All-Optical Methods To Study Neuronal Function
All-Optical Methods To Study Neuronal Function
All-Optical
Methods to
Study Neuronal
Function
NEUROMETHODS
Series Editor
Wolfgang Walz
University of Saskatchewan
Saskatoon, SK, Canada
Edited by
Eirini Papagiakoumou
Institut de la Vision, Sorbonne Université, INSERM, CNRS, Paris, France
Editor
Eirini Papagiakoumou
Institut de la Vision
Sorbonne Université, INSERM, CNRS
Paris, France
This Humana imprint is published by the registered company Springer Science+Business Media, LLC, part of Springer
Nature.
The registered company address is: 1 New York Plaza, New York, NY 10004, U.S.A.
Preface to the Series
Experimental life sciences have two basic foundations: concepts and tools. The Neuro-
methods series focuses on the tools and techniques unique to the investigation of the
nervous system and excitable cells. It will not, however, shortchange the concept side of
things as care has been taken to integrate these tools within the context of the concepts and
questions under investigation. In this way, the series is unique in that it not only collects
protocols but also includes theoretical background information and critiques which led to
the methods and their development. Thus, it gives the reader a better understanding of the
origin of the techniques and their potential future development. The Neuromethods
publishing program strikes a balance between recent and exciting developments like those
concerning new animal models of disease, imaging, in vivo methods, and more established
techniques, including, for example, immunocytochemistry and electrophysiological tech-
nologies. New trainees in neurosciences still need a sound footing in these older methods in
order to apply a critical approach to their results.
Under the guidance of its founders, Alan Boulton and Glen Baker, the Neuromethods
series has been a success since its first volume published through Humana Press in 1985. The
series continues to flourish through many changes over the years. It is now published under
the umbrella of Springer Protocols. While methods involving brain research have changed a
lot since the series started, the publishing environment and technology have changed even
more radically. Neuromethods has the distinct layout and style of the Springer Protocols
program, designed specifically for readability and ease of reference in a laboratory setting.
The careful application of methods is potentially the most important step in the process
of scientific inquiry. In the past, new methodologies led the way in developing new dis-
ciplines in the biological and medical sciences. For example, physiology emerged out of
anatomy in the nineteenth century by harnessing new methods based on the newly discov-
ered phenomenon of electricity. Nowadays, the relationships between disciplines and meth-
ods are more complex. Methods are now widely shared between disciplines and research
areas. New developments in electronic publishing make it possible for scientists that
encounter new methods to quickly find sources of information electronically. The design
of individual volumes and chapters in this series takes this new access technology into
account. Springer Protocols makes it possible to download single protocols separately. In
addition, Springer makes its print-on-demand technology available globally. A print copy
can therefore be acquired quickly and for a competitive price anywhere in the world.
v
Preface
Control and monitoring of neuronal activity with light, what is often called all-optical
manipulation of neurons, is admittedly the most adequate method for addressing questions
regarding communication of neurons in a neural circuit or between different circuits. The
big and continuously expanding toolbox of molecular photosensitive probes that activate/
inhibit or image (through membrane voltage or calcium changes) neuronal activity, in
combination with the development of original light-microscopy methods for stimulating
these probes, has tremendously contributed to this direction and led to innovative experi-
mental concepts, where neurons can be manipulated either as entities or as ensembles.
Indeed, the use of light offers suitable spatiotemporal resolution to manipulate neurons at
single-cell specificity. At the same time, it gives access to a large population of cells simulta-
neously via scanless, parallel illumination methods.
Neural circuit studies are more conclusive for addressing biological questions when
performed in vivo. In this sense, they necessitate three-dimensional (3D) accessibility both
for activation and imaging, at physiological time scales (few-ms scale activation and imag-
ing). 3D imaging approaches enable today using complementary strategies to access
volumes extending up to hundreds of micrometers in the axial direction. On the contrary,
the development of 3D photoactivation methods is more recent. These systems use
computer-generated holography (CGH), a technique based on phase modulation of the
excitation beam’s wavefront, to create multiple excitation regions of interest. Thanks to
3D-CGH, used either solely (parallel methods) or in its diffraction-limit version in combi-
nation with scanning of the holographic beamlets, it is nowadays possible to simultaneously
activate multiple neurons providing both the adequate temporal resolution, as well as the
spatial resolution for near single-cell precision. High spatial resolution and selectivity is often
assured by implementing those methods with two-photon excitation.
Although the first experiments of all-optical manipulation of neuronal activity per-
formed activation of neurons via uncaging of caged glutamate, the term all-optical today
is mostly related to the combination of functional imaging and optogenetic activation. There
is a growing number of studies using optogenetics and calcium imaging to explore several
hypotheses in cellular and systems neuroscience, nevertheless a full optical neuronal control
remains a challenge in terms of achieving reliable delivery and expression of sensors and
actuators in the same neurons, eliminating the crosstalk between imaging and activation,
and recording and stimulating with single-neuron and single-action-potential precision.
In this volume, we opt to give an overview of the methods that have been used so far in
all-optical experiments, but also to present other promising approaches potentially useful in
this domain. The book is addressed to people experienced in different disciplines, such as
physicists, engineers, and neuroscientists; therefore, it starts by providing some basic but
fundamental background information in terms of both physiology and optics in the context
of all-optical two-photon neurophysiology experiments (Chap. 1), followed by some
prompts for the selection of appropriate actuators and sensors, and functional imaging
methodologies that drive the choice of both, together with the suitable laser sources for
two-photon excitation (Chaps. 1 and 2).
vii
viii Preface
We then present, in detail, optical methods that have been used for photoactivation and
imaging. The reader can find the design principles and, in some cases, hardware implemen-
tation for methods like generalized phase contrast (Chap. 1), computer-generated hologra-
phy and scanning approaches (Chaps. 3 and 4), temporal focusing (Chaps. 1 and 4), as well
as guidance to the entire workflow for an all-optical experiment in circuit neuroscience
(Chap. 5). In Chap. 6, possibilities and limitations of optogenetic actuators are discussed
within the context of an all-optical single-beam experiment by giving insights into the
photophysical properties of actuators. Detailed methods are provided in Chap. 7 on a
miniature head mounted two-photon fiber-coupled microscope for imaging neuronal activ-
ity in vivo in freely moving animals, while Chap. 8 discusses how 3D holographic optoge-
netics can be added to a home-built light sheet microscope.
This book also attributes a part on innovative imaging techniques that could be
implemented in the framework of an all-optical electrophysiology experiment. Chapter 9
pronounces the theories of temporal focusing in combination with single-pixel detection for
imaging of fast collective biological processes at depth, over a widefield and at high
spatiotemporal resolution. Chapter 10 presents alternative imaging methods at synaptic
resolution, such as two-photon fluorescence microscopy equipped with Bessel focus scan-
ning technology and widefield fluorescence microscopy with optical sectioning ability.
The implementation of simultaneous two-photon imaging and holographic optoge-
netics in conjunction with population analytical tools or with psychophysical measurements
of evoked synthetic percepts to confirm a precise relationship between optical manipulations
and behavior is presented in Chaps. 11 and 12.
Finally, in Chap. 13, an approach for label-free imaging is presented: stimulated Raman
scattering (SRS) microscopy, a non-linear imaging method for visualizing a molecule based
on its chemical properties, and the way to integrate it in a commercial multiphoton
microscope, eventually useful for label-free functional imaging.
The use of all-optical methods for studying the neural function is a multiparametric and
arduous project to setup. It entails a multidisciplinary know-how both for developing the
optical system and the adequate biological preparation, and sometimes expensive equip-
ment, especially when multiphoton excitation is considered. We hope that this book can
serve as a guide to facilitate the first requirement, establishing a useful reference for groups
starting their activity in this domain, and give insights on the optical systems, the choice of
actuators and sensors, but also stimulate ideas for ground-breaking configurations and
experiments.
Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 417
Contributors
xi
xii Contributors
ALBRECHT STROH • Leibniz Institute for Resilience Research, Mainz, Germany; Institute of
Pathophysiology, University Medical Center Mainz, Mainz, Germany
DIMITRII TANESE • Wavefront-Engineering Microscopy Group, Photonics Department,
Institut de la Vision, Sorbonne Université, INSERM, CNRS, Paris, France
TAMÁS VÁCZI • Institute for Solid State Physics and Optics, Wigner Research Centre for
Physics, Budapest, Hungary
MIKLÓS VERES • Institute for Solid State Physics and Optics, Wigner Research Centre for
Physics, Budapest, Hungary
ANNA WIERCZEIKO • Leibniz Institute for Resilience Research, Mainz, Germany
PHILIP WIJESINGHE • SUPA, School of Physics and Astronomy, University of St Andrews,
Scotland, UK
WEIJIAN YANG • University of California at Davis, Davis, CA, USA
MICHAEL D. YOUNG • Department of Bioengineering, University of Colorado Denver,
Aurora, CO, USA
RAFAEL YUSTE • NeuroTechnology Center, Columbia University, New York, NY, USA
QINRONG ZHANG • Department of Physics, University of California, Berkeley, CA, USA
Chapter 1
Abstract
One of the central goals of neuroscience is to decipher the specific contributions of neural mechanisms to
different aspects of sensory perception. Since achieving this goal requires tools capable of precisely
perturbing and monitoring neural activity across a multitude of spatiotemporal scales, this aim has inspired
the innovation of many optical technologies capable of manipulating and recording neural activity in a
minimally invasive manner. The interdisciplinary nature of neurophotonics requires a broad knowledge base
in order to successfully develop and apply these technologies, and one of the principal aims of this chapter is
to provide some basic but fundamental background information in terms of both physiology and optics in
the context of all-optical two-photon neurophysiology experiments. Most of this information is expected to
be familiar to readers experienced in either domain, but is presented here with the aim of bridging the divide
between disciplines in order to enable physicists and engineers to develop useful optical technologies or for
neuroscientists to select appropriate tools and apply them to their maximum potential.
The first section of this chapter is dedicated to a brief overview of some basic principles of neural
physiology relevant for controlling and recording neuronal activity using light. Then, the selection of
appropriate actuators and sensors for manipulating and monitoring particular neural signals is discussed,
with particular attention paid to kinetics and sensitivity. Some considerations for minimizing crosstalk in
optical neurophysiology experiments are also introduced. Next, an overview of the state-of-the-art optical
technologies is provided, including a description of suitable laser sources for two-photon excitation
according to particular experimental requirements. Finally, some detailed, technical, information regarding
the specific wavefront engineering approaches known as Generalized Phase Contrast (GPC) and temporal
focusing is provided.
Key words All-optical neurophysiology, Light shaping, Temporal focusing, Generalized phase con-
trast, Computer-generated holography, Functional imaging, Optogenetics, Molecular tools, GECIs,
GEVIs
1 Introduction
Eirini Papagiakoumou (ed.), All-Optical Methods to Study Neuronal Function, Neuromethods, vol. 191,
https://doi.org/10.1007/978-1-0716-2764-8_1, © The Author(s) 2023
1
2 Ruth R. Sims et al.
Fig. 2 Two-photon characterization of channelrhodopsins. (a) In the simplest conceptual model of the opsin
photocycle, ion channels reside in the closed state. Upon light absorption, the channels open, allowing the
exchange of ions between the cytosol and extracellular space. Depending on the ion selectivity of the channel
and the electrochemical gradient, this flow of ions will hyperpolarize or depolarize the cell membrane and
ultimately inhibit or excite the cell. (b) In reality, the opsin photocycle is more complex, but can reasonably be
approximated by the so-called four-state model. For more details refer to [34, 35]. (c) Opsins can be
characterized using whole-cell voltage patch clamp to measure the currents that flow across the cell
membrane under different conditions. Inset: visualization of a characteristic 12-μm diameter holographic
spot, typically used for parallel 2P- photoactivation. (d) Photocurrent traces recorded in whole cell voltage
patch clamp from CHO (Chinese Hamster Ovary) cells expressing ChRmine as a function of increasing 2P
excitation power (920 nm, 12-μm diameter excitation spot, 200 ms continuous illumination, incident powers
varied between 0 and 50 mW as indicated in the color bar). The characteristic features of the photocurrent
traces (kinetics and peak/stationary photocurrent) are labeled. The magnitude of the photocurrent increases
with power density to saturation. (e) 2P-LSM image of AAV9-CaMKIIa-somBiPOLES-mCerulean expressed in
hippocampal organotypic slice cultures by bulk infection (scale bar represents 50 μm). (f) Photostimulation
(upper, 1100-nm illumination, 0.44 mW/μm2, 5 ms continuous illumination (red bar)) and inhibition (lower,
920-nm illumination, 0.3 mW/μm2, 200 ms illumination during constant current injection (gray bar)) of a single
neuron expressing somBiPOLES with a 12-μm diameter holographic spot
Fig. 3 Calcium and voltage indicators as reporters of neuronal activity. (a) Cytosolic calcium concentration
increases temporarily as a result of the change in membrane potential that occurs during an action potential.
The intensity-based fluorescent probe GCaMP binds to calcium. This alters the conformation of the circularly
permuted GFP chromophore and results in an increase in fluorescence intensity. (b) Left: 2P-LSM image of
AAV9-Syn-jGCaMP7s expressed in hippocampal organotypic slice cultures by bulk viral infection (scale bar
represents 25 μm). Right: fluorescent jGCaMP7s traces in response to trains of action potentials (5, 15, 40 Hz)
evoked by pulsed current injection into a patched neuron (indicated below). (c) In the case of voltage-sensing
domain (VSD)-based voltage indicators, a change in membrane potential causes a change in conformation of
the VSD, which is covalently linked to a circularly permuted fluorophore. The change in conformation of the
fluorophore typically results in a decrease in fluorescence intensity. (d) Left: 2P-LSM image of AAV8-hSyn-
ASAP3b expressed in hippocampal organotypic slice cultures by bulk viral infection (scale bar represents
25 μm). Right: simulated ASAP3b traces in response to trains of action potentials (5, 15, 40 Hz) as in (b)
Fig. 4 All-optical electrophysiology in mice hippocampal organotypic slices. (a) Two-photon fluorescence
image showing the co-expression of the high performance and soma-targeted cation channelrhodopsin
ST-ChroME, here tagged to the chromophore mRuby3 (red colour corresponds to the nuclear localization of
mRuby3 reporter) and the genetically encoded Ca2+ indicator GCaMP7s (green) in the CA3 region of a
hippocampal organotypic slice. White circles represent the two-photon temporally focused spots delivered
to excite 50 different neurons (12 μm spot diameter, 1040 nm wavelength, 0.26 mW/μm2 incident power).
Scale bar: 50 μm. (b) Two-photon imaging of GCaMP7s fluorescence signals evoked by the sequential
stimulation of the cells (interstimulus interval ~3 s). Gray bars represent the stimulation protocol which
consisted of a train of 5 pulses of 5-ms duration at 4 Hz. The identity of the cells during the sequential
stimulation is denoted by the blue numbers on top. In this experiment, 28 out of 50 cells yielded calcium
transients in response to stimulation (green horizontal arrowheads). During the acquisition time (~160 s) two
synchronous network-wide bursting events were observed (vertical arrowheads at the bottom), the first one
seemed to be triggered by the direct activation of a hub-like cell (cell number 15 in the sequence; see pink
inset), while the second network-wide event seemed to be triggered by the spontaneous activation of a
hub-like cell in the circuit. Pink and orange arrowheads denote the evoked or spontaneous nature of the
events, respectively. A single event (in only 1 neuron) with similar characteristics to the network-wide bursting
events in terms of amplitude and kinetics was observed near the end of the acquisition time (horizontal orange
arrowhead). The large amplitude of these events reflects the large number and/or frequency of action potential
firing in comparison to the fine-tuned control of firing activity evoked by single-cell resolution and
sub-millisecond precision patterned photostimulation as it is observed in the inset in (c)
3 Implementation of Methods
3.2 Beam Shaping As outlined in Sect. 2.3, many parallel two-photon excitation
with Generalized approaches rely on lateral beam sculpting. A correspondingly wide
Phase Contrast variety of methods based on amplitude or phase modulation have
been conceived of and demonstrated experimentally. Phase modu-
lation is generally preferable since it is more power efficient than
amplitude modulation. Computer-generated holography (CGH) is
currently the most common phase modulation method used for
photoactivation or imaging in all-optical neurophysiology experi-
ments. Since CGH is described in detail in other chapters of this
book (Chaps. 3, 4, and 11), this section will focus on the principles
and implementation of an alternative phase modulation approach:
generalized phase contrast (GPC) [187].
