Open Access
6 March 2024 Multimode fiber endoscopes for computational brain imaging
Author Affiliations +
Abstract

Advances in imaging tools have always been a pivotal driver for new discoveries in neuroscience. An ability to visualize neurons and subcellular structures deep within the brain of a freely behaving animal is integral to our understanding of the relationship between neural activity and higher cognitive functions. However, fast high-resolution imaging is limited to sub-surface brain regions and generally requires head fixation of the animal under the microscope. Developing new approaches to address these challenges is critical. The last decades have seen rapid progress in minimally invasive endo-microscopy techniques based on bare optical fibers. A single multimode fiber can be used to penetrate deep into the brain without causing significant damage to the overlying structures and provide high-resolution imaging. Here, we discuss how the full potential of high-speed super-resolution fiber endoscopy can be realized by a holistic approach that combines fiber optics, light shaping, and advanced computational algorithms. The recent progress opens up new avenues for minimally invasive deep brain studies in freely behaving mice.

1.

Introduction

The field of neurophotonics is built upon continuous technological advances in light microscopy. Ever since Antoni van Leeuwenhoek observed cells under an early microscope, optical microscopy remains the key instrument in neuroscience.1 Technological breakthroughs in optical imaging are constantly changing the way neural circuits can be examined and visualized.2 Modern benchtop systems provide high-resolution multifunctional imaging but require head fixation of the animal under the microscope objective,3 which is unfortunately incompatible with many behavioral studies. Miniaturized microscopes have been developed to enable measurements in freely moving animals.4,5 However, miniscopes as well as many other state-of-the-art optical techniques work well only at surface or sub-surface areas up to a few hundred micrometers in depth,6 but neurophotonics research requires high-resolution images in vivo in deeper layers of the brain.7 Gradient-index (GRIN) lenses are used for deep imaging but due to the relatively large size the implantation includes tissue removal.8 Optical imaging at truly unlimited depths has been enabled by minimally invasive endo-microscopy based on bare optical fibers.9 The most straightforward approach of lensless fiber-imaging is to use fiber bundles where each fiber transmits one pixel of an image.1012 However, the spatial resolution is relatively low due to large core-to-core spacing.

To summarize, the problem of understanding the relationship between neural activity in deep brain structures and unrestrained behavior remains unsolved. One of the most promising research directions addressing this challenge is novel imaging approaches based on a multimode fiber (MMF).13 The full potential of minimally invasive MMF endoscopy for neuroimaging—ultimate performance when used for deep-tissue imaging—can only be realized by combining optimal probes, advanced light control, and computational post-processing algorithms.

2.

Multimode Fiber Endoscopes

An MMF is a flexible waveguide that simultaneously supports tens to thousands of guided modes propagating with different speeds.14 The interesting feature of light transmission through an MMF is that the process is highly complex and seemingly random but yet linear and deterministic. Coupling coherent light to an MMF results in a complex interference pattern with diffraction-limited features known as speckles but the information is only scrambled and not lost.15 The recent emergence of computational holography and wavefront shaping allowed for precise manipulation of the speckles by controlling the incident wavefront with a spatial light modulator.16,17 The capability to engineer an optical field at the MMF output to any desired pattern,18 e.g., a focal spot, as shown in Fig. 1(a), provides a new imaging modality: raster-scan imaging of a tissue on the MMF output facet.19,20 A conventional MMF can now be utilized as an ultra-thin (usually about 100  μm in diameter) aberration-free imaging probe.2124 Multimode light propagation via a single core guarantees the best spatial resolution for a given footprint.25,26 However, it remains diffraction-limited, meaning that some subcellular structures cannot be visualized since diffraction of light blurs them to a single feature, as shown in Fig. 1(c).

