). Google Scholar

65. 

Z. Ramezani, K. J. Seo and H. Fang, “Hybrid electrical and optical neural interfaces,” J. Micromech. Microeng., 31 (4), 044002 https://doi.org/10.1088/1361-6439/abeb30 JMMIEZ 0960-1317 (2021). Google Scholar

66. 

Y. Sych et al., “High-density multi-fiber photometry for studying large-scale brain circuit dynamics,” Nat. Methods, 16 (6), 553 –560 https://doi.org/10.1038/s41592-019-0400-4 1548-7091 (2019). Google Scholar

67. 

Z. Guo et al., “Neural dynamics in the limbic system during male social behaviors,” Neuron, 111 (20), 3288 –3306 https://doi.org/10.1016/j.neuron.2023.07.011 NERNET 0896-6273 (2023). Google Scholar

Biography

Nicole Byron is a PhD student at Dr. Shuzo Sakata’s Laboratory at the University of Strathclyde, United Kingdom. She earned her MSci degree in biochemistry with distinction from the University of Strathclyde in 2020. Her current research focuses on optical interrogation of Alzheimer’s disease pathology. She utilizes state-of-the-art tapered fiber-based photometry to optically monitor Alzheimer’s pathology in freely behaving mice.

Shuzo Sakata is a reader at the University of Strathclyde, United Kingdom. He earned his PhD in biophysics from Kyoto University in 2002. His research group applies a wide range of in vivo approaches to investigate state-dependent and cell-type-specific information processing in the brain. He is also a Scientific Advisory Board member of NeuroVLC. The topic in this perspective is not related to this company.

CC BY: © The Authors. Published by SPIE under a Creative Commons Attribution 4.0 International License. Distribution or reproduction of this work in whole or in part requires full attribution of the original publication, including its DOI.
Nicole Byron and Shuzo Sakata "Fiber photometry-based investigation of brain function and dysfunction," Neurophotonics 11(S1), S11502 (14 November 2023). https://doi.org/10.1117/1.NPh.11.S1.S11502
Received: 22 August 2023; Accepted: 17 October 2023; Published: 14 November 2023
Advertisement
Advertisement
KEYWORDS
Photometry

Brain

Calcium

Optical sensing

Electrophysiology

Fiber optics

Sensors

Back to Top