GPC is an efficient approach for transverse beam shaping and
has been applied to imaging [188, 189], photomanipulation [190–
192], and atom trapping [193]. GPC patterns have smooth,
speckle-free intensity profiles and can be combined with temporal
focusing for depth-resolved, robust excitation, deep in scattering
tissue [147, 194]. As demonstrated in Fig. 5a, in GPC, the phase
imprinted on a beam (using a phase mask or an SLM) is mapped to
intensity variations in a conjugate image plane by engineered con-
structive and destructive interference. The simplest implementa-
tions of GPC are based on 4f arrangements of lenses, constructed as
follows (Fig. 5a): the first phase modulating element (hereafter
SLM) is located a distance f1 prior to the first lens (L1), which has
focal length f1, and is referred to hereafter as the Fourier lens. The
necessary SLM phase (ϕxy(x,y)) depends on the spatial profile of the
desired pattern. For binary GPC, ϕxy = ϕ1 for SLM pixels inside the
pattern and ϕxy = ϕ2 for SLM pixels outside of the pattern, ϕ1 = π
and ϕ2 = 0 is a simple (and useful) choice. An element known as a
phase contrast filter (PCF) is located in the Fourier plane (FP) of
L1, and a distance f2 prior to the second lens (L2), which has focal
length f2. The PCF applies a selective phase shift to the field in the
Fourier plane. The phase shift imparted by the PCF depends on its
thickness (d) and refractive index of the substrate (n2): ϕPCF =
(2πd(n2-n1))/λ, where n1 is the refractive index of the medium
surrounding the PCF (usually n1 = 1 for air, and n2 = 1.45 for a
PCF fabricated with fused silica). For binary GPC, and ϕ1 = π, ϕ2 =
0, constructive interference in the output pattern occurs for ϕPCF =
π. The resulting interference pattern is formed in the image plane
(IP) of the second lens, a distance f2 from L2.
22 Ruth R. Sims et al.
Fig. 5 Wavefront engineering based on Generalized Phase Contrast. (a) (i) Schematic representation of a
common configuration for Generalized Phase Contrast. The beam is modulated using an SLM, which is used to
impart a phase shift to the portion of the beam corresponding to the desired pattern. The SLM phase should
match the desired pattern (up to a magnification factor according to the respective focal lengths of L1 and L2).
In the binary case, the SLM is usually used to impart a π phase shift to the pixels within the pattern and 0 to
those outside. The synthetic reference beam is the portion that is phase shifted by the phase contrast filter
(PCF), which typically imparts a π phase shift relative to the field that does not pass through the PCF, referred
to here as the modulated beam. The different portions of the beam are recombined by L2 in the Image Plane
(IP), where the modulated and synthetic reference fields interfere to form the desired pattern. (ii) Cartoon
representations of the ideal 2D amplitudes and phases of the electric fields in the input (SLM) plane and the
output (Image) plane. The phase profile of a typical PCF is shown centrally, with the filter diameter indicated by
dashed black lines. (iii) 1D cross sections of the amplitudes and phases of the electric fields in the case of
binary circle GPC. (b) 2-photon excited fluorescence from a thin rhodamine layer for two different patterns:
circle and ring GPC. Scale bars represent 10 μm
Fig. 6 Intuitive optimization of phase contrast filter for GPC. (a) The ideal reference field would generate total
constructive interference at all positions in the image plane within the desired pattern and total destructive
interference at all positions outside. The ideal reference field can be calculated by subtracting the modulated
field (i.e., the magnified image of the field at the SLM plane) from the field corresponding to the desired
pattern. The colors of the field profiles represent their phase ϕ, (blue: ϕ = 0 and red: ϕ = π). (b) The Fourier
transform (denoted F ) of this ideal reference field gives its profile in the Fourier plane, where the PCF is
located. The profiles of the ideal reference field and modulated field in the Fourier plane are used to guide the
All-Optical Neuronal Manipulation with Wavefront Engineering 25
Fig. 6 (continued) choice of an optimal PCF filter in GPC. The optimal PCF parameters are those for which the
synthetic reference field most closely matches the ideal reference field. It is clear that this occurs when the
PCF imparts a π phase shift and its edges coincide with the first zero crossings of the modulated field
(indicated by black dashed lines). (c) Since the synthetic reference field cannot completely match the ideal
reference field, there exist some differences between the ideal output field and that which is obtained. In most
cases, there is a mismatch between the beam waists of the synthetic reference and the modulated fields,
resulting in a “ring-of-light” surrounding the output pattern (highlighted by gray arrows). This is normally
blocked by an iris positioned in a conjugate image plane. Furthermore, since the synthetic reference field is
typically composed of the low spatial frequency components of the field, there are no small features and the
Gaussian profile of the input beam is not compensated for (highlighted by black arrows)
26 Ruth R. Sims et al.
3.3 Implementing Due to their interferometric character, GPC patterns suffer from a
Temporal Focusing lack of axial confinement and optical sectioning [147] (this is in
notable contrast to patterns generated using CGH). Temporal
focusing has been used to restore axial resolution for GPC, and
other extended light patterns that have been used for 2P
All-Optical Neuronal Manipulation with Wavefront Engineering 27
Fig. 7 Implementation of temporal focusing with light-shaping methods. (a) Temporal focusing of a Gaussian
beam. A diffraction grating placed in a conjugate image plane of the optical path is used to separate the
spectral frequencies (“colors”) of the femtosecond laser pulses. The grating is illuminated with a parallel
Gaussian beam of the appropriate size adjusted through a beam expander, for giving the desired beam size at
the sample plane. The orientation of the grating is usually chosen such that the 1st order is diffracted
perpendicular to the grating (θdiff = 0° ) to avoid tilted illumination of the image plane. Conjugation of the
grating (image) plane to the sample image plane is realized by a telescope consisting of a lens and the
microscope objective. (b) In temporal focusing of CGH beams the grating is placed at the image plane of CGH,
illuminated with the holographic pattern generated by addressing the corresponding phase on the SLM (inset).
(c) Similarly as in CGH, in temporal focusing of GPC beams the grating is illuminated with the intensity pattern
generated at the output (image) plane of the GPC configuration, when addressing the SLM with the
appropriate, in the simplest case binary, pattern (inset). In all panels θ denotes the incident angle of the
light-shaped beam onto the grating. In (b) and (c) the beam expander prior to the SLM is omitted for simplicity.
(Adapted from Ref. [201])
Fig. 8 Multiplexed temporally focused CGH patterns. Projection of temporally focused patterns in multiple
planes consists in a 3-step-approach: 1. beam amplitude shaping, here by CGH, 2. performing temporal
focusing, and 3. spatial multiplexing by using a SLM (SLM2) and 3D point-cloud CGH, at a Fourier plane after
dispersion of the spectral frequencies on the grating. The pattern generated through phase modulation (inset
SLM1) is projected onto the grating (F(X,Y) inset) and replicated by 3D point-cloud CGH to different 3D
positions (G(X,Y,Z) inset SLM2). The way SLM2 is illuminated is also shown in the inset. The resulting pattern
at the sample is a convolution of patterns F and G. (Reproduced from Ref. [201])
Fig. 9 Organotypic slices. (a) Organization of the hood for the dissection of organotypic slices. (b) Dissected
organotypic slices on PTFE membranes placed on inserts in a 6-well plate. (c) Transmitted light image of a
patched cell in a hippocampal organotypic slice. (d) Expression of a nuclear targeted fluorescent protein
(mRuby) following bulk infection with an AAV. The architecture of the hippocampus is maintained
3.4 Preparation of While the benefits of using organotypic slice cultures for prototyp-
Organotypic ing preparations for all-optical neurophysiology experiments were
Hippocampal Slice described in Sect. 2.2.1 of this chapter, it is important to highlight
Cultures that this preparation can also be used to address many fundamental
questions central to modern neuroscience.
All animal experiments must comply with national regulations.
Organotypic hippocampal slices are prepared from mice (here from
Janvier Labs, C57Bl6J WT) at post-natal day 8 (P8).
3.4.1 Solutions Table 1 presents the quantities of reagents required for the solu-
tions used during the preparation of organotypic hippocampal
slices (their product references are also listed).
Table 1
Solutions for organotypic hippocampal slice cultures preparation
11. Gently take apart the skull and transfer the brain in the Petri
dish containing the filter paper using a short and rounded
spatula. Place the brain on the filter paper.
12. Insert the spatula between the two hemispheres and gently
separate them.
13. Separate the cortex with the underlying hippocampus from the
brainstem, midbrain, and striatum using the spatula without
touching the hippocampus.
14. Place the cortex so that the hippocampus is exposed. Flip the
hippocampus over and out using the spatula.
15. Repeat steps 13–14 for the hippocampus of the remaining
hemisphere.
16. Use the wide-bore pipette to transfer the hippocampi to the
stage of the tissue chopper and align them perpendicularly to
the blade.
17. Remove any excess dissection medium from the stage to mini-
mize any motion of the hippocampi during the slicing process.
18. Cut 300 μm thick slices.
19. Using the narrow-bore pipette, flush the slices with some
dissection medium and transfer them to the 3rd petri dish.
20. Gently separate the slices with the pipette.
21. Transfer the best slices to the 4th petri dish and incubate at 4 °
C (or on ice) for at least 30 min (refer to Fig. 3 from Gogolla
et al. 2006 [203] for criteria of selection). Another pup can be
dissected during this time.
22. Following a minimum of 30 min of incubation at 4 °C, retrieve
the 6-well plate from the incubator and place each selected slice
in the middle of each square membrane, using the narrow-bore
pipette (Fig. 9b).
23. Remove any excess dissection medium around the slice using a
200 μL micropipette and put the plate back in the incubator
immediately.
Excess dissection medium affects gas exchange and, conse-
quently, slice health.
24. After 3 days, remove all culture medium below the insert and
replace it with 1 mL of fresh, warm neurobasal-A culture
medium.
25. Replace the culture medium every 3–4 days.
3.4.3 Bulk Infection Before infecting, it is essential to let the slices recover and adhere to
the membrane for at least 3 days in the incubator following slicing.
If possible, opt for AAVs as they are non-pathogenic, non-cyto-
toxic, and do not integrate in the host genome. Alternatively, to
avoid the use of viruses, one can utilize electroporation of plasmids.
36 Ruth R. Sims et al.
3.4.4 Troubleshooting – The slices should flatten and become transparent after a few days
in culture (Fig. 9c). Dead slices remain whitish and opaque.
– To know: WT slices exhibit autofluorescence.
– Depending on type of experiment to conduct, it is essential to be
aware that the use of antibiotics can affect the physiology of the
cells.
3.4.5 What Is Essential – Use an upright microscope because the membrane will make it
on the Day of Experiment? hard to see the cells.
– For optimal results, oxygenate the external solution.
– pH should be kept around 7.4 (with HEPES or bicarbonate &
bubbling).
– Take the slice out of the incubator only once everything on the
setup is ready.
– Under optimal conditions the slices may be re-used across mul-
tiple experimental sessions.
4 Notes
5 Outlook
Acknowledgments
References
1. Koch C (1997) Computation and the mind. 9. Dana H, Mohar B, Sun Y et al (2016) Sensi-
Nature 385:207–210. https://doi.org/10. tive red protein calcium indicators for imaging
1177/0002764297040006002 neural activity. elife 5:1–24. https://doi.org/
2. Emiliani V, Cohen AE, Deisseroth K, H€ausser 10.7554/eLife.12727
M (2015) All-optical interrogation of neural 10. Wang W, Kim CK, Ting AY (2019) Molecular
circuits. J Neurosci 35:13917–13926. tools for imaging and recording neuronal
https://doi.org/10.1523/JNEUROSCI. activity. Nat Chem Biol 15:101–110.
2916-15.2015 https://doi.org/10.1038/s41589-018-
3. Prakash R, Yizhar O, Grewe B et al (2012) 0207-0
Two-photon optogenetic toolbox for fast 11. Tian L, Hires SA, Mao T et al (2009) Imaging
inhibition, excitation and bistable modula- neural activity in worms, flies and mice with
tion. Nat Methods 9:1171–1179. https:// improved GCaMP calcium indicators. Nat
doi.org/10.1038/nmeth.2215 Methods 6:875–881. https://doi.org/10.
4. Klapoetke NC, Murata Y, Kim SS et al (2014) 1038/nmeth.1398
Independent optical excitation of distinct 12. Fan LZ, Kheifets S, Böhm UL et al (2020)
neural populations. Nat Methods 11:338– All-optical electrophysiology reveals the role
346. https://doi.org/10.1038/nmeth.2836 of lateral inhibition in sensory processing in
5. Mardinly AR, Oldenburg IA, Pégard NC et al cortical layer 1. Cell 180:521–535.e18.
(2018) Precise multimodal optical control of https://doi.org/10.1016/j.cell.2020.
neural ensemble activity. Nat Neurosci 21: 01.001
881–893. https://doi.org/10.1038/ 13. Abdelfattah AS, Kawashima T, Singh A et al
s41593-018-0139-8 (2019) Bright and photostable chemigenetic
6. Shemesh OA, Tanese D, Zampini V et al indicators for extended in vivo voltage imag-
(2017) Temporally precise single-cell resolu- ing. Science (80) 365:699–704. https://doi.
tion optogenetics. Nat Neurosci 20:1796– org/10.1126/science.aav6416
1806 14. Villette V, Chavarha M, Dimov IK et al
7. Baker CA, Elyada YM, Parra-Martin A, Bol- (2019) Ultrafast two-photon imaging of a
ton M (2016) Cellular resolution circuit high-gain voltage indicator in awake behaving
mapping in mouse brain with temporal- mice. Cell 179:1590–1608.e23. https://doi.
focused excitation of soma-targeted channelr- org/10.1016/j.cell.2019.11.004
hodopsin. eLife 5:1–15. https://doi.org/10. 15. Denk W, Strickler JH, Webb WW (1990)
7554/eLife.14193 Two-photon laser scanning fluorescence
8. Marshel JH, Kim YS, Machado TA, et al microscopy. Science (80) 248:73–76
(2019) Cortical layer–specific critical dynam- 16. Theer P, Hasan MT, Denk W (2003)
ics triggering perception. Science 365 Two-photon imaging to a depth of 1000 mu
40 Ruth R. Sims et al.
m in living brains by use of a Ti: Al2O3 regen- neuroscience. J Neurosci 40:4264. https://
erative amplifier. Opt Lett 28:1022–1024 doi.org/10.1523/JNEUROSCI.0103-20.