Fig. 1

(a) Wavefront engineering on the MMF input using a spatial light modulator creates a focal spot on the fiber output facet. (b) Random patterns created within an MMF represent nearly ideal illumination for computational compressive sensing. (c), (d) Images of fluorescent beads 1.5  μm in diameter obtained through an MMF probe (NA = 0.1, pump wavelength, λ=532  nm) by (c) raster scan wavefront shaping-based endo-microscopy, which is unable to resolve beads that are <1.5  μm apart and (d) computational compressive sensing with super-resolution. The scale bars are equal to the diffraction limit λ/(2NA)=2.66  μm. Images in (c) and (d) are adapted from Ref. 27.

NPh_11_S1_S11509_f001.png

In the most popular raster scan imaging approach, the ability to image a large brain region with high optical resolution is achieved at the cost of an acquisition speed. Sequential scanning of every point of interest puts a technological but still very hard limit on temporal resolution, and it is important to visualize rapid interactions between different elements of complex neuronal networks. Another issue of the state-of-the-art MMF imaging probes is its extreme sensitivity to external perturbations, such as fiber bending, movements, and temperature drifts. Even small changes in the fiber configuration destroy the imaging abilities. Therefore, despite a significant effort, use of an MMF as a high-resolution flexible probe is still far away.15,2832 To fully exploit the complexity of light transport through an MMF for neuroimaging, the new technological insights are needed. Most likely it will be based on the emergence of smart and powerful computing algorithms.

3.

Optical Imaging in the Age of Computation

Computation is becoming an integral part of imaging systems, giving rise to a new concept of designing the hardware and software components together.33 A joint optimization of optical setups and computational algorithms opens up new ways to overcome state-of-the-art limits of optical microscopy. While nearly all conventional signal acquisition protocols are based on the famous Nyquist–Shannon theorem (the sampling rate must be at least twice the maximum frequency of the signal), surprisingly, we rarely use all the information acquired. Storing images with lossy codecs, such as jpeg, in fact, discards the majority of acquired data. Implementing the compression already at the signal acquisition step leads to faster imaging. Computational compressive sensing facilitates signal acquisition with a large reduction in sampling for signals that have a sparse representation, vastly reducing the number of measurements beyond the Nyquist limit.34,35

Compressive sensing enables super-resolution imaging.36,37 The mechanism behind sub-diffraction compressive imaging relies on computational bandwidth interpolation. In contrast to alternative computational approaches that mainly fail because of noise, compression-based interpolation serves as an effective way for rejecting common noise types.38 As a result, by using an incomplete measurement set consisting of only low spatial frequency components and certain constraints (such as sparsity and continuity), the high spatial frequencies and therefore sub-diffraction features can be numerically reconstructed during post-processing. Super-resolution up to five times higher than the diffraction limit has been demonstrated in proof-of-principle experiments.37,3941 Combined with structured illumination microscopy, compressive sensing helped to achieve 60-nm resolution in live-cell imaging.42 The spatial resolution depends on various factors, such as the measurement matrix size, the level of noise and stability, and the choice of an algorithm. A critical constraint is the sparsity of a sample. While the most natural images may not appear sparse, they still have a sparse representation, implying sparsity in a certain basis, e.g., after a wavelet transform. Therefore, identifying a proper basis and/or suitable algorithm is key for imaging biological tissues and samples that are not sparse in the regular domain.43

Compressive sensing protocols require the sampling to fulfill specific conditions to gather enough information from all parts of the object. Therefore, practical design guidelines depend on the optical system and the sample. One of the easiest and most general implementation with good reconstruction guarantee is based on randomized illumination since the random matrix is highly incoherent with any analytically fixed basis.44 By using an MMF, we can create the desired random illumination “for free,” as shown in Fig. 1(b) as one of the main properties of an MMF is to randomly scramble light without losing power. It makes an MMF a unique hair-thin instrument for computational imaging in vivo deep inside the living tissues, as shown in Fig. 2(a).

Fig. 2

(a) Illustration of an exciting application area of computational imaging through an MMF: minimally invasive fast and super-resolution deep tissue 3D imaging in freely moving animals. (b)–(d) Three main steps of computational imaging through an MMF: (a) pre-calibration that includes recording of random patterns without the sample, (b) illuminating the sample and measuring of the total signal for each illumination pattern, and (d) computational image reconstruction using iterative algorithms of machine learning frameworks.