17. Theer P, Denk W (2006) On the fundamental 2020
imaging-depth limit in two-photon micros- 31. Nagel G, Ollig D, Fuhrmann M et al (2002)
copy. J Opt Soc Am A Opt Image Sci Vis 23: Channelrhodopsin-1: a light-gated proton
3139–3149 channel in green algae. Science (80) 296:
18. Helmchen F, Denk W (2005) Deep tissue 2395–2398. https://doi.org/10.1126/sci
two-photon microscopy. Nat Methods 2: ence.1072068
932–940. https://doi.org/10.1038/ 32. Nagel G, Szellas T, Huhn W et al (2003)
NMETH818 Channelrhodopsin-2, a directly light-gated
19. Ustione A, Piston DW (2011) A simple intro- cation-selective membrane channel. Proc
duction to multiphoton microscopy. J Natl Acad Sci U S A 100:13940–13945
Microsc 243:221–226. https://doi.org/10. 33. Boyden ES, Zhang F, Bamberg E et al (2005)
1111/j.1365-2818.2011.03532.x Millisecond-timescale, genetically targeted
20. Svoboda K, Yasuda R (2006) Principles of optical control of neural activity. Nat Neurosci
two-photon excitation microscopy and its 8:1263–1268
applications to neuroscience. Neuron 50: 34. Nikolic K, Grossman N, Grubb MS et al
823–839. https://doi.org/10.1016/j.neu (2009) Photocycles of Channelrhodopsin-2.
ron.2006.05.019 Photochem Photobiol 85:400–411. https://
21. Kandel ER, Schwartz JH, Jessell TM et al doi.org/10.1111/j.1751-1097.2008.
(2013) Principles of neural science, 5th edn. 00460.x
McGraw-Hill, New York 35. Kuhne J, Vierock J, Tennigkeit SA et al
22. Petersen CCH (2017) Whole-cell recording (2019) Unifying photocycle model for light
of neuronal membrane potential during adaptation and temporal evolution of cation
behavior. Neuron 95:1266–1281. https:// conductance in channelrhodopsin-2. Proc
doi.org/10.1016/j.neuron.2017.06.049 Natl Acad Sci U S A 116:9380–9389.
23. Poulet JFA, Petersen CCH (2008) Internal h t t p s : // d o i . o r g / 1 0 . 1 0 7 3 / p n a s .
brain state regulates membrane potential syn- 1818707116
chrony in barrel cortex of behaving mice. 36. Nagel G, Brauner M, Liewald JF et al (2005)
Nature 454:881–885. https://doi.org/10. Light activation of channelrhodopsin-2 in
1038/nature07150 excitable cells of Caenorhabditis elegans trig-
24. Gerstner W, Kreiter AK, Markram H, Herz gers rapid behavioral responses. Curr Biol 15:
AVM (1997) Neural codes: firing rates and 2279–2284. https://doi.org/10.1016/j.
beyond. Proc Natl Acad Sci U S A 94: cub.2005.11.032
12740–12741. https://doi.org/10.1073/ 37. Bi A, Cui J, Ma YP et al (2006) Ectopic
pnas.94.24.12740 expression of a microbial-type rhodopsin
25. Stein RB, Gossen ER, Jones KE (2005) Neu- restores visual responses in mice with photo-
ronal variability: noise or part of the signal? receptor degeneration. Neuron 50:23–33.
Nat Rev Neurosci 6:389–397. https://doi. https://doi.org/10.1016/j.neuron.2006.
org/10.1038/nrn1668 02.026
26. Engel AK, Fries P, Singer W (2001) Dynamic 38. Mahn M, Saraf-Sinik I, Patil P et al (2021)
predictions: oscillations and synchrony in Optogenetic silencing of neurotransmitter
top–down processing. Nat Rev Neurosci 2: release with a naturally occurring invertebrate
704–716. https://doi.org/10.1038/ rhodopsin. bioRxiv:2021.02.18.431673
35094565 39. Sridharan S, Gajowa MA, Ogando MB et al
27. Hubel DH, Wiesel TN (1959) Receptive (2022) High-performance microbial opsins
fields of single neurones in the cat’s striate for spatially and temporally precise perturba-
cortex. J Physiol 148:574–591. https://doi. tions of large neuronal networks.
org/10.1113/jphysiol.1959.sp006308 Neuron:1–17. https://doi.org/10.1016/j.
neuron.2022.01.008
28. Yuste R (2015) From the neuron doctrine to
neural networks. Nat Rev Neurosci 16:487– 40. Kishi KE, Kim YS, Fukuda M et al (2022)
497. https://doi.org/10.1038/nrn3962 Structural basis for channel conduction in
the pump-like channelrhodopsin ChRmine.
29. Salinas E, Sejnowski TJ, Medical HH et al Cell 185:672–689.e23. https://doi.org/10.
(2001) Salinas et al., 2001. 2 1016/j.cell.2022.01.007
30. Altimus CM, Marlin BJ, Charalambakis NE 41. Govorunova EG, Gou Y, Sineshchekov OA
et al (2020) The next 50 years of et al (2021) Kalium rhodopsins: natural
All-Optical Neuronal Manipulation with Wavefront Engineering 41
88. Wiegert JS, Gee CE, Oertner TG (2017) imaging in awake behaving animals. Nat
Single-cell electroporation of neurons. Cold Methods 13:1001–1004. https://doi.org/
Spring Harb Protoc 2017:135–138. https:// 10.1038/nMeth.4033
doi.org/10.1101/pdb.prot094904 100. Shain WJ, Vickers NA, Goldberg BB et al
89. LoTurco J, Manent JB, Sidiqi F (2009) New (2017) Novel deconvolution kernel for
and improved tools for in utero electropora- extended depth-of-field microscopy with a
tion studies of developing cerebral cortex. high-speed deformable mirror. Opt Lett 42:
Cereb Cortex 19:i120–i125. https://doi. 995–998. https://doi.org/10.1364/
org/10.1093/cercor/bhp033 MATH.2017.MW1C.2
90. Humpel C (2015) Neuroscience forefront 101. Dal Maschio M, De Stasi AM, Benfenati F,
review organotypic brain slice cultures: a Fellin T (2011) Three-dimensional in vivo
review. Neuroscience 305:86–98. https:// scanning microscopy with inertia-free focus
doi.org/10.1016/j.neuroscience.2015. control. Opt Lett 36:3503–3505. https://
07.086 doi.org/10.1364/OL.36.003503
91. Jutras MJ, Buffalo EA (2010) Synchronous 102. Packer AM, Peterka DS, Hirtz JJ et al (2012)
neural activity and memory formation. Curr Two-photon optogenetics of dendritic spines
Opin Neurobiol 20:150–155. https://doi. and neural circuits. Nat Methods 9:1171–
org/10.1016/j.conb.2010.02.006 1179. https://doi.org/10.1038/nmeth.
92. Fan GY, Fujisaki H, Miyawaki A et al (1999) 2249
Video-rate scanning two-photon excitation 103. Carrillo-reid L, Yang W, Bando Y et al (2016)
fluorescence microscopy and ratio imaging Imprinting and recalling cortical ensembles.
with cameleons. Biophys J 76:2412–2420 Science (80) 353:691–694. https://doi.org/
93. Amos WB, White JG (2003) How the confo- 10.1126/science.aaf7560
cal laser scanning microscope entered 104. Nikolenko V, Watson BO, Araya R et al
biological research. Biol Cell 95:335–342. (2008) SLM microscopy: scanless
https://doi.org/10.1016/S0248-4900(03) two-photon imaging and photostimulation
00078-9 with spatial light modulators. Front Neural
94. Bouchard MB, Voleti V, Mendes CS et al Circuits 2:5. https://doi.org/10.3389/
(2015) Swept confocally-aligned planar exci- neuro.04.005.2008
tation (SCAPE) microscopy for high-speed 105. Göbel W, Kampa BM, Helmchen F, Go W
volumetric imaging of behaving organisms. (2007) Imaging cellular network dynamics
Nat Photonics 9:113–119. https://doi.org/ in three dimensions using fast 3D laser scan-
10.1038/nphoton.2014.323 ning. Nat Methods 4:73–79. https://doi.
95. Salomé R, Kremer Y, Dieudonné S et al org/10.1038/NMETH989
(2006) Ultrafast random-access scanning in 106. Blochet B, Bourdieu L, Gigan S (2017)
two-photon microscopy using acousto-optic Focusing light through dynamical samples
deflectors. J Neurosci Methods 154:161– using fast closed-loop wavefront optimiza-
174. https://doi.org/10.1016/j.jneumeth. tion. Opt Lett 42:4994–4997. https://doi.
2005.12.010 org/10.1364/OL.42.004994
96. Otsu Y, Bormuth V, Wong J et al (2008) 107. Grewe BF, Voigt FF, van ’t Hoff M et al
Optical monitoring of neuronal activity at (2011) Fast two-layer two-photon imaging
high frame rate with a digital random-access of neuronal cell populations using an electri-
multiphoton (RAMP) microscope. J Neurosci cally tunable lens. Biomed Opt Express 2:
Methods 173:259–270 2035–2046. https://doi.org/10.1364/
97. Reddy GD, Saggau P (2005) Fast three- BOE.2.002035
dimensional laser scanning scheme using 108. Weisenburger S, Tejera F, Demas J et al
acousto-optic deflectors. J Biomed Opt 10: (2019) Volumetric Ca2+ imaging in the
064038. https://doi.org/10.1117/1. mouse brain using hybrid multiplexed
2141504 sculpted light microscopy. Cell 177:1–17.
98. Reddy GD, Kelleher K, Fink R, Saggau P https://doi.org/10.1016/j.cell.2019.
(2008) Three-dimensional random access 03.011
multiphoton microscopy for functional imag- 109. Kong L, Tang J, Little JP et al (2015) Con-
ing of neuronal activity. Nat Neurosci 11: tinuous volumetric imaging via an optical
713–720. https://doi.org/10.1038/nn. phase-locked ultrasound lens. Nat Methods
2116 12:759–762. https://doi.org/10.1038/
99. Nadella KMNS, Roš H, Baragli C et al (2016) nmeth.3476
Random access scanning microscopy for 3D
44 Ruth R. Sims et al.
using stereoscopy (vTwINS). Nat Methods 145. Lutz C, Otis TS, DeSars V et al (2008) Holo-
14:420–426. https://doi.org/10.1101/ graphic photolysis of caged neurotransmit-
073742 ters. Nat Methods 5:821–827. https://doi.
134. Haist T, Schönleber M, Tiziani H (1997) org/10.1038/nmeth.1241
Computer-generated holograms from 146. Papagiakoumou E, de Sars V, Oron D, Emi-
3D-objects written on twisted-nematic liquid liani V (2008) Patterned two-photon illumi-
crystal displays. Opt Commun 140:299–308. nation by spatiotemporal shaping of
https://doi.org/10.1016/S0030-4018(97) ultrashort pulses. Opt Express 16:22039–
00192-2 22047
135. Polin M, Ladavac K, Lee S-H et al (2005) 147. Papagiakoumou E, Anselmi F, Bègue A et al
Optimized holographic optical traps. Opt (2010) Scanless two-photon excitation of
Express 13:5831–5845 channelrhodopsin-2. Nat Methods 7:848–
136. Anselmi F, Ventalon C, Begue A et al (2011) 854. https://doi.org/10.1038/nmeth.1505
Three-dimensional imaging and photostimu- 148. Oron D, Papagiakoumou E, Anselmi F, Emi-
lation by remote-focusing and holographic liani V (2012) Two-photon optogenetics.
light patterning. Proc Natl Acad Sci U S A Prog Brain Res 196:119–143. https://doi.
108:19504–19509. https://doi.org/10. org/10.1016/B978-0-444-59426-6.
1073/pnas.1109111108 00007-0
137. Yang S, Papagiakoumou E, Guillon M et al 149. Hernandez O, Papagiakoumou E, Tanese D
(2011) Three-dimensional holographic et al (2016) Three-dimensional spatiotempo-
photostimulation of the dendritic arbor. J ral focusing of holographic patterns. Nat
Neural Eng 8:46002. S1741-2560(11) Commun 7:11928. https://doi.org/10.
87640-8 [pii]. https://doi.org/10.1088/ 1038/ncomms11928
1741-2560/8/4/046002 150. Accanto N, Molinier C, Tanese D et al (2018)
138. Thalhammer G, Bowman RW, Love GD et al Multiplexed temporally focused light shaping
(2013) Speeding up liquid crystal SLMs using for high-resolution multi-cell targeting.
overdrive with phase change reduction. Opt Optica 5:1478–1491
Express 21:1779–1797 151. Prevedel R, Verhoef AJ, Pernı́a-Andrade AJ
139. Xue Y, Berry KP, Boivin JR et al (2019) Scan- et al (2016) Fast volumetric calcium imaging
less volumetric imaging by selective access across multiple cortical layers using sculpted
multifocal multiphoton microscopy. Optica light. Nat Methods 13:1021–1028. https://
6:76. https://doi.org/10.1364/OPTICA.6. doi.org/10.1038/nmeth.4040
000076 152. Zahid M, Velez-Fort M, Papagiakoumou E
140. Zhang T, Hernandez O, Chrapkiewicz R et al et al (2010) Holographic photolysis for mul-
(2019) Kilohertz two-photon brain imaging tiple cell stimulation in mouse hippocampal
in awake mice. Nat Methods 16:1119–1122. slices. PLoS One 5:e9431
https://doi.org/10.1038/s41592-019- 153. Bègue A, Papagiakoumou E, Leshem B et al
0597-2 (2013) Two-photon excitation in scattering
141. Zhang Z, Russell LE, Packer AM, et al (2018) media by spatiotemporally shaped beams and
Closed-loop all-optical manipulation of neu- their application in optogenetic stimulation.
ral circuits in vivo. https://doi.org/10.1038/ Biomed Opt Express 4:2869–2879
s41592-018-0183-z 154. Dal Maschio M, Donovan JC, Helmbrecht
142. Yang W, Carrillo-reid L, Bando Y et al (2018) TO, Baier H (2017) Linking neurons to net-
Simultaneous two-photon optogenetics and work function and behavior by two-photon
imaging of cortical circuits in three dimen- holographic optogenetics and volumetric
sions. elife 7:e32671. https://doi.org/10. imaging. Neuron 94:774–789.e5. https://
7554/eLife.32671 doi.org/10.1016/j.neuron.2017.04.034
143. Carrillo-Reid L, Han S, Yang W et al (2019) 155. Moretti C, Antonini A, Bovetti S et al (2016)
Controlling visually guided behavior by holo- Scanless functional imaging of hippocampal
graphic recalling of cortical ensembles. Cell networks using patterned two-photon illumi-
178:447–457.e5. https://doi.org/10.1016/ nation through GRIN lenses. Biomed Opt
j.cell.2019.05.045 Express 7:3958. https://doi.org/10.1364/
144. Yang W, Miller JK, Carrillo-Reid L et al BOE.7.003958
(2016) Simultaneous multi-plane imaging of 156. Förster D, Dal Maschio M, Laurell E, Baier H
neural circuits. Neuron 89:269–284. https:// (2017) An optogenetic toolbox for unbiased
doi.org/10.1016/j.neuron.2015.12.012 discovery of functionally connected cells in
46 Ruth R. Sims et al.
neural circuits. Nat Commun 8:116. https:// 167. Papadopoulos IN, Jouhanneau JS, Takahashi
doi.org/10.1038/s41467-017-00160-z N et al (2020) Dynamic conjugate F-SHARP
157. Forli A, Vecchia D, Binini N et al (2018) microscopy. Light Sci Appl 9. https://doi.
Two-photon bidirectional control and imag- org/10.1038/s41377-020-00348-x
ing of neuronal excitability with high spatial 168. May MA, Kummer KK, Kress M et al (2021)
resolution in vivo. Cell Rep 22:2809–2817. Fast holographic scattering compensation for
https://doi.org/10.1016/j.celrep.2018. deep tissue biological imaging. 1–6
02.063 169. Abrahamsson S, Chen J, Hajj B et al (2012)
158. Chen I-W, Ronzitti E, Lee BR et al (2019) In Fast multicolor 3D imaging using aberration-
vivo submillisecond two-photon optogenetics corrected multifocus microscopy. Nat Meth-
with temporally focused patterned light. J ods 1–6. https://doi.org/10.1038/nmeth.