NPh_11_S1_S11509_f002.png

4.

Compressive Fiber Imaging: State-of-the-Art and Outlook

Computational imaging via an MMF can be realized by illuminating the sample with a set of random patterns generated by, e.g., raster scanning of the input fiber facet with a focused spot.45,46 Imaging procedure consists of three main steps as shown in Figs. 2(b)2(d). During the pre-calibration, intensity distributions of a large set of random illumination patterns are recorded, as shown in Fig. 2(b). Every image is flattened and all are assembled to two-dimensional matrix A. It is the most time-consuming experimental step, as it requires recording hundreds of images and may take from seconds to a minute depending on a camera frame rate. Fortunately, the pre-calibration is needed only once. Then, an MMF probe is inserted to the region of interest, as shown in Fig. 2(c). The sample is illuminated by the same set of patterns and the total signal (e.g., fluorescent response) for each pattern is recorded (vector, y). Compressive sensing facilitates a significant reduction of acquisition time: the lengths of y is much smaller than the lengths of flattened sample x to be reconstructed.46 Moreover, a camera is not necessary anymore, allowing the use of much faster detectors. The measurement rate is limited by sensitivity, field of view, and scanning speed. Utilizing a DMD (22 kHz) can provide video rate imaging at over 25 fps for up to 880 speckle patterns, well-suited for a conventional 50  μm MMF. Finally, the flattened image of a sample (x) can be reconstructed by computational algorithms that essentially perform the pseudo-inversion of the (under-determined) linear system: Ax=y, as shown in Fig. 2(d). The reconstruction speed typically ranges from seconds to minutes, depending on the number of pixels and computational power.

Computational compressive imaging, in contrast to many other super-resolution approaches, is not only integrable with an ultrathin probe, as shown in Fig. 1(d)27,47 but also does not require special fluorescent marks and can be used to increase resolution more than twofold beyond the diffraction limit label-free.48 It opens up ways to create flexible probes that do not require recalibration or access to the distal end of the fiber during imaging. Flexible probes have been demonstrated based on GRIN49,50 or step-index MMFs with low51 and high spatial resolution.52,53 Wavelength-dependent scattering is utilized to create a flexible probe based on a single-mode fiber.54 Compressive sensing improves imaging through multicore fibers by reducing acquisition time, preventing photo-bleaching, and increasing space-bandwidth product.55,56 Compression is also beneficial for conventional raster-scan imaging modalities helping to improve quality by harnessing “muddy” modes57 and to speed up pre-calibration measurements.58

MMFs are already used as minimally invasive probes for neuroscience and clinical applications. Wavefront shaping through an MMF probe have allowed for minimally invasive in vivo imaging of neurons in deep-brain regions5962 and gastrointestinal imaging.63 Computational post-processing for a non-imaging MMF probe was used for tracking the activity of neurons.64 Compressive imaging through a single MMF visualized accumulation of lipofuscin in Alzheimer’s disease human brain with sub-Nyquist speed.65 Although many recent publications have shown the feasibility and potential of MMF imaging to become a key technology for deep-tissue brain imaging in freely moving animals, there are many challenges and associated opportunities for advancing the field. Currently, conventional diffraction limited imaging with raster scanned foci still offers better resilience to noise in a low-photon regime.66 Future research directions include the development of fast, precise, and robust to noise algorithms and machine learning frameworks.67,68

Fast high-resolution imaging of a large field of view in 3D is highly demanded.69 However, the transition to 3D presents several major challenges.70,71 The signal falls off rapidly with a distance and dense labeling may hinder deeper layers complicating the reconstruction workflow. Exploring hybrid imaging approaches, e.g., using blinking molecules, is a promising direction. Neuroscience applications require monitoring of multiple markers, therefore we need to visualize various contrast mechanisms simultaneously including quantitative label-free phase imaging.72 The next great challenge is to integrate different microscopy techniques into a single ultra-thin MMF probe for parallel multifunctional imaging. Finally, real-life video-rate brain imaging in freely moving animals through an ultimately thin fiber probe to be demonstrated.

To summarize, in many aspects, this new computational MMF imaging paradigm has already exceeded the current state-of-the-art. Ongoing progress in experimental design, fiber probes, and algorithms is rapidly improving both the performance and applicability. We will soon witness the computational MMF brain imaging to reach the level of readiness for technology to be transferred in the domain of in vivo neuroscience.

Disclosures

No conflicts of interest to disclose.

Code and Data Availability

Data sharing is not applicable to this article, as no new data were created.

Acknowledgments

I thank all the members of my research group at the Advanced Research Center for Nanolithography (ARCNL) and VU Amsterdam for their support and fruitful discussions.

References

1. 

V. Ntziachristos, “Going deeper than microscopy: the optical imaging frontier in biology,” Nat. Methods, 7 (8), 603 –614 https://doi.org/10.1038/nmeth.1483 1548-7091 (2010). Google Scholar

2. 

N. Ji, “The practical and fundamental limits of optical imaging in mammalian brains,” Neuron, 83 (6), 1242 –1245 https://doi.org/10.1016/j.neuron.2014.08.009 NERNET 0896-6273 (2014). Google Scholar

3. 

M. Minderer et al., “Virtual reality explored,” Nature, 533 (7603), 324 –325 https://doi.org/10.1038/nature17899 (2016). Google Scholar

4. 

S. Chen et al., “Miniature fluorescence microscopy for imaging brain activity in freely-behaving animals,” Neurosci. Bull., 36 (10), 1182 –1190 https://doi.org/10.1007/s12264-020-00561-z NRPBA2 0028-3967 (2020). Google Scholar

5. 

W. Zong et al., “Large-scale two-photon calcium imaging in freely moving mice,” Cell, 185 (7), 1240 –1256.e30 https://doi.org/10.1016/j.cell.2022.02.017 CELLB5 0092-8674 (2022). Google Scholar

6. 

N. G. Horton et al., “In vivo three-photon microscopy of subcortical structures within an intact mouse brain,” Nat. Photonics, 7 (3), 205 –209 https://doi.org/10.1038/nphoton.2012.336 NPAHBY 1749-4885 (2013). Google Scholar

7. 

S. Gigan, “Optical microscopy aims deep,” Nat. Photonics, 11 (1), 14 –16 https://doi.org/10.1038/nphoton.2016.257 NPAHBY 1749-4885 (2017). Google Scholar

8. 

L. Zhang et al., “Miniscope grin lens system for calcium imaging of neuronal activity from deep brain structures in behaving animals,” Curr. Protoc. Neurosci., 86 (1), e56 https://doi.org/10.1002/cpns.56 (2019). Google Scholar

9. 

H. H. Hopkins and N. S. Kapany, “A flexible fibrescope, using static scanning,” Nature, 173 (4392), 39 –41 https://doi.org/10.1038/173039b0 (1954). Google Scholar

10. 

W. Choi et al., “Flexible-type ultrathin holographic endoscope for microscopic imaging of unstained biological tissues,” Nat. Commun., 13 (1), 4469 https://doi.org/10.1038/s41467-022-32114-5 NCAOBW 2041-1723 (2022). Google Scholar

11. 

E. R. Andresen et al., “Toward endoscopes with no distal optics: video-rate scanning microscopy through a fiber bundle,” Opt. Lett., 38 (5), 609 –611 https://doi.org/10.1364/OL.38.000609 OPLEDP 0146-9592 (2013). Google Scholar

12. 

L. V. Doronina-Amitonova et al., “High-resolution wide-field Raman imaging through a fiber bundle,” Appl. Phys. Lett., 102 (16), 161113 https://doi.org/10.1063/1.4801847 APPLAB 0003-6951 (2013). Google Scholar

13. 

H. Cao et al., “Controlling light propagation in multimode fibers for imaging, spectroscopy, and beyond,” Adv. Opt. Photon., 15 524 –612 https://doi.org/10.1364/AOP.484298 (2023). Google Scholar

14. 