Neurosci 39:3484–3497. https://doi.org/ 2277
10.1523/JNEUROSCI.1785-18.2018 170. Quirin S, Peterka DS, Yuste R (2013) Instan-
159. Qian Y, Piatkevich KD, Larney B et al (2019) taneous three-dimensional sensing using spa-
A genetically encoded near-infrared fluores- tial light modulator illumination with
cent calcium ion indicator. Nat Methods. extended depth of field imaging. Opt Express
https://doi.org/10.1038/s41592-018- 21:16007. https://doi.org/10.1364/OE.
0294-6 21.016007
160. Shemetov AA, Monakhov MV, Zhang Q et al 171. Rickgauer JP, Deisseroth K, Tank DW (2014)
(2021) A near-infrared genetically encoded Simultaneous cellular-resolution optical per-
calcium indicator for in vivo imaging. Nat turbation and imaging of place cell firing
Biotechnol 39:368–377. https://doi.org/ fields. Nat Neurosci 17:1816–1824. https://
10.1038/s41587-020-0710-1 doi.org/10.1038/nn.3866
161. Jackson CT, Jeong S, Dorlhiac GF, Landry 172. Denk W, Strickler J, Webb W (1990)
MP (2021) Advances in engineering near- Two-photon laser scanning fluorescence
infrared luminescent materials. iScience 24: microscopy. Science (80) 52:1778–1779
102156. https://doi.org/10.1016/j.isci. 173. Denk W, Piston DW, Webb WW (1995) Two
2021.102156 photon molecular excitation in laser-
162. Nöbauer T, Skocek O, Pernı́a-Andrade AJ scannning microscopy. In: Pawley JB
et al (2017) Video rate volumetric Ca2+ imag- (ed) Handbook of confocal microscopy. Ple-
ing across cortex using seeded iterative num, pp 445–458
demixing (SID) microscopy. Nat Methods 174. Denk W, Svoboda K (1997) Photon upman-
14:811–818. https://doi.org/10.1038/ ship: why multiphoton imaging is more than a
nmeth.4341 gimmick. Neuron 18:351–357
163. Taylor MA, Nöbauer T, Pernia-Andrade A 175. Xu C, Webb WW (1996) Measurement of
et al (2018) Brain-wide 3D light-field imag- two-photon excitation cross sections of
ing of neuronal activity with speckle- molecular fluorophores with data from
enhanced resolution. Optica 5:345. https:// 690 to 1050 nm. J Opt Soc Am B 13:481.
doi.org/10.1364/OPTICA.5.000345 https://doi.org/10.1364/JOSAB.13.
164. Ducros M, Goulam Houssen Y, Bradley J et al 000481
(2013) Encoded multisite two-photon 176. Xu C, Zipfel W, Shear JB et al (1996) Multi-
microscopy. Proc Natl Acad Sci U S A 110: photon fluorescence excitation: new spectral
13138–13143. https://doi.org/10.1073/ windows for biological nonlinear microscopy.
pnas.1307818110 Proc Natl Acad Sci U S A 93:10763–10768
165. Tsyboulski D, Orlova N, Saggau P (2017) 177. Rohrbacher A, Olarte OE, Villamaina V et al
Amplitude modulation of femtosecond laser (2017) Multiphoton imaging with blue-
pulses in the megahertz range for frequency- diode-pumped SESAM-modelocked Ti:sap-
multiplexed two-photon imaging. Opt phire oscillator generating 5 nJ 82 fs pulses.
Express 25:9435. https://doi.org/10.1364/ Opt Express 25:10677. https://doi.org/10.
oe.25.009435 1364/oe.25.010677
166. Papadopoulos IN, Jouhanneau JS, Poulet 178. Rehbinder J, Brückner L, Wipfler A et al
JFA, Judkewitz B (2017) Scattering compen- (2014) Multimodal nonlinear optical micros-
sation by focus scanning holographic aberra- copy with shaped 10 fs pulses. Opt Express
tion probing (F-SHARP). Nat Photonics 11: 22:28790. https://doi.org/10.1364/oe.22.
116–123. https://doi.org/10.1038/ 028790
nphoton.2016.252
All-Optical Neuronal Manipulation with Wavefront Engineering 47
179. Chaigneau E, Ronzitti E, Gajowa AM et al 193. Lee JG, Hill WT (2014) Spatial shaping for
(2016) Two-photon holographic stimulation generating arbitrary optical dipole traps for
of ReaChR. Front Cell Neurosci 10:234 ultracold degenerate gases. Rev Sci Instrum
180. Schmidt E, Oheim M (2020) Infrared excita- 85:103106. https://doi.org/10.1063/1.
tion induces heating and calcium microdo- 4895676
main hyperactivity in cortical astrocytes. 194. Papagiakoumou E, Bègue A, Leshem B et al
Biophys J 119:2153–2165. https://doi.org/ (2013) Functional patterned multiphoton
10.1016/j.bpj.2020.10.027 excitation deep inside scattering tissue. Nat
181. Podgorski K, Ranganathan G (2016) Brain Photonics 7:274–278
heating induced by near infrared lasers during 195. Bañas A, Kopylov O, Villangca M et al (2014)
multi-photon microscopy. J Neurophysiol GPC light shaper: static and dynamic experi-
116:1012–1023. https://doi.org/10.1101/ mental demonstrations. Opt Express 22:
057364 23759. https://doi.org/10.1364/OE.22.
182. Patterson GH, Piston DW (2000) Photo- 023759
bleaching in two-photon excitation micros- 196. Bañas A, Palima D, Villangca M et al (2014)
copy. Biophys J 78:2159–2162. https://doi. GPC light shaper for speckle-free one- and
org/10.1016/S0006-3495(00)76762-2 two-photon contiguous pattern excitation.
183. Koester HJ, Baur D, Uhl R, Hell SW (1999) Opt Express 22:5299. https://doi.org/10.
Ca2+ fluorescence imaging with pico- and 1364/OE.22.005299
femtosecond two-photon excitation: signal 197. Bañas A, Glückstad J (2017) Holo-GPC:
and photodamage. Biophys J 77:2226– holographic generalized phase contrast. Opt
2236. https://doi.org/10.1016/S0006- Commun 392:190–195. https://doi.org/
3495(99)77063-3 10.1016/j.optcom.2017.01.036
184. Hopt A, Neher E (2001) Highly nonlinear 198. Oron D, Silberberg Y (2005) Spatiotemporal
photodamage in two-photon fluorescence coherent control using shaped, temporally
microscopy. Biophys J 80:2029–2036. focused pulses. Opt Express 13:9903–9908
https://doi.org/10.1016/S0006-3495(01) 199. Oron D, Silberberg Y (2015) Temporal
76173-5 focusing microscopy. Cold Spring Harb Pro-
185. Picot A, Dominguez S, Liu C et al (2018) toc 2015:145–151. https://doi.org/10.
Thermal model of temperature rise under 1101/pdb.top085928
in vitro and in vivo two-photon optogenetics 200. Durst ME, Zhu G, Xu C (2008) Simulta-
brain stimulation. Cell Rep 24:1243–1253 neous spatial and temporal focusing in non-
186. Maioli V, Boniface A, Mahou P et al (2020) linear microscopy. Opt Commun 281:1796–
Fast in vivo multiphoton light-sheet micros- 1805. https://doi.org/10.1016/j.optcom.
copy with optimal pulse frequency. Biomed 2007.05.071
Opt Express 11:6012–6026. https://doi. 201. Papagiakoumou E, Ronzitti E, Emiliani V
org/10.1364/boe.400113 (2020) Scanless two-photon excitation with
187. Glückstad J (1996) Phase contrast image syn- temporal focusing. Nat Methods 17:571–
thesis. Opt Commun 130:225–230 581. https://doi.org/10.1038/s41592-
188. Zernike F (1958) How I discovered phase 020-0795-y
contrast. Science (80) 121:345–349 202. Llobet L, Montoya J, López-Gallardo E,
189. Wang Z, Millet L, Mir M et al (2011) Spatial Ruiz-Pesini E (2015) Side effects of culture
light interference microscopy (SLIM). Opt media antibiotics on cell differentiation. Tis-
Express 19:1016–1026 sue Eng – Part C Methods 21:1143–1147.
190. Rodrigo PJ, Daria VR, Glückstad J (2005) https://doi.org/10.1089/ten.tec.2015.
Four-dimensional optical manipulation of 0062
colloidal particles. Appl Phys Lett 86:74103 203. Gogolla N, Galimberti I, DePaola V, Caroni P
191. Rodrigo PJ, Daria VR, Glückstad J (2004) (2006) Preparation of organotypic hippocam-
Real-time three-dimensional optical micro- pal slice cultures for long-term live imaging.
manipulation of multiple particles and living Nat Protoc 1:1165–1171. https://doi.org/
cells. Opt Lett 29:2270–2272 10.1038/nprot.2006.168
192. Rodrigo PJ, Perch-Nielsen IR, Alonzo CA, 204. Horton NG, Wang K, Kobat D et al (2013)
Glückstad J (2006) GPC-based optical micro- In vivo three-photon microscopy of subcorti-
manipulation in 3D real-time using a single cal structures within an intact mouse brain.
spatial light modulator. Opt Express 14: Nat Photonics 7:205–209. https://doi.org/
13107–13112 10.1038/nphoton.2012.336
48 Ruth R. Sims et al.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International
License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing, adaptation, distribution
and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative Commons license,
unless indicated otherwise in a credit line to the material. If material is not included in the chapter’s Creative
Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use,
you will need to obtain permission directly from the copyright holder.
Chapter 2
Abstract
The goal of this chapter is to establish a framework to evaluate imaging methodologies for all-optical
neurophysiology experiments. This is not an exhaustive review of fluorescent indicators and imaging
modalities but rather aims to distill the functional imaging principles driving the choice of both. Scientific
priorities determine whether the imaging strategy is based on an “optimal fluorescent indicator” or
“optimal imaging modality.” The choice of the first constrains the choice of the second due to each’s
contributions to the fluorescence budget and signal-to-noise ratio of time-varying fluorescence changes.
1 Introduction
Eirini Papagiakoumou (ed.), All-Optical Methods to Study Neuronal Function, Neuromethods, vol. 191,
https://doi.org/10.1007/978-1-0716-2764-8_2, © The Author(s) 2023
49
50 Peter Quicke et al.
Fig. 1 Widefield, one-photon imaging (left) and multiphoton scanning (right) imaging modalities; not to scale
Balancing the Fluorescence Budget April 2020 51
Fig. 2 The challenges of 3D imaging in brain tissue. (a) Techniques that are not optically sectioning suffer from
blur caused by the collection of light from out-of-focus planes. (b) Light rays can be scattered before exiting
the tissue, changing their apparent origin. In aggregate, this leads to a blurring effect that drastically increases
with depth. Reproduced from [88], CC BY-SA 4.0
2.2 Challenge 2: While certain model organisms (e.g., larval zebrafish, C. elegans)
(Most) Brains Scatter are transparent, most brains, especially mammalian brains, scatter
Light light strongly. Fluorescence excitation efficiency decreases with
increasing imaging depth as excitation light is scattered and
absorbed, resulting in the degradation of the excitation point-
spread function (PSF). Moreover, light excited in one spatial loca-
tion can be scattered before collection and detected as though
arising from somewhere else (Fig. 2, right). This degrades image
contrast and confuses analysis of fluorescence time courses in adja-
cent areas.
Together, out-of-focus fluorescence and scattered fluorescence
produce blurring, which drastically increases with imaging depth
[46]. The need for optical sectioning and robustness to scattering
has driven development and application of two- [30] and three-
photon [45, 121] (or “multiphoton”) point scanning modalities,
which achieve both. Due to non-linear dependence on excitation
52 Peter Quicke et al.
1
Intensity is often normalized to photon energy in the multiphoton imaging literature [130], different from the
standard definition of W/m2.
54 Peter Quicke et al.
1
Δt = , ð6Þ
Racq N Z
where Racq is the frame acquisition rate and NZ is the number of
planes imaged per volume. In contrast, multiphoton point rastering
modalities integrate fluorescence from each location for only a small
fraction of the acquisition period
1
Δt = , ð7Þ
Racq N X N Y N Z
where NX and NY are the lateral frame dimensions.
Consider the case where NZ = 1 to image a single plane, such
that Racq is the frame rate, and compare the widefield to the
multiphoton raster scan case. For the widefield case, Δt = 1/Racq.
For the multiphoton scanning case, digitizing each frame for exam-
ple as NX = NY = 256 pixels, we see that Δt = 1/(2562Racq). If we
base our SNR comparison solely on differences
pffiffiffiffiffiffiffiffiffiffiffi in Δt, we see that
the SNR for the widefield case is 2562 = 256 times that of
multiphoton scanning! Note that this factor could be even bigger
if the scanning modality involves significant “dead time” (i.e.,
galvonometric scanner turn around). Cameras also require read-
out time, during which they are not integrating photons; however,
this is typically around 10 μs per line for a modern sCMOS camera,
a small fraction of the frame period, while scanning dead time can
be around 10–20% of each frame.
The optical sectioning and robustness to scattering achieved by
2P and 3P point scanning modalities over widefield comes at high
cost to the fluorescence budget, F0. Note that in the multiphoton
point scanning case, F0 is inversely proportional to the product of
the number of locations (pixels or voxels, NXNYNZ) monitored,
and the rate at which they are monitored, Racq. This implies that F0
can be maintained by performing high Racq acquisition of few
locations or low Racq acquisition of many locations. For example,
at the high rate extreme, 10 kilohertz acquisition of voltage tran-
sients from a single “voxel” has been achieved in two photon by
parking the beam over a dendritic spine [3]. At the other extreme,
two-photon “mesoscopes” image FOVs multiple millimeters wide
at low frame rates (0.1–10 Hz) [102, 106] using slow and highly
sensitive calcium indicators [25, 29].
From Eq. 5, we see that SNR can be increased by: 1. use of a
high sensitivity (large ΔF/F0) fluorescence reporter and/or
2. increasing F0. F0 can be increased by: maximizing ϕ with efficient
photo-sensors, high collection NA, etc.; selecting a bright fluoro-
phore and exciting it at wavelengths for which its brightness, σ n, is
highest; maximizing fluorophore concentration, CFl, and exciting
and collecting fluorescence over the largest possible voxel, VFl;
exciting the fluorophore with the highest feasible intensity,
Balancing the Fluorescence Budget April 2020 55
Table 1
Summary of variables determining temporal SNR for functional fluorescence imaging
4.1 Temporal Here, indicator “temporal profile” refers to the convolution of the
Considerations time course of a biophysical process (e.g., fast membrane potential
variations vs. slow changes in calcium or neurotransmitter concen-
tration) with the kinetics of the fluorescent indicator. Here we refer
to τOff as a general term for the rate of decay back to baseline of the
fluorescence transient induced by an action potential or other brief
event. We note, however, that many indicators’ kinetics are not
described by a simple mono-exponential decay. Importantly, τOff
56 Peter Quicke et al.
Fig. 4 The challenges of voltage imaging. Three issues make voltage imaging
more challenging than calcium imaging. First, faster intrinsic kinetics limit the
photon integration period. Second, voltage indicators must lie in the membrane
or degrade the signal; this limits the volume of indicators that can be integrated
to measure the signal. Lastly, membranes where signal arises are tightly packed
in the brain; fluorescent signals from overlapping membranes wash out single-
cell signals. Adapted from original in [88], CC BY-SA 4.0
5.1 Fluorescence Regarding SNR, the fluorescence excitation subsystem can be char-
Excitation acterized in terms of light intensity, < In > , and the per pixel, per
integration period excited volume, VFl. Together with the fluoro-
phore cross-section (σ n), concentration (CFl), integration period
(Δt), and spatial distribution, these parameters determine the max-
imum available fluorescence excitation budget (Fgen; Eqs. 3, 4). The
illumination wavelength and fluorophore cross-sections determine
whether the system favors one-photon, two-photon, or three-
photon fluorescence excitation. Two- and three-photon excitation
requires ultrafast pulsed lasers with megahertz2 repetition frequen-
cies to achieve RFl sufficient for imaging. One-photon fluorescence
excitation rates, RFl, vary widely depending on indicator brightness
(σ 1) and illumination intensity (< I > ) and generally exceed RFl for
two- and three-photon modalities, which typically excite < 0.1
photons per laser pulse [101].
5.1.1 Fluorescence The fluorescence excitation rate critically depends on the degree to
Excitation Volume, VFl which the fluorescence excitation is parallelized. Widefield excita-
tion, which excites fluorescence in all locations throughout a vol-
ume simultaneously, features the highest degree of parallelization,
thus maximizing the fluorescence excitation budget. Focusing a
laser beam to a diffraction-limited point and serially scanning that
point correspond to lowest parallelization and excitation budget. In
between widefield and scanned point excitation, the excitation light
can be sculpted into many forms, including a large point (scanned-
temporal focusing; S-TeFo, [87, 118]), multiple scanned (spinning
disk confocal [108, 126], multifocal 2P [15, 55, 78, 89, 98, 120,
127]) or static (computer-generated holography, [23, 32, 85,
122]) points, a line (TeFo line scanning [28], SLAP [53], vTWINS
[103], Bessel beams [19, 66]), whole planes [5, 20, 51, 54, 82, 96,
97, 128], and extended shapes patterned directly onto structures of
interest [21, 39, 79, 109, 110]. It is important to note that not all
fluorescence photons contribute useful signal. For example, a
higher proportion of photons excited through two-photon point
scanning contribute to image formation compared to widefield
imaging, where photons excited outside the plane of focus are not
imaged and can smear the temporal signals extracted from in-focus
ROIs. It is also important to note that VFl introduced earlier in this
chapter refers to the per location fluorescence volume, not the total
spot, line, or sheet volume, and therefore depends on the spatial
discretization performed by the collection subsystem.
2
Imaging a 128 × 128 pixel FOV at 10 Hz with one pulse per pixel, the lower limit, requires a
10 × 1282 = 0.16 MHz repetition rate.
62 Peter Quicke et al.
Table 2
Δt for the different excitation volume shapes in scanning configurations
1
Scan type Δt = R acq ×
Single spot, rastered 1
NXNY NZ
k spots, rastered k
NXNY NZ
Line, scanned 1
NY NZ
Sheet, scanned 1
NZ
5.2 Fluorescence The fluorescence detection system determines how the fluores-
Detection cence excitation budget is exploited to form images. Sensors for
fluorescence detection fall into two categories: single and multi-
channel.
Fig. 5 Balancing indicator and imaging strategy contributions to SNR. The imaging modality (upper triangle)
and fluorescent indicator (lower triangle) properties together determine fluorescence transient SNR (horizontal
axis; equation, top). Importantly, the SNR of a “low fluorescence budget” imaging modality can be compen-
sated by a high ΔF/F0 or “high fluorescence budget” indicator and vice versa. The vertical dashed lines situate
three example indicator/modality pairings with respect to each’s relative contribution to SNR
7 Notes
Fig. 7 Demonstration of the effect of Poisson noise on SNR. Reproduced from [88], CC BY-SA 4.0. Inspired by
[31]
Acknowledgements
References
1. Abdelfattah AS, Farhi SL, Zhao Y, Brinks D, 12. Beaulieu DR, Davison IG, Kılıç K, Bifano TG,
Zou P, Ruangkittisakul A, Platisa J, Pieribone Mertz J (2020) Simultaneous multiplane
VA, Ballanyi K, Cohen AE, et al. (2016) A imaging with reverberation two-photon
bright and fast red fluorescent protein voltage microscopy. Nature Methods 17(3):283–286
indicator that reports neuronal activity in 13. Beck C, Zhang D, Gong Y (2019) Enhanced
organotypic brain slices. J Neurosci 36(8): genetically encoded voltage indicators
2458–2472 advance their applications in neuroscience.
2. Abdelfattah AS, Kawashima T, Singh A, Curr Opin Biomed Eng 12:111–117
Novak O, Liu H, Shuai Y, Huang Y-C, 14. Berridge MJ, Lipp P, Bootman MD (2000)
Campagnola L, Seeman SC, Yu J, et al. The versatility and universality of calcium sig-
(2019) Bright and photostable chemigenetic nalling. Nat Rev Mol cell Biol 1(1):11–21
indicators for extended in vivo voltage imag- 15. Bewersdorf J, Pick R, Hell SW (1998) Multi-
ing. Science 365(6454):699–704 focal multiphoton microscopy. Optics Letters
3. Acker CD, Yan P, Loew LM (2011) Single- 23(9):655–657
voxel recording of voltage transients in den- 16. Bindocci E, Savtchouk I, Liaudet N,
dritic spines. Biophysical Journal 101(2): Becker D, Carriero G, Volterra A (2017)
L11–3 Three-dimensional ca2+ imaging advances
4. Adam Y, Kim JJ, Lou S, Zhao Y, Xie ME, understanding of astrocyte biology. Science
Brinks D, Wu H, Mostajo-Radji MA, 356(6339):eaai8185
Kheifets S, Parot V, et al. (2019) Voltage 17. Bloodgood BL, Sabatini BL (2007) Ca2+ sig-
imaging and optogenetics reveal behaviour- naling in dendritic spines. Curr Opin Neuro-
dependent changes in hippocampal dynamics. biol 17(3):345–351
Nature 569(7756):413–417
18. Blunck R, Chanda B, Bezanilla F (2005)
5. Ahrens MB, Orger MB, Robson DN, Li JM, Nano to micro-fluorescence measurements
Keller PJ (2013) Whole-brain functional of electric fields in molecules and genetically
imaging at cellular resolution using light- specified neurons. J Membrane Biol 208(2):
sheet microscopy. Nature Methods 10(5):413 91–102
6. Aimon S, Katsuki T, Jia T, Grosenick L, 19. Botcherby EJ, Booth MJ, Juškaitis R, Wilson
Broxton M, Deisseroth K, Sejnowski TJ, T (2008) Real-time extended depth of field
Greenspan RJ (2019) Fast near-whole– brain microscopy. Optics Express 16(26):
imaging in adult Drosophila during responses 21843–21848. https://doi.org/10.1364/
to stimuli and behavior. PLoS Biology 17(2): OE.16.021843. http://www.opticsexpress.
e2006732 org/abstract.cfm?URI=oe-16-26-21843
7. Akerboom J, Carreras Calderón N, Tian L, 20. Bouchard MB, Voleti V, Mendes CS,
Wabnig S, Prigge M, Tolö J, Gordus A, Lacefield C, Grueber WB, Mann RS, Bruno
Orger MB, Severi KE, Macklin JJ, et al. RM, Hillman EMC (2015) Swept confocally-
(2013) Genetically encoded calcium indica- aligned planar excitation (scape) microscopy
tors for multi-color neural activity imaging for high-speed volumetric imaging of behav-
and combination with optogenetics. Front ing organisms. Nature Photonics 9(2):113
Mol Neurosci 6:2
21. Bovetti S, Moretti C, Zucca S, Dal
8. Anselmi F, Ventalon C, Bègue A, Ogden D, Maschio M, Bonifazi P, Fellin T (2017)
Emiliani V (2011) Three-dimensional imag- Simultaneous high-speed imaging and opto-
ing and photostimulation by remote-focusing genetic inhibition in the intact mouse brain.
and holographic light patterning. Proc Natl Scientific Reports 7:40041
Acad Sci 108(49):19504–19509
22. Breuninger T, Greger K, Stelzer EHK (2007)
9. Antic S, Zecevic D (1995) Optical signals Lateral modulation boosts image quality in
from neurons with internally applied voltage- single plane illumination fluorescence micros-
sensitive dyes. J Neurosci Offic J Soc Neurosci copy. Optics Letters 32(13):1938–1940
15(2):1392–1405
23. Castanares ML, Gautam V, Drury J,
10. Baker CA, Elyada YM, Parra A, Bolton MM Bachor H, Daria VR (2016) Efficient multi-
(2016) Cellular resolution circuit mapping site two-photon functional imaging of neuro-
with temporal-focused excitation of soma- nal circuits. Biomed Opt Exp 7(12):
targeted channelrhodopsin. Elife 5:e14193 5325–5334
11. Bando Y, Sakamoto M, Kim S, Ayzenshtat I, 24. Chen JL, Voigt FF, Javadzadeh M,
Yuste R (2019) Comparative evaluation of Krueppel R, Helmchen F (2016) Long-
genetically encoded voltage indicators. Cell
Reports 26(3):802–813
70 Peter Quicke et al.
47. Hochbaum DR, Zhao Y, Farhi SL, (2014) Independent optical excitation of dis-
Klapoetke N, Werley CA, Kapoor V, Zou P, tinct neural populations. Nature Methods
Kralj JM, Maclaurin D, Smedemark- 11(3):338
Margulies N, et al. (2014) All-optical electro- 58. Knöpfel T, Dı́ez-Garcı́a J, Akemann W (2006)
physiology in mammalian neurons using engi- Optical probing of neuronal circuit dynamics:
neered microbial rhodopsins. Nature genetically encoded versus classical fluores-
Methods 11(8):825 cent sensors. Trends Neurosci 29(3):
48. Hontani Y, Xia F, Xu C (2021) Multicolor 160–166. ISSN:0166-2236. https://doi.
three-photon fluorescence imaging with org/10.1016/j.tins.2006.01.004. http://
single-wavelength excitation deep in mouse www.sciencedirect.com/science/article/pii/
brain. Science Advances 7(12):eabf3531. S0166223606000051
https://doi.org/10.1126/sciadv.abf3531 59. Kondo M, Kobayashi K, Ohkura M, Nakai J,
49. Hoover EE, Young MD, Chandler EV, Matsuzaki M (2017) Two-photon calcium
Luo A, Field JJ, Sheetz KE, Sylvester AW, imaging of the medial prefrontal cortex and
Squier JA (2011) Remote focusing for pro- hippocampus without cortical invasion.
grammable multi-layer differential multipho- eLIFE 6:e26839
ton microscopy. Biomed Opt Exp 2(1): 60. Kulkarni RU, Miller EW (2017) Voltage
113–122 imaging: pitfalls and potential. Biochemistry
50. Howe CL, Quicke P, Song P, Verinaz-Jadan- 56(39):5171–5177
H, Dragotti PL, Foust AJ (2022) Comparing 61. Lamy CM, Chatton J-Y (2011) Optical prob-
synthetic refocusing to deconvolution for the ing of sodium dynamics in neurons and astro-
extraction of neuronal calcium transients from cytes. Neuroimage 58(2):572–578
light fields. Neurophotonics 9(4):041404 62. Lecoq J, Orlova N, Grewe BF (2019) Wide.
51. Huisken J, Swoger J, Del Bene F, Wittbrodt J, fast. deep: Recent advances in multiphoton
Stelzer EHK (2004) Optical sectioning deep microscopy of in vivo neuronal activity. J Neu-
inside live embryos by selective plane illumi- rosci 39(46):9042–9052. https://doi.org/
nation microscopy. Science 305(5686): 10.1523/jneurosci.1527-18.2019
1007–1009 63. Lim D, Mertz J, Chu KK (2008) Wide-field
52. Jing M, Zhang P, Wang G, Feng J, Mesik L, fluorescence sectioning with hybrid speckle
Zeng J, Jiang H, Wang S, Looby JC, Gua- and uniform-illumination microscopy. Optics
gliardo NA, et al. (2018) A genetically Letters 33(16):1819–1821
encoded fluorescent acetylcholine indicator 64. Lim DH, Mohajerani MH, LeDue J, Boyd J,
for in vitro and in vivo studies. Nature Bio- Chen S, Murphy TH (2012) In vivo large-
technology 36(8):726–737 scale cortical mapping using
53. Kazemipour A, Novak O, Flickinger D, Mar- channelrhodopsin-2 stimulation in transgenic
vin JS, Abdelfattah AS, King J, Borden PM, mice reveals asymmetric and reciprocal rela-
Kim JJ, Al-Abdullatif SH, Deal PE, et al. tionships between cortical areas. Front Neural
(2019) Kilohertz frame-rate two-photon Circuits 6:11
tomography. Nature Methods 16(8):778 65. Lin JY, Knutsen PM, Muller A, Kleinfeld D,
54. Keller PJ, Ahrens MB (2015) Visualizing Tsien RY (2013) ReaChR: a red-shifted vari-
whole-brain activity and development at the ant of channelrhodopsin enables deep tran-
single-cell level using light-sheet microscopy. scranial optogenetic excitation. Nature
Neuron 85(3):462–483 Neuroscience 16(10):1499
55. Kim KH, Buehler C, Bahlmann K, Ragan T, 66. Lu R, Sun W, Liang Y, Kerlin A, Bierfeld J,
Lee W-CA, Nedivi E, Heffer EL, Fantini S, So Seelig JD, Wilson DE, Scholl B, Mohar B,
PTC (2007) Multifocal multiphoton micros- Tanimoto M, et al. (2017) Video-rate volu-
copy based on multianode photomultiplier metric functional imaging of the brain at syn-
tubes. Optics Express 15(18):11658–11678 aptic resolution. Nature Neuroscience 20(4):
56. Kim MK, Lim CS, Hong JT, Han JH, Jang 620
H-Y, Kim HM, Cho BR (2010) Sodium-ion- 67. Mahn M, Gibor L, Patil P, Cohen-Kashi
selective two-photon fluorescent probe for Malina K, Oring S, Printz Y, Levy R,
in vivo imaging. Angewandte Chemie Int Lampl I, Yizhar O (2018) High-efficiency
Edn 49(2):364–367 optogenetic silencing with soma-targeted
57. Klapoetke NC, Murata Y, Kim SS, Pulver SR, anion-conducting channelrhodopsins. Nature
Birdsey-Benson A, Cho YK, Morimoto TK, Communications 9(1):1–15
Chuong AS, Carpenter EJ, Tian Z, et al.