A. W. Snyder and J. D. Love, Optical Waveguide Theory, Chapman and Hall, London; New York (1983). Google Scholar

15. 

M. Plöschner, T. Tyc and T. Čižmár, “Seeing through chaos in multimode fibres,” Nat. Photonics, 9 (8), 529 –535 https://doi.org/10.1038/nphoton.2015.112 NPAHBY 1749-4885 (2015). Google Scholar

16. 

I. M. Vellekoop and A. P. Mosk, “Focusing coherent light through opaque strongly scattering media,” Opt. Lett., 32 (16), 2309 –2311 https://doi.org/10.1364/OL.32.002309 OPLEDP 0146-9592 (2007). Google Scholar

17. 

I. M. Vellekoop, A. Lagendijk and A. P. Mosk, “Exploiting disorder for perfect focusing,” Nat. Photonics, 4 (5), 320 –322 https://doi.org/10.1038/nphoton.2010.3 NPAHBY 1749-4885 (2010). Google Scholar

18. 

M. Mounaix et al., “Time reversed optical waves by arbitrary vector spatiotemporal field generation,” Nat. Commun., 11 (1), 5813 https://doi.org/10.1038/s41467-020-19601-3 NCAOBW 2041-1723 (2020). Google Scholar

19. 

T. Čižmár and K. Dholakia, “Shaping the light transmission through a multimode optical fibre: complex transformation analysis and applications in biophotonics,” Opt. Express, 19 (20), 18871 –18884 https://doi.org/10.1364/OE.19.018871 OPEXFF 1094-4087 (2011). Google Scholar

20. 

R. Di Leonardo and S. Bianchi, “Hologram transmission through multi-mode optical fibers,” Opt. Express, 19 (1), 247 –254 https://doi.org/10.1364/OE.19.000247 OPEXFF 1094-4087 (2011). Google Scholar

21. 

T. Čižmár and K. Dholakia, “Exploiting multimode waveguides for pure fibre-based imaging,” Nat. Commun., 3 1027 https://doi.org/10.1038/ncomms2024 NCAOBW 2041-1723 (2012). Google Scholar

22. 

A. Descloux, L. V. Amitonova and P. W. Pinkse, “Aberrations of the point spread function of a multimode fiber due to partial mode excitation,” Opt. Express, 24 (16), 18501 –18512 https://doi.org/10.1364/OE.24.018501 OPEXFF 1094-4087 (2016). Google Scholar

23. 

R. Turcotte et al., “Focusing light in biological tissue through a multimode optical fiber: refractive index matching,” Opt. Lett., 44 (10), 2386 –2389 https://doi.org/10.1364/OL.44.002386 OPLEDP 0146-9592 (2019). Google Scholar

24. 

D. Loterie et al., “Digital confocal microscopy through a multimode fiber,” Opt. Express, 23 (18), 23845 https://doi.org/10.1364/OE.23.023845 OPEXFF 1094-4087 (2015). Google Scholar

25. 

L. V. Amitonova et al., “High-resolution wavefront shaping with a photonic crystal fiber for multimode fiber imaging,” Opt. Lett., 41 (3), 497 https://doi.org/10.1364/OL.41.000497 OPLEDP 0146-9592 (2016). Google Scholar

26. 

I. T. Leite et al., “Three-dimensional holographic optical manipulation through a high-numerical-aperture soft-glass multimode fibre,” Nat. Photonics, 12 (1), 33 –39 https://doi.org/10.1038/s41566-017-0053-8 NPAHBY 1749-4885 (2018). Google Scholar

27. 

L. V. Amitonova and J. F. de Boer, “Endo-microscopy beyond the Abbe and Nyquist limits,” Light Sci. Appl., 9 (1), 81 https://doi.org/10.1038/s41377-020-0308-x (2020). Google Scholar

28. 

R. Y. Gu, R. N. Mahalati and J. M. Kahn, “Design of flexible multi-mode fiber endoscope,” Opt. Express, 23 (21), 26905 https://doi.org/10.1364/OE.23.026905 OPEXFF 1094-4087 (2015). Google Scholar

29. 

A. M. Caravaca-Aguirre et al., “Real-time resilient focusing through a bending multimode fiber,” Opt. Express, 21 (10), 12881 https://doi.org/10.1364/OE.21.012881 OPEXFF 1094-4087 (2013). Google Scholar

30. 

M. W. Matthès et al., “Learning and avoiding disorder in multimode fibers,” Phys. Rev. X, 11 (2), 021060 https://doi.org/10.1103/PhysRevX.11.021060 PRXHAE 2160-3308 (2021). Google Scholar

31. 

P. Ambichl et al., “Super- and anti-principal-modes in multimode waveguides,” Phys. Rev. X, 7 (4), 041053 https://doi.org/10.1103/PhysRevX.7.041053 PRXHAE 2160-3308 (2017). Google Scholar

32. 

W. Xiong, C. W. Hsu and H. Cao, “Long-range spatio-temporal correlations in multimode fibers for pulse delivery,” Nat. Commun., 10 (1), 2973 https://doi.org/10.1038/s41467-019-10916-4 NCAOBW 2041-1723 (2019). Google Scholar

33. 

J. N. Mait, G. W. Euliss and R. A. Athale, “Computational imaging,” Adv. Opt. Photonics, 10 (2), 409 https://doi.org/10.1364/AOP.10.000409 AOPAC7 1943-8206 (2018). Google Scholar

34. 

E. J. Candès, J. K. Romberg and T. Tao, “Stable signal recovery from incomplete and inaccurate measurements,” Commun. Pure Appl. Math., 59 (8), 1207 –1223 https://doi.org/10.1002/cpa.20124 CPMAMV 0010-3640 (2006). Google Scholar

35. 

O. Katz, Y. Bromberg and Y. Silberberg, “Compressive ghost imaging,” Appl. Phys. Lett., 95 (13), 131110 https://doi.org/10.1063/1.3238296 APPLAB 0003-6951 (2009). Google Scholar

36. 

D. Donoho, “Superresolution via sparsity constraints,” SIAM J. Math. Anal., 23 (5), 1309 –1331 https://doi.org/10.1137/0523074 SJMAAH 0036-1410 (1992). Google Scholar

37. 

S. Gazit et al., “Super-resolution and reconstruction of sparse sub-wavelength images,” Opt. Express, 17 (26), 23920 –23946 https://doi.org/10.1364/OE.17.023920 OPEXFF 1094-4087 (2009). Google Scholar

38. 

J. W. Goodman, Introduction to Fourier Optics, Roberts and Company Publishers( (2005). Google Scholar

39. 

M. Pascucci et al., “Compressive three-dimensional super-resolution microscopy with speckle-saturated fluorescence excitation,” Nat. Commun., 10 (1), 1327 https://doi.org/10.1038/s41467-019-09297-5 NCAOBW 2041-1723 (2019). Google Scholar

40. 

W. Li et al., “Single-frame wide-field nanoscopy based on ghost imaging via sparsity constraints,” Optica, 6 (12), 1515 –1523 https://doi.org/10.1364/OPTICA.6.001515 (2019). Google Scholar

41. 

W. Gong and S. Han, “High-resolution far-field ghost imaging via sparsity constraint,” Sci. Rep., 5 9280 https://doi.org/10.1038/srep09280 SRCEC3 2045-2322 (2015). Google Scholar

42. 

W. Zhao et al., “Sparse deconvolution improves the resolution of live-cell super-resolution fluorescence microscopy,” Nat. Biotechnol., 60 606 –617 https://doi.org/10.1038/s41587-021-01092-2 NABIF9 1087-0156 (2021). Google Scholar

43. 

A. Liutkus et al., “Imaging with nature: compressive imaging using a multiply scattering medium,” Sci. Rep., 4 5552 https://doi.org/10.1038/srep05552 SRCEC3 2045-2322 (2014). Google Scholar

44. 

D. L. Donoho, “Compressed sensing,” IEEE Trans. Inf. Theory, 52 (4), 1289 –1306 https://doi.org/10.1109/TIT.2006.871582 IETTAW 0018-9448 (2006). Google Scholar

45. 