72 Peter Quicke et al.
68. Manita S, Miyazaki K, Ross WN (2011) Syn- A (2017) Video rate volumetric Ca2+ imag-
aptically activated Ca2+ waves and NMDA ing across cortex using seeded iterative demix-
spikes locally suppress voltage-dependent ing (SID) microscopy. Nature Methods
ca2+ signalling in rat pyramidal cell dendrites. 14(8):811–818
J Physiol 589(20):4903–4920 81. Olarte OE, Andilla J, Artigas D, Loza-Alvarez
69. Marvin JS, Borghuis BG, Tian L, Cichon J, P (2015) Decoupled illumination detection in
Harnett MT, Akerboom J, Gordus A, Rennin- light sheet microscopy for fast volumetric
ger SL, Chen T-W, Bargmann CI, et al. imaging. Optica 2(8):702
(2013) An optimized fluorescent probe for 82. Oron D, Tal E, Silberberg Y (2005) Scanning-
visualizing glutamate neurotransmission. less depth-resolved microscopy. Optics
Nature Methods 10(2):162 Express 13(5):1468–1476
70. Marvin JS, Scholl B, Wilson DE, Podgorski K, 83. Packer AM, Russell LE, Dalgleish HWP,
Kazemipour A, Müller JA, Schoch S, José H€ausser M (2014) Simultaneous all-optical
Urra Quiroz F, Rebola N, Bao H, et al. manipulation and recording of neural circuit
(2018) Stability, affinity, and chromatic var- activity with cellular resolution in vivo. Nature
iants of the glutamate sensor iGluSnFR. Methods 12(2):140–146. https://doi.org/
Nature Methods 15(11):936–939 10.1038/nmeth.3217
71. Marvin JS, Shimoda Y, Magloire V, Leite M, 84. Pégard NC, Liu H-Y, Antipa N, Gerlock M,
Kawashima T, Jensen TP, Kolb I, Knott EL, Adesnik H, Waller L (2016) Compressive
Novak O, Podgorski K, et al. (2019) A genet- light-field microscopy for 3d neural activity
ically encoded fluorescent sensor for in vivo recording. Optica 3(5):517–524
imaging of GABA. Nature Methods, p. 1 85. Pozzi P, Gandolfi D, Tognolina M, Chirico G,
72. McMahon SM, Jackson MB (2018) An incon- Mapelli J, D’Angelo E (2015) High-
venient truth: calcium sensors are calcium throughput spatial light modulation
buffers. Trends Neurosci 41(12):880–884 two-photon microscopy for fast functional
73. Mertz J, Kim J (2010) Scanning light-sheet imaging. Neurophotonics 2(1):015005
microscopy in the whole mouse brain with 86. Prevedel R, Yoon Y-G, Hoffmann M, Pak N,
HiLo background rejection. J Biomed Opt Wetzstein G, Kato S, Schrödel T, Raskar R,
15(1):016027 Zimmer M, Boyden ES, Vaziri A (2014)
74. Miyazaki K, Manita S, Ross WN (2012) Simultaneous whole-animal 3D imaging of
Developmental profile of localized spontane- neuronal activity using light-field microscopy.
ous ca2+ release events in the dendrites of rat Nature Methods 11(7):727–730
hippocampal pyramidal neurons. Cell Cal- 87. Prevedel R, Verhoef AJ, Pernia-Andrade AJ,
cium 52(6):422–432 Weisenburger S, Huang BS, Nöbauer T,
75. Murakami T, Yoshida T, Matsui T, Ohki K Fernández A, Delcour JE, Golshani P,
(2015) Wide-field ca2+ imaging reveals visu- Baltuska A, et al. (2016) Fast volumetric cal-
ally evoked activity in the retrosplenial area. cium imaging across multiple cortical layers
Front Mol Neurosci 8:20 using sculpted light. Nature Methods
76. Neil MAA, Juškaitis R, Wilson T (1997) 13(12):1021–1028
Method of obtaining optical sectioning by 88. Quicke P (2019) Improved methods for func-
using structured light in a conventional tional neuronal imaging with genetically
microscope. Optics Letters 22(24): encoded voltage indicators. PhD thesis,
1905–1907 Imperial College, London
77. Ng R, Adams A, Footer M, Horowitz M, 89. Quicke P, Reynolds S, Neil M, Knöpfel T,
Levoy M (2006) Light field microscopy. Schultz SR, Foust AJ (2018) High speed
ACM Trans Graph 25(3):924–934 functional imaging with source localized mul-
78. Nielsen T, Fricke M, Hellweg D, Andresen P tifocal two-photon microscopy. Biomed Opt
(2001) High efficiency beam splitter for mul- Exp 9(8):3678–3693
tifocal multiphoton microscopy. J Microscopy 90. Quicke P, Song C, McKimm EJ, Milosevic
201(3):368–376 MM, Howe CL, Neil M, Schultz SR, Antic
79. Nikolenko V, Watson BO, Araya R, SD, Foust AJ, Knöpfel T (2019) Single-
Woodruff A, Peterka DS, Yuste R (2008) neuron level one-photon voltage imaging
SLM microscopy: scanless two-photon imag- with sparsely targeted genetically encoded
ing and photostimulation using spatial light voltage indicators. Front Cell Neurosci 13:
modulators. Front Neural Circuits 2:5 39. ISSN:1662-5102. https://doi.org/10.
80. Nobauer T, Skocek O, Pernia-Andrade AJ, 3389/fncel.2019.00039. https://www.
Weilguny L, Traub FM, Molodtsov MI, Vaziri
Balancing the Fluorescence Budget April 2020 73
K (2015) SPED light sheet microscopy: fast Wetzstein G, Deisseroth K (2015) Extended
mapping of biological system structure and field-of-view and increased-signal 3d holo-
function. Cell 163(7):1796–1806 graphic illumination with time-division multi-
114. Tsuda S, Kee MZL, Cunha C, Kim J, Yan P, plexing. Optics Express 23(25):
Loew LM, Augustine GJ (2013) Probing the 32573–32581
function of neuronal populations: combining 123. Yang W, Miller J-EK, Carrillo-Reid L,
micromirror-based optogenetic photostimu- Pnevmatikakis E, Paninski L, Yuste R, Peterka
lation with voltage-sensitive dye imaging. DS (2016) Simultaneous multi-plane imaging
Neuroscience Research 75(1):76–81 of neural circuits. Neuron 89(2):269–284
115. van Drongelen W (2007) Signal processing 124. Yang W, Carrillo-Reid L, Bando Y, Peterka
for neuroscientists: Introduction to the anal- DS, Yuste R (2018) Simultaneous
ysis of physiological signals. J Motor Behav two-photon imaging and two-photon opto-
39(2):158 genetics of cortical circuits in three dimen-
116. Venkatachalam V, Cohen AE (2014) Imaging sions. Elife 7:e32671
GFP-based reporters in neurons with multi- 125. Yizhar O, Fenno LE, Prigge M, Schneider F,
wavelength optogenetic control. Biophys J Davidson TJ, O’shea DJ, Sohal VS, Goshen I,
107(7):1554–1563 Finkelstein J, Paz JT, et al. (2011) Neocorti-
117. Villette V, Chavarha M, Dimov IK, Bradley J, cal excitation/inhibition balance in informa-
Pradhan L, Mathieu B, Evans SW, tion processing and social dysfunction.
Chamberland S, Shi D, Yang R, et al. (2019) Nature 477(7363):171–178
Ultrafast two-photon imaging of a high-gain 126. Yoshida E, Terada S-I, Tanaka YH,
voltage indicator in awake behaving mice. Kobayashi K, Ohkura M, Nakai J, Matsuzaki
Cell 179(7):1590–1608 M (2018) In vivo wide-field calcium imaging
118. Weisenburger S, Tejera F, Demas J, Chen B, of mouse thalamocortical synapses with an 8 k
Manley J, Sparks FT, Traub FM, Daigle T, ultra-high-definition camera. Scientific
Zeng H, Losonczy A, et al. (2019) Volumet- Reports 8(1):1–15
ric ca2+ imaging in the mouse brain using 127. Zhang T, Hernandez O, Chrapkiewicz R,
hybrid multiplexed sculpted light microscopy. Shai A, Wagner MJ, Zhang Y, Wu C-H, Li
Cell 177(4):1050–1066 JZ, Inoue M, Gong Y, et al. (2019) Kilohertz
119. Willadt S, Canepari M, Yan P, Loew LM, Vogt two-photon brain imaging in awake mice.
KE (2014) Combined optogenetics and volt- Nature Methods 16(11):1119–1122
age sensitive dye imaging at single cell resolu- 128. Zhu G, Van Howe J, Durst M, Zipfel W, Xu C
tion. Front Cell Neurosci 8:311 (2005) Simultaneous spatial and temporal
120. Wu J, Liang Y, Chen S, Hsu C-L, focusing of femtosecond pulses. Optics
Chavarha M, Evans SW, Shi D, Lin MZ, Tsia express 13(6):2153–2159
KK, Ji N (2020) Kilohertz two-photon fluo- 129. Zinter JP, Levene MJ (2011) Maximizing
rescence microscopy imaging of neural activ- fluorescence collection efficiency in multipho-
ity in vivo. Nature Methods 17(3):287–290 ton microscopy. Optics Express 19(16):
121. Xu C, Zipfel W, Shear JB, Williams RM, Webb 15348–15362
WW (1996) Multiphoton fluorescence excita- 130. Zipfel WR, Williams RM, Webb WW (2003)
tion: new spectral windows for biological Nonlinear magic: multiphoton microscopy in
nonlinear microscopy. Proc Natl Acad Sci the biosciences. Nature Biotechnology
USA 93(20):10763–10768 21(11):1369–1377
122. Yang SJ, Allen WE, Kauvar I, Andalman AS,
Young NP, Kim CK, Marshel JH,
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International
License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing, adaptation, distribution
and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative Commons license,
unless indicated otherwise in a credit line to the material. If material is not included in the chapter’s Creative
Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use,
you will need to obtain permission directly from the copyright holder.
Chapter 3
Abstract
Understanding how the brain orchestrates neuronal activity to finely produce and regulate behavior is an
intriguing yet challenging task. In the last years, the progressive refinement of optical techniques and light-
based molecular tools allowed to start addressing open questions in cellular and systems neuroscience with
unprecedented resolution and specificity. Currently, all-optical experimental protocols for simultaneous
recording of the activity of large cell populations with the concurrent modulation of the firing rate at cellular
resolution represent an invaluable tool. In this scenario, it is becoming everyday more evident the
importance of sampling and probing the circuit mechanisms not just in a single plane, but extending the
exploration to the entire volume containing the involved circuit components. Here, we focus on the design
principles and the hardware architectures of all-optical approaches allowing for studying the neuronal
dynamics at cellular resolution across a volume of the brain.
1 Introduction
Eirini Papagiakoumou (ed.), All-Optical Methods to Study Neuronal Function, Neuromethods, vol. 191,
https://doi.org/10.1007/978-1-0716-2764-8_3, © The Author(s) 2023
75
76 Matteo Bruzzone et al.
2 Methods
2.1 Molecular and In general, all-optical probing of brain circuits relies on the possi-
Technical Constraints bility of concurrently recording and modulating neuronal activity
with high resolution [14]. This assumes the compresence of light-
gated actuators and light-based activity reporters within the same
preparation and, in some cases, within the same excitation volume.
Then, combining these two families of tools together in the same
experimental scenario requires evaluating the biochemical and bio-
physical properties of these molecules along with technical aspects
and working constraints associated with their use.
78 Matteo Bruzzone et al.
Fig. 1 Spectral properties of the most common actuator-reporter pairs, used for all-optical approaches. In the
upper part are reported the light-driven actuators. The range with a 2P action spectrum greater than 60% is
shown with a darkened mark indicating the wavelengths corresponding to the activation peak. On the right
side, the corresponding τOFF is reported. This indicates the photocurrent decay time at the offset of the
illumination. In the lower part, excitation and emission spectra are presented for the most common activity
reporters. Black lines highlight the actuator-reporter pairs reported [16–19]
Optical Approaches to Probe Neuronal Circuits in 3D 79
2.4 Off-Target The possibility of using all-optical approaches to probe brain cir-
Activation and Somatic cuits critically depends, along with recording the neuronal activity
Opsin Targeting without perturbing the system, on the capability to target the
neuronal modulation with high spatial precision and selectivity.
This assumes that the photostimulation impacts exclusively or
mostly on the identified targets, i.e., sets of neuronal somata. This
is a goal not straightforward to achieve as dendrites and axons from
many different cells surround the targeted cell body and constitute
a dense mesh of neuronal processes, frequently expressing them-
selves the opsin. This represents a possible issue of off-target acti-
vation, i.e., indirect photostimulation of neuronal components
other than those targeted. This is beyond the limit of the hardware
design, independently from the specific optical implementation to
drive the light-based actuator. To overcome this limitation, in the
last years, many labs developed light-driven actuators specifically
targeted to the cell soma, importantly reducing the expression
along the axons and the other processes [28–30]. This frequently
results in more efficient stimulation (in terms of required power
density), and the strong reduction of indirect effects mediated by
passing-by neuronal processes.
Fig. 2 General hardware layout for 3D imaging. This typically includes three elements: the first one with the
source and the intensity modulation unit, a component of the optical path designed for scanning the beam
along the longitudinal direction (in blue), and a module for scanning the excitation beam along the lateral
dimension (in red). The different elements are conjugated by means of relay optics
3.2 Methods for The most common way for scanning the sample along the lateral
Scanning the Sample dimension relies on a pair of galvanometric mirrors, conjugated to
Along the Lateral the BFP of the objective. These are 3–6 mm wide mirrors mounted
Dimension with an orthogonal optical axis whose orientations can be tuned by
proper control signals. A change in the orientation with respect to
the propagation of the incoming beam introduces a linear phase
gradient to the wavefront, resulting in the lateral offset of the beam
with respect to the center of the field of view (FOV). In a raster
scanning scheme, one mirror sweeps over a line, and the other
jumps from one line to the following one. With a typical pixel
dwell time of about 4 μs, this design results in a frame acquisition
time lower than a second for a mesh of 512 512 pixels. An
8–12 kHz resonant galvanometric mirror can replace the line scan-
ning galvo, reducing the line acquisition time to 62–41 μs, respec-
tively, and increasing the acquisition rates up to 30–45 frames per
second for a pixel matrix of 512 512 elements. In this second
scenario, there is no direct possibility to adjust the pixel dwell time
and its extremely short value, typically 120–80 ns, heavily impacts
the number of collected photons per pixel and, ultimately, the
image SNR. In parallel to the solutions described above, to steer
the light beams, it is possible to modulate the light wavefront in an
inertia-free approach using acousto-optic deflectors (AODs)
[37, 38]. These are active optical elements with a crystal window
Optical Approaches to Probe Neuronal Circuits in 3D 83
3.3 Methods for Adjusting remotely the longitudinal position of an excitation spot
Scanning the Sample generally relies on the possibility to impose a curvature on the
Along the Longitudinal wavefront profile in a plane optically conjugated to the BFP of the
Dimension objective. Considering the sole imaging purposes, possibly the
simplest implementation is to place a lens with a controllable focal
length, either at the level of the BFP downstream of the XY scanner
and the scan lens-tube lens pair or in a plane optically conjugated to
the BFP, upstream the scanner. An electrically tunable lens (ETL) is
an example of such a device, which is composed of a liquid volume
enclosed between a glass and an elastic polymer membrane, i.e.,
effectively a plano-convex liquid lens [40]. An electromagnetic coil
driven by an electric current exerts pressure on the liquid, increas-
ing the membrane curvature and thus the lens focal power. ETLs
are capable of relatively fast settling times, typically between 5 and
15 ms, in response to a step-like control signal and can easily reach,
depending on the objective properties, a few hundreds of microns
of focal range with limited distortion of the point spread function
(PSF) (Fig. 3). ETL driving, either with staircase-like or sawtooth
control signals, in coordination with the frame acquisition allows
for the plane-after-plane acquisition of a volume. According to a
similar scheme, a tunable acoustic gradient index of refraction lens
(or TAG lens) can be used to quickly scan the sample longitudinal
dimension [41, 42]. In this case, high-frequency driving of a piezo-
electric actuator generates in a viscous medium a cylindrical acous-
tic wave whose constructive interference produces periodic changes
in medium density and, consequently, in the lens optical power.