R. N. Mahalati, R. Y. Gu and J. M. Kahn, “Resolution limits for imaging through multi-mode fiber,” Opt. Express, 21 (2), 1656 –1668 https://doi.org/10.1364/OE.21.001656 OPEXFF 1094-4087 (2013). Google Scholar

46. 

L. V. Amitonova and J. F. de Boer, “Compressive imaging through a multimode fiber,” Opt. Lett., 43 (21), 5427 https://doi.org/10.1364/OL.43.005427 OPLEDP 0146-9592 (2018). Google Scholar

47. 

B. Lochocki et al., “Ultimate resolution limits of speckle-based compressive imaging,” Opt. Express, 29 (3), 3943 –3955 https://doi.org/10.1364/OE.413831 OPEXFF 1094-4087 (2021). Google Scholar

48. 

K. Abrashitova and L. V. Amitonova, “High-speed label-free multimode-fiber-based compressive imaging beyond the diffraction limit,” Opt. Express, 30 (7), 10456 –10469 https://doi.org/10.1364/OE.444796 OPEXFF 1094-4087 (2022). Google Scholar

49. 

A. Abdulaziz et al., “Robust real-time imaging through flexible multimode fibers,” Sci. Rep., 13 (1), 11371 https://doi.org/10.1038/s41598-023-38480-4 SRCEC3 2045-2322 (2023). Google Scholar

50. 

D. E. Boonzajer Flaes et al., “Robustness of light-transport processes to bending deformations in graded-index multimode waveguides,” Phys. Rev. Lett., 120 (23), 233901 https://doi.org/10.1103/PhysRevLett.120.233901 PRLTAO 0031-9007 (2018). Google Scholar

51. 

D. Choudhury et al., “Computational optical imaging with a photonic lantern,” Nat. Commun., 11 (1), 5217 https://doi.org/10.1038/s41467-020-18818-6 NCAOBW 2041-1723 (2020). Google Scholar

52. 

Z. Lyu et al., “Sub-diffraction computational imaging via a flexible multicore-multimode fiber,” Opt. Express, 31 (7), 11249 –11260 https://doi.org/10.1364/OE.481052 OPEXFF 1094-4087 (2023). Google Scholar

53. 

B. Lochocki et al., “Swept-source multimode fiber imaging,” Sci. Rep., 13 (1), 8071 https://doi.org/10.1038/s41598-023-34062-6 SRCEC3 2045-2322 (2023). Google Scholar

54. 

J. Shin, B. T. Bosworth and M. A. Foster, “Single-pixel imaging using compressed sensing and wavelength-dependent scattering,” Opt. Lett., 41 (5), 886 https://doi.org/10.1364/OL.41.000886 OPLEDP 0146-9592 (2016). Google Scholar

55. 

S. Guérit et al., “Compressive imaging through optical fiber with partial speckle scanning,” SIAM J. Imaging Sci., 15 (2), 387 –423 https://doi.org/10.1137/21M1407586 (2022). Google Scholar

56. 

J. Shin et al., “A minimally invasive lens-free computational microendoscope,” Sci. Adv., 5 (12), eaaw5595 https://doi.org/10.1126/sciadv.aaw5595 STAMCV 1468-6996 (2019). Google Scholar

57. 

T. Tučková et al., “Computational image enhancement of multimode fibre-based holographic endo-microscopy: harnessing the muddy modes,” Opt. Express, 29 (23), 38206 –38220 https://doi.org/10.1364/OE.434848 OPEXFF 1094-4087 (2021). Google Scholar

58. 

S. Li et al., “Compressively sampling the optical transmission matrix of a multimode fibre,” Light Sci. Appl., 10 (1), 88 https://doi.org/10.1038/s41377-021-00514-9 (2021). Google Scholar

59. 

S. Ohayon et al., “Minimally invasive multimode optical fiber microendoscope for deep brain fluorescence imaging,” Biomed. Opt. Express, 9 (4), 1492 –1509 https://doi.org/10.1364/BOE.9.001492 BOEICL 2156-7085 (2018). Google Scholar

60. 

S. A. Vasquez-Lopez et al., “Subcellular spatial resolution achieved for deep-brain imaging in vivo using a minimally invasive multimode fiber,” Light Sci. Appl., 7 (1), 110 https://doi.org/10.1038/s41377-018-0111-0 (2018). Google Scholar

61. 

R. Turcotte et al., “Volumetric two-photon fluorescence imaging of live neurons using a multimode optical fiber,” Opt. Lett., 45 (24), 6599 –6602 https://doi.org/10.1364/OL.409464 OPLEDP 0146-9592 (2020). Google Scholar

62. 

M. Stibůrek 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

63. 

Z. Wen et al., “Single multimode fibre for in vivo light-field-encoded endoscopic imaging,” Nat. Photonics, 17 679 –687 https://doi.org/10.1038/s41566-023-01240-x NPAHBY 1749-4885 (2023). Google Scholar

64. 

C. V. Rimoli et al., “Demixing fluorescence time traces transmitted by multimode fibers,” (2024). Google Scholar

65. 

B. Lochocki et al., “Epi-fluorescence imaging of the human brain through a multimode fiber,” APL Photonics, 7 (7), 071301 https://doi.org/10.1063/5.0080672 (2022). Google Scholar

66. 

R. Y. Gu, R. N. Mahalati and J. M. Kahn, “Noise-reduction algorithms for optimization-based imaging through multi-mode fiber,” Opt. Express, 22 (12), 15118 –15132 https://doi.org/10.1364/OE.22.015118 OPEXFF 1094-4087 (2014). Google Scholar

67. 

W. Li et al., “Generative adversarial network for superresolution imaging through a fiber,” Phys. Rev. Appl., 18 (3), 034075 https://doi.org/10.1103/PhysRevApplied.18.034075 PRAHB2 2331-7019 (2022). Google Scholar

68. 

W. Li, K. Abrashitova and L. V. Amitonova, “Super-resolution multimode fiber imaging with an untrained neural network,” Opt. Lett., 48 (13), 3363 –3366 https://doi.org/10.1364/OL.491375 OPLEDP 0146-9592 (2023). Google Scholar

69. 

K. Yanny et al., “Miniscope3D: optimized single-shot miniature 3D fluorescence microscopy,” Light Sci. Appl., 9 (1), 171 https://doi.org/10.1038/s41377-020-00403-7 (2020). Google Scholar

70. 

S.-Y. Lee et al., “Confocal 3D reflectance imaging through multimode fiber without wavefront shaping,” Optica, 9 (1), 112 –120 https://doi.org/10.1364/OPTICA.446178 (2022). Google Scholar

71. 

D. Stellinga et al., “Time-of-flight 3D imaging through multimode optical fibers,” Science, 374 (6573), 1395 –1399 https://doi.org/10.1126/science.abl3771 SCIEAS 0036-8075 (2021). Google Scholar

72. 

J. Sun et al., “Quantitative phase imaging through an ultra-thin lensless fiber endoscope,” Light Sci. Appl., 11 (1), 204 https://doi.org/10.1038/s41377-022-00898-2 (2022). Google Scholar

Biography

Lyubov V. Amitonova is an assistant professor at Vrije University Amsterdam and a group leader at the Advanced Research Center for Nanolithography, the Netherlands. She received her PhD from the Lomonosov Moscow State University (Russia) in 2013. Her research interests include computational optical imaging, fiber endoscopy, and advanced wavefront shaping.

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.
Liubov V. Amitonova "Multimode fiber endoscopes for computational brain imaging," Neurophotonics 11(S1), S11509 (6 March 2024). https://doi.org/10.1117/1.NPh.11.S1.S11509
Received: 1 September 2023; Accepted: 31 January 2024; Published: 6 March 2024
Advertisement
Advertisement
KEYWORDS
Multimode fibers

Brain

Brain imaging

Neuroimaging

Endoscopes

Animals

Compressed sensing

Back to Top