The resonance frequency of these devices, typically a few hundreds
84 Matteo Bruzzone et al.
Fig. 3 Multiphoton imaging across multiple planes in the olfactory bulb of a mouse expressing GCaMP6f. A
basic configuration for multiphoton imaging is complemented with an electrically tunable lens to sample
quasi-simultaneously different planes of the brain volume. Four different planes are shown with activity
profiles extracted from the corresponding cells
Fig. 4 Scheme representing some of the hardware configurations for photostimulation in two- or three-
dimensions. Starting from the left side, we report the solution for sequential photostimulation with diffraction-
limited spot with spatially extended photostimulation patches (in gray). Then, the parallel configurations are
indicated, relaying on scanning of multiplexed beamlets (yellow), scanless simultaneous excitation of multiple
patches without (green), and with temporal focusing (red)
4.1 The General The core components for 3D light-assisted modulation of neuronal
Features of a circuits are optical configurations engineered to generate arbitrary
Photostimulation Train light distributions in the sample space and are generally grouped
into two main classes [26]: sequential and parallel excitation
approaches (Fig. 4). In the first design, a single beam, either
diffraction-limited or with an engineered PSF, quickly travels across
88 Matteo Bruzzone et al.
Fig. 5 Computer-Generated Holography. Engineering the light wavefront at the BFP allows for rendering
arbitrary light intensity distributions at the sample. Introducing a phase correction resembling a Fresnel lens
results in a change in the convergence properties of a propagating beam, moving the position of the focus
longitudinally. Similarly, with a linear gradient of phase delay applied at the BFP, the position of the focus is
moved laterally. Multiplexing different diffractive optical elements (DOEs) allows for rendering arbitrary light
distribution at the sample
5.1 The Hardware Different can be the optical configurations to support an all-optical
Integration circuit probing framework. It is clear that integrating and coordi-
nating two components requires the identifying the appropriate
Optical Approaches to Probe Neuronal Circuits in 3D 93
Fig. 6 Layout of the integration scheme of the imaging and photostimulation arms. In cerulean is indicated the
path of the imaging beam with the assembly to control the z-position upstream of the XY scanner. In red are
shown the possible insertion points of the photostimulation path with respect to the optical components
required for imaging
5.2 Beam Co- The first aspect to consider is evaluating how the three light beams
registration used for the two approaches (excitation of the light-gated actuator,
Procedures excitation of the fluorescent reporter, and fluorescence emission by
the reporter) are reaching or leaving the sample, whether these
beams travel through the same objective, and whether the objective
is kept still during the acquisition. This is usually the most common
case [20, 23, 54, 69] but are also possible architectures relying on
independent arms, one for photostimulation and fluorescence col-
lection with a moving objective, the other for reporter excitation
[62]. Having one objective still assumes the use of remote optical
Optical Approaches to Probe Neuronal Circuits in 3D 95
6 Notes
7 Conclusions
Acknowledgments
References
1. Weisenburger S, Vaziri A (2018) A guide to 10. Fenno L, Yizhar O, Deisseroth K (2011) The
emerging technologies for large-scale and development and application of optogenetics.
whole brain optical imaging of neuronal activ- Annu Rev Neurosci 34:389–412. https://doi.
ity. Annu Rev Neurosci 41:431–452. https:// org/10.1146/annurev-neuro-
doi.org/10.1146/annurev-neuro- 061010-113817
072116-031458 11. Kim CK, Adhikari A, Deisseroth K (2017)
2. Tian L, Hires SA, Looger LL (2012) Imaging Integration of optogenetics with complemen-
neuronal activity with genetically encoded cal- tary methodologies in systems neuroscience.
cium indicators. Cold Spring Harb Protoc Nat Rev Neurosci 18:222–235. https://doi.
2012:pdb.top069609. https://doi.org/10. org/10.1038/nrn.2017.15
1101/pdb.top069609 12. Hegemann P, Nagel G (2013) From channelr-
3. Bando Y, Sakamoto M, Kim S et al (2019) hodopsins to optogenetics. EMBO Mol Med
Comparative evaluation of genetically encoded 5:173–176. https://doi.org/10.1002/
voltage indicators. Cell Rep 26:802–813.e4. emmm.201202387
https://doi.org/10.1016/j.celrep.2018. 13. Wiegert JS, Mahn M, Prigge M et al (2017)
12.088 Silencing neurons: tools, applications, and
4. Nakai J, Ohkura M, Imoto K (2001) A high experimental constraints. Neuron 95:
signal-to-noise Ca2+ probe composed of a sin- 504–529. https://doi.org/10.1016/j.neu
gle green fluorescent protein. Nat Biotechnol ron.2017.06.050
19:137–141. https://doi.org/10.1038/ 14. Maschio MD, Difato F, Beltramo R et al
84397 (2010) Simultaneous two-photon imaging
5. Chen T-W, Wardill TJ, Sun Y et al (2013) and photo-stimulation with structured light
Ultrasensitive fluorescent proteins for imaging illumination. Opt Express 18:18720–18731.
neuronal activity. Nature 499:295–300. https://doi.org/10.1364/OE.18.018720
https://doi.org/10.1038/nature12354 15. Venkatachalam V, Cohen AE (2014) Imaging
6. Dana H, Sun Y, Mohar B et al (2019) High- GFP-based reporters in neurons with multi-
performance calcium sensors for imaging activ- wavelength optogenetic control. Biophys J
ity in neuronal populations and microcompart- 107:1554–1563. https://doi.org/10.1016/j.
ments. Nat Methods 16:649–657. https://doi. bpj.2014.08.020
org/10.1038/s41592-019-0435-6 16. Carrillo-Reid L, Yang W, Bando Y et al (2016)
7. Dana H, Mohar B, Sun Y et al Sensitive red Imprinting and recalling cortical ensembles.
protein calcium indicators for imaging neural Science 353:691–694. https://doi.org/10.
activity. elife 5. https://doi.org/10.7554/ 1126/science.aaf7560
eLife.12727 17. Carrillo-Reid L, Han S, Yang W et al (2019)
8. Mohr MA, Bushey D, Aggarwal A et al (2020) Controlling visually guided behavior by holo-
jYCaMP: an optimized calcium indicator for graphic recalling of cortical ensembles. Cell
two-photon imaging at fiber laser wavelengths. 178:447–457.e5. https://doi.org/10.1016/j.
Nat Methods 17:694–697. https://doi.org/ cell.2019.05.045
10.1038/s41592-020-0835-7 18. Szabo V, Ventalon C, De Sars V et al (2014)
9. Panzeri S, Harvey CD, Piasini E et al (2017) Spatially selective holographic photoactivation
Cracking the neural code for sensory percep- and functional fluorescence imaging in freely
tion by combining statistics, intervention, and behaving mice with a fiberscope. Neuron 84:
behavior. Neuron 93:491–507. https://doi. 1157–1169. https://doi.org/10.1016/j.neu
org/10.1016/j.neuron.2016.12.036 ron.2014.11.005
98 Matteo Bruzzone et al.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International
License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing, adaptation, distribution
and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative Commons license,
unless indicated otherwise in a credit line to the material. If material is not included in the chapter’s Creative
Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use,
you will need to obtain permission directly from the copyright holder.
Chapter 4
Abstract
Understanding brain function requires technologies that can monitor and manipulate neural activity with
cellular resolution and millisecond precision in three dimensions across large volumes. These technologies
are best designed using interdisciplinary approaches combining optical techniques with reporters and
modulators of neural activity. While advances can be made by separately improving optical resolution or
opsin effectiveness, optimizing both systems together matches the strengths and constraints of different
approaches to create a solution optimized for the needs of neuroscientists. To achieve this goal, we first
developed a new multiphoton photoexcitation method, termed 3D-Scanless Holographic Optogenetics
with Temporal focusing (3D-SHOT), that enables simultaneous photoactivation of arbitrary sets of
neurons in 3D. Our technique uses point-cloud holography to place multiple copies of a temporally focused
disc, matched to the dimensions of a neuron’s cell body, anywhere within the operating volume of the
microscope. However, since improved placement of light, on its own, is not sufficient to allow precise
control of neural firing patterns, we also developed and tested optogenetic actuators ST-ChroME and
ST-eGtACR1 that fully leverage the new experimental capabilities of 3D-SHOT. The synergy of fast opsins
matched with our technology allows reliable, precisely timed control of evoked action potentials and
enables on-demand read-write operations with unprecedented precision. In this chapter, we review the
steps necessary to implement 3D-SHOT and provide a guide to selecting ideal opsins that will work with
it. Such collaborative, interdisciplinary approaches will be essential to develop the experimental capabilities
needed to gain causal insight into the fundamental principles of the neural code underlying perception and
behavior.
1 Introduction
Eirini Papagiakoumou (ed.), All-Optical Methods to Study Neuronal Function, Neuromethods, vol. 191,
https://doi.org/10.1007/978-1-0716-2764-8_4, © The Author(s) 2023
101
102 Ian Antón Oldenburg et al.
2 Methods
2.1 3D-SHOT Optical Our experimental setup, shown in Fig. 1, is based on the standard
System Design design of a holographic microscope with a spatial light modulator
(SLM) in the Fourier domain (pupil plane). The SLM shapes the
phase of a coherent femtosecond laser light source to synthesize
custom 3D shapes [54, 55] digitally synthesized with Computer
Generated Holography [35–37] (CGH). Unlike scanning
approaches, CGH wide-area holograms matched to the dimensions
of each neuron’s soma enable simultaneous, flash-based, activation
of a large number of opsin molecules, yielding photocurrents with
fast kinetics [39].
In most brain structures, neurons are distributed continuously
in 3D, not in discrete layers. Therefore, the inability of 2D opto-
genetics approaches (i.e., 2D CGH with temporal focusing) to
target neurons at any arbitrary number of axial planes simulta-
neously is a major obstacle for large-scale optogenetic interrogation
of neural circuits. To overcome this outstanding challenge
3D-SHOT leverages the advantages of 3D-CGH, to simulta-
neously address neurons in custom locations. To make 3D CGH
and temporal focusing mutually compatible, 3D-SHOT forgoes
the ability to create custom patterns. Instead, the optical path is
optimized to holographically replicate multiple identical copies of a
temporally focused excitation pattern, termed “custom temporally
106 Ian Antón Oldenburg et al.
Fig. 1 Experimental setup for 3D-SHOT. This is made of two consecutive optical systems. First, a diffraction
grating and a rotating diffuser are imaged onto each other by a f-f optical relay. This assembly shapes
femtosecond laser pulses both spatially and spectrally to create a custom temporally focused pattern (CTFP)
matched to the dimensions of a neuron soma. The resulting engineered point spread function is then spatially
modulated by a second system that enables 3D computer-generated holography (CGH). A spatial light
modulator (SLM) placed in the Fourier domain modulates the phase of the CTFP to target custom 3D locations
with a point-cloud hologram. The resulting sculpted illumination pattern replicates identical copies of the CTFP
at each targeted location. The 3D hologram is further demagnified by a tube lens and a microscope objective.
A zero-order block eliminates any remaining undiffracted light from the hologram. The grating frequency
determines the spectral dispersion, “a”, and the diffuser determines the beam dimension “b” at the surface of
the SLM. Those parameters, along with the focal lengths of lenses L1–L4, are adjusted to match the desired
addressable volume and CTFP dimensions within constraints imposed by the SLM size, the laser source, and
the numerical aperture of the microscope
2.1.1 3D-SHOT Design 1. The size of the CTFP must be adjusted to match the dimensions
Parameters “d” of a neuron. In a typical implementation of 3D SHOT, the
3D hologram (see Fig. 1) is relayed into a microscope with an
additional tube lens (L5, f ¼ f5) and microscope objective (L6,
108 Ian Antón Oldenburg et al.
b ¼ f 3 αd
a ¼ δλ f 1 f 3 =ð l f 2 Þ:
The numerical aperture NA of the microscope objective is
generally the limiting factor that defines the accessible volume
and CTFP minimal dimensions for 3D-SHOT. The SLM pattern
(see Fig. 1) of width a + 2b is imaged onto the back aperture of the
microscope objective. Hence, to fully capture the light modulated
by the SLM, a suitable design constraint is to ensure (a + 2b)f5/
f4 < NA f6. In this configuration, the characteristic size, δz, of the
CTFP along the (z) axis in the demagnified hologram under the
microscope objective is given by
2
δz ¼ λ f 4 f 6 = a f 5
2.2 Characterization Our primary goal is not to render a visually accurate hologram but
and Performance instead to increase contrast for two-photon excitation at selected
Metrics for 3D-SHOT locations while avoiding inadvertent photoactivation of
non-targeted areas, which relaxes constraints on hologram compu-
tation. Hence, the traditional metrics used to characterize imaging
systems such as resolution, contrast, and speckle are not adequate
to evaluate the capabilities of 3D-SHOT in experimental
applications.
Instead, to evaluate the capabilities of 3D-SHOT and quantify
how two-photon absorption is spatially distributed in 3D, we
placed a thin fluorescent film on a microscope slide under the
excitation objective to record the corresponding two-photon fluo-
rescence image with a fixed sub-stage objective coupled to a camera
High-Speed All-Optical Neural Interfaces with 3D Temporally Focused Holography 111
Fig. 2 Optical characterization of the spatial resolution of CGH vs 3D-SHOT. (a) We used a fluorescent
calibration slide and an inverted microscope to quantify two-photon excitation in 3D. (b) For conventional
holography, we consider a 10 μm diameter disk target, and show (from top to bottom) projection views of
two-photon absorption in the (x,y), (z,y), and (z,x) planes. (c) With 3D-SHOT, the CTFP was adjusted to a 10 μm
diameter target and the same projection views were recorded. (d) We measured the FWHM of the radial (top)
or axial (bottom) PSF measured through brain slices of varying thickness (* indicates p < 0.05, Kruskal-Wallis
Test with multiple comparison correction. Data represent the mean and standard deviation of n 5
observations for each thickness of brain tissue). (Adapted from Pegard et al. [7])
Fig. 3 3D-SHOT generates axially confined photoactivation. (a) A photostimulation pattern generated with CGH
(top) or 3D-SHOT (bottom) was mechanically stepped along the optical axis (z) and passed through a cell
expressing opsin. Photocurrents were recorded in the whole-cell voltage-clamp configuration in neurons. (b)
FWHM of the characteristic response profile for both methods at various power levels. (c) Photocurrent
response profile for CGH (left) and 3D-SHOT (right) with a 10 μm disk target and different power levels. (d)
Spatial profile of two-photon evoked spiking of a L2/3 pyramidal neuron in a mouse brain slice (left) in the
radial dimension. Black: CGH; red: 3D-SHOT, p < 0.56 Mann-Whitney U-test, and (right), along the axial
dimension ( p < 0.006, Mann-Whitney U-test). (e) Quantification of the FWHM comparing CGH and 3D-SHOT.
(f) Full volumetric assessment of photostimulation resolution, points throughout the volume were tested, but
only points that elicited spike probability greater than zero are shown. (Adapted from Pegard et al. [7])
2.2.2 3D-SHOT We next quantified the physiological spatial resolution of CGH and
Photostimulation with 3D-SHOT in neurons by measuring the spiking probability along
Single-Neuron Resolution the radial direction in the imaging (x,y) plane and along the optical
(z) axis. We compared holography and 3D-SHOT by projecting a
single photostimulation target placed at a distance (x,y,z) from a
patched neuron in mouse brain slice, either with single copy of the
114 Ian Antón Oldenburg et al.
Fig. 4 3D-SHOT provides cellular resolution photostimulation in a large volume through digital focusing. (a) To
quantify the spatial resolution of 3D-SHOT as a function of hologram target depth, we recorded the spike
probability in cortical neurons while digitally targeting varying positions along the optical axis (z), and
measuring resolution by mechanically sweeping the objective over the entire (z) range and measuring the
response at each point. (b) Spike probability in cortical neurons while targeting the same cell from different
axial displacements ( p ¼ 0.2, Kruskal-Wallis Test). (c) Spike probability resolution as a function of digital
displacement – shaded green colors denote mechanical sweeps across the optical axis for different digital
displacements. (d) Quantification of the FWHM for the axial fit of spike probability as a function of digital
defocus from the focal plane ( p ¼ 0.17, Kruskal-Wallis Test). (Adapted from Pegard et al. [7])
Fig. 5 Spatial resolution with simultaneous targets throughout a large volume. (a) 3D-SHOT was tested by
simultaneously targeting 75 randomly distributed targets within the full operating volume defined by the SLM’s
spatial range for the first diffracted order. Projection views of the 3D reconstruction of 2P-induced fluorescence in
a calibration slide are shown along the (y, z), (x, z) axis, with a 3D projection. (b) Similar experiments were
repeated with 20, 30, 50, and 75 targets. The FWHMs of the two-photon response were computed for each target,
and show that spatial resolution and axial confinement are not significantly degraded by increasing the number of
simultaneous targets in any given hologram (axial FWHM: p ¼ 0.34, Kruskal-Wallis; radial FWHM: p < 0.001 for
75 ROIs; p > 0.05 for all other comparisons). (Adapted from Pegard et al. [7])
2.3 Calibration of The calibration and alignment of the optical system is critical to the
3D-SHOT with Imaging successful use of any multiphoton stimulation system, this is made
System even more challenging when improving the resolution of such a
system. Furthermore, whereas it is customary to report the best
possible resolution in optics publications (to explain the potential
of the technique), it is also known that the resolution is not con-
stant throughout the entire working volume. However, for
biological experiments, it is necessary to know what the actual
resolution is at any given point in the working volume.
High-Speed All-Optical Neural Interfaces with 3D Temporally Focused Holography 117
Calibration Procedure
(a) Manually position the substage camera such that the slide is in
focus, and the illuminated area is in the center of the substage
camera’s field of view. Tip: It may help to zoom in the 2P image
and/or reduce the line count to get more visible fluorescence on
the substage camera. However, care should be made not to
photobleach the slide, and it will be necessary to return the
imaging conditions to their normal settings before the rest of
the calibration.
(b) Compute 500–1000 test holograms containing a single target
randomly placed throughout the imaging accessible volume.
Tips: Random spots work slightly better than regularly placed
spots to avoid overfitting.
(c) Coarse Alignment. Sequentially illuminate each hologram
with the same power and record the fluorescence on the
substage camera (Fig. 6a). Move the objective mechanically
in 25-μm steps increments throughout the useable volume to
get a stack of images for each hologram. From this stack you
can get the expected XY location of each hologram, and the
peak fluorescence depth. Tips: Make sure that the power level
you use here is neither saturating the substage camera, nor too
dim to be visible. As holograms near the center of the optical axis
will be brighter (better diffraction efficiency) we recommend
118 Ian Antón Oldenburg et al.
Fig. 6 Calibration protocol for 3D-SHOT. (a) Substage camera assembly for calibration with a uniform
fluorescent thin film on a microscope slide. (b) A single hologram is imaged at 13 different planes by moving
the hologram with respect to the thin fluorescent slide. The full range is 40 μm from the estimated center of
the hologram. (c) We fit a Gaussian curve to the measured fluorescence at each plane for each hologram
recorded in (b). Relevant resolution characterization values (peak intensity, FWHM, and depth) are extracted
for each hologram. (d) We first identify the relationship between the predicted SLM defocus and the detected
depth of the corresponding holograms. (e) Mapped relationship between hologram FWHM and the hologram
depth (left) is measured in the entire volume accessible by the SLM (right). (f) Hologram diffraction efficiency is
measured throughout the field of view. (g) True depth of the two-photon imaging planes, as detected by the
substage camera. We note that the planes are neither flat, evenly spaced, nor parallel to the axis of the
camera, but that the calibration will account for all those discrepancies. (h) The final hole pattern in SLM space
accounts for aberrations and curvature from both the SLM/stimulation path and imaging path. (i) Images of the
holes ablated in the first plane, and for a subsequent plane. The hole pattern is asymmetric, so that
subsequent planes will not burn in the same location. (j) Simulated targeting error over the full calibrated
volume
High-Speed All-Optical Neural Interfaces with 3D Temporally Focused Holography 119
using a test hologram near the zero order block and set the power
such that it is just below saturating the camera pixels.
(d) Fine Alignment. Take a z stack at 5–6 μm spacing for the
40 μm around the expected Z location of each test holo-
gram (Fig. 6b). Extract the fluorescence at the center of the
hologram across the measured depths and fit this value to a
Gaussian. The width of this Gaussian is the axial FWHM for
the measured XYZ depth, and the peak of the Gaussian will be
used to determine the diffraction efficiency for this point
(Fig. 6c). The radial FWHM is determined by the fluores-
cence profile of the holographic spot at the best depth. It is
best measured by fitting the observed fluorescence to a
Gaussian, but as it is measured at a high spatial resolution
already, this is not critical. Tip: while there may be aberrations
or curvature from the SLM or other optical elements, they don’t
need to be explicitly corrected as the general mapping strategy
will account for all smooth distortions.
(e) Fit SLM locations to substage cameras. For every SLM XYZ
coordinate we now have a corresponding XYZ location
measured by the substage camera with the depth determined
from the z stack. Some holograms may need to be excluded if
they were not properly detected (see Note 2.3.3). Use a
polynomial fit to map SLMXYZ to CameraXYZ (Fig. 6d).
Tip: When fitting make sure to scale both the camera pixels and
the z depth to similar size units (such as μm) to prevent over-
weighting one axis. It is best practices to use ‘hold out’ data to
test the fidelity of your fit. This will allow you to detect if there are
systematic problems in your procedure.
(f) At this point you will have collected all the necessary infor-
mation to detect any variation in the shape of the holograms
throughout the useable volume. Axial FWHM is often
degraded when reaching the limits of your optical system
(both radially near the edges, and axially at far defocus levels).
It is important to restrict your experiments to locations where
this FWHM is acceptable for your application (Fig. 6e).
(g) To determine the diffraction efficiency as a function of target
location in the accessible volume, you may divide the
observed peak fluorescence from the fine calibration by the
best fluorescence observed in the experiment. Here again, we
recommend a polynomial interpolation to map the scalar
“diffraction efficiency as a function of the SLMXYZ coordi-
nates” (Fig. 6f). Tips: The diffraction efficiency accounts for
many spatially dependent losses of observed fluorescence beyond
holographic diffraction efficiency itself. This however works out
to the more useful measure when running biological experi-
ments. Furthermore, it would be more correct to get the best
120 Ian Antón Oldenburg et al.
targeted opsin constructs”, i.e., those that are only expressed in the
soma and proximal dendrites. Without soma targeting, the spatial
resolution is compromised [11, 19, 56]. Furthermore, other opsin
properties, such as photocurrent amplitude, absorption spectrum,
and photocurrent kinetics, will strongly affect the experimental
abilities of a 3D-SHOT system. These biophysical properties will
interact with both the imaging and 3D-SHOT system and will alter
the resolution, precision, and scale of neural control that is possible.
We will briefly summarize how these properties interact, before
describing a protocol for assessing opsin properties with regard to
how to best activate or suppress a neuron. While many techniques
are employed in this evaluation process, we will focus on those steps
that are germane to opsin evaluation and two-photon optogenetics
while directing the reader elsewhere for some technical procedures.
While we will focus on selection of activating opsins, we will briefly
discuss the additional concerns surrounding selection of a suppres-
sing opsin.
We will also include, where possible and relevant, data on
existing opsins. While many studies focus on individual features of
opsin behavior, proper evaluation of an opsin requires a holistic
understanding of many properties of those opsins. There are rela-
tively few commonly used activating opsins used with multiphoton
optogenetics the most common variants are ChR2 [28], C1V1
[8, 13], ChrimsonR, Chronos [15], ChroME [11], ReaChR
[57], CoChR [56], and ChRmine [14] each with their own advan-
tages and disadvantages. Opsins for two-photon suppression are
less well characterized with only Arch [13], NpHR3, PsuACR, iC+
+, GtACR1 [11], and GtACR2 [58] being described.
We will select for opsins that:
l Are well-trafficked to the somatic membrane with little toxicity.
l Have large photocurrents.
l Have fast kinetics.
l Are well activated by 1030-nm light.
l Are compatible with the all-optical approach of choice.
l Can reliably drive spiking in vivo.
Procedure
1. Subcellular Targeting.
When expressed in neurons, the opsin must be properly traf-
ficked to the cell membrane but restricted to whatever extent
possible away from the distal dendrites and axons (Fig. 7a). A
sequence from the Kir2.0 channel [59] can be very helpful to
export from endoplasmic reticulum, while a portion of the Kv2.1
channel [60] has become the standard (but not the only [56])
“soma targeting” motif, facilitating both membrane expression
and restriction to the soma and proximal dendrites. It is advisable
to use all subcellular targeting motifs even when testing opsins in
124 Ian Antón Oldenburg et al.
Fig. 7 Characterizing opsin characteristics for use with 3D-SHOT. (a) Comparison of non-soma-targeted opsin
localization (left) to soma-targeted opsin localization (right). (Adapted from Mardinly et al. [11]). (b) Compari-
son of photocurrent FWHM (right) at different points on the opsin response function (left). Closer to saturation
(dark blue), the actual FWHM of the photocurrent is larger than the theoretical opsin FWHM. (c) Comparison of
photocurrent amplitude and kinetics for several commonly used opsins expressed in CHO cells. (Adapted from
Sridharan et al. [61]). (d) Schematic of three common opsin kinetic metrics: left, time to peak used to measure
opening kinetic. Center, desensitization. Right, tau off, a metric of decay kinetic. (e) Schematized absorption
spectra for three opsin types compared with GCaMP absorption spectrum (dashed green). (f) Schematic of
how, with fast imaging and slow opsins (top left) scan-induced photocurrents can accumulate to produce
unwanted spiking, while in other conditions they decay and do not produce spiking. (g) Schematic displaying
how under different stimulation conditions, short laser light pulses (left) can produce more or less reliable
spike trains depending on opsin characteristics
High-Speed All-Optical Neural Interfaces with 3D Temporally Focused Holography 125
less than one millisecond [11, 61, 70]. Nevertheless, in the correct
conditions even much slower opsins such as ChrimsonR [52],
CoChR [56], or ChRmine [14] can reach jitter of about 1 ms.
However, cells expressing these last three opsins struggle to follow
rates over 20 Hz [14, 52], probably due to their slower off kinetics.
4. Two-Photon Spectra.
The two-photon absorption spectrum of an opsin affects its
compatibility with imaging approaches and suitability for use with
the high-power lasers typically used for 3D-SHOT. A chief concern
with simultaneous imaging and holographic activation is the phe-
nomenon of crosstalk, in which the scanning diffraction-limited
spot used to excite GCaMP fluorescence also activates opsin mole-
cules (the reader may also refer to Chaps. 2, 3, 5, 6, and 11 of this
book for an extended overview of this phenomenon). Opsins with
low absorption in the range of wavelengths used to image GCaMP
(typically 910–930 nm) reduce crosstalk. Unfortunately, many tra-
ditional opsins are highly activated by blue light, so this has
required the development of many new opsins. Alternate calcium
indicators that absorb in other wavelengths are also available but are
much dimmer than available GCaMPs [71–73]. In addition, most
commercially available high-energy lasers emit around 1030 nm
[63]. Thus, the optimal opsin would have a peak photocurrent
around 1030 nm with a comparative minimum at the wavelengths
to image GCaMP (~930 nm) (Fig. 7e, see Note 2.4.5 for further
discussion of alternate strategies).
It is well known that fluorophores have two-photon absorption
spectra that are considerably different from their one-photon coun-
terparts [74]. To assess the sensitivity at various 2 photon wave-
lengths, the simplest approach is to use a 2p imaging microscope
with femtosecond laser (e.g., Ti: Sapphire oscillator) that is tunable
across a large range of wavelengths. Since spectral response is not
known to be affected by the light delivery method, using a scanning
imaging system to photoactivate opsins is a suitable approach to
compare the relative activation at different wavelengths and mea-
sure the spectral response profile. Recording photocurrents from
CHO cells at different wavelengths while scanning will provide a
two-photon spectrum for the opsin. Emission power varies with
wavelength, so be sure to test power out of the objective at all
testing wavelengths.
Of the common excitatory opsins, ChR2 [28] and CoChR
[56] are blue-shifted making them suboptimal for pairing with
GCaMP imaging (Fig. 7e). Several 2p-optimized opsins, including
Chronos, ChroME [11], and ReachR [57] peak around 1000 nm,
and more red-shifted opsins such as C1V1 [13], ChrimsonR [11],
and ChRmine [14] peak beyond 1040 nm.
128 Ian Antón Oldenburg et al.
3 Conclusion
Notes
2.4.1: While observing aggregates in structural imaging is a clear
indicator of possible toxicity, many exogenous proteins can
have deleterious effects on the health of cells without necessar-
ily forming aggregates. Thus, we recommend further assess-
ment of cell health for all preparations used for experiments.
Viral overexpression and/or combination with other stressors,
including calcium buffering from GCaMP, may further lead to
cells no longer responding physiologically. It is important to
benchmark the health of cells with your chosen opsin and
expression strategy. There are two broad categories to examine:
First the intrinsic properties of cells, and second their physio-
logical responses to stimuli. Intrinsic properties, such as a cell’s
input resistance, resting membrane potential, membrane time
constant, action potential threshold, and shape of an action
potential should all remain unchanged, between opsin expres-
sing and opsin negative cells, and are easily measurable in
ex vivo whole-cell recordings. Similarly, the resting firing rate
of cells in vivo should not be altered by the presence of an
opsin, nor when imaging with standard GCaMP imaging con-
ditions. Physiological assessment of cell health is more chal-
lenging and is often overlooked. Ideally, one would measure
the in vivo responses to given sensory stimuli and show that
they do not change with presence or activation of the opsin.
2.4.2: Opsin expression. While AAV delivery of opsin is a popular
and effective strategy for expression in neurons, it can
High-Speed All-Optical Neural Interfaces with 3D Temporally Focused Holography 131
References
1. Adam Y et al (2019) Voltage imaging and opto- 6. Naka A et al (2019) Complementary networks
genetics reveal behaviour-dependent changes of cortical somatostatin interneurons enforce
in hippocampal dynamics. Nature. https:// layer specific control. elife. https://doi.org/
doi.org/10.1038/s41586-019-1166-7 10.7554/eLife.43696
2. Szabo V, Ventalon C, De Sars V, Bradley J, 7. Pégard NC et al (2017) Three-dimensional
Emiliani V (2014) Spatially selective holo- scanless holographic optogenetics with tempo-
graphic photoactivation and functional fluores- ral focusing (3D-SHOT). Nat Commun.
cence imaging in freely behaving mice with a https://doi.org/10.1038/s41467-017-
fiberscope. Neuron 84:1157–1169 01031-3
3. Carrillo-Reid L, Han S, Yang W, Akrouh A, 8. Packer AM et al (2012) Two-photon optoge-
Yuste R (2019) Controlling visually guided netics of dendritic spines and neural circuits.
behavior by holographic recalling of cortical Nat Methods 9:1202–1205
ensembles. Cell 178:447–457.e5 9. Nikolenko V, Poskanzer KE, Yuste R (2007)
4. Sohal VS, Zhang F, Yizhar O, Deisseroth K Two-photon photostimulation and imaging of
(2009) Parvalbumin neurons and gamma neural circuits. Nat Methods. https://doi.org/
rhythms enhance cortical circuit performance. 10.1038/nmeth1105
Nature 459:698–702 10. Ronzitti E, Emiliani V, Papagiakoumou E
5. Liu X et al (2012) Optogenetic stimulation of a (2018) Methods for three-dimensional all-op-
hippocampal engram activates fear memory tical manipulation of neural circuits. Front Cell
recall. Nature. https://doi.org/10.1038/ Neurosci. https://doi.org/10.3389/fncel.
nature11028 2018.00469
High-Speed All-Optical Neural Interfaces with 3D Temporally Focused Holography 133
11. Mardinly AR et al (2018) Precise multimodal 25. Svoboda K, Yasuda R (2006) Principles of
optical control of neural ensemble activity. Nat two-photon excitation microscopy and its
Neurosci. https://doi.org/10.1038/s41593- applications to neuroscience. Neuron 50:
018-0139-8 823–839
12. Lin JY, Knutsen PM, Muller A, Kleinfeld D, 26. Svoboda K, Denk W, Kleinfeld D, Tank DW
Tsien RY (2013) ReaChR: a red-shifted variant (1997) In vivo dendritic calcium dynamics in
of channelrhodopsin enables deep transcranial neocortical pyramidal neurons. Nature 385:
optogenetic excitation. Nat Neurosci 16: 161–165
1499–1508 27. Denk W, Strickler JH, Webb WW (1990)
13. Prakash R et al (2012) Two-photon optoge- Two-photon laser scanning fluorescence
netic toolbox for fast inhibition, excitation and microscopy. Science (80- ). https://doi.org/
bistable modulation. Nat Methods 9: 10.1126/science.2321027
1171–1179 28. Rickgauer JP, Tank DW (2009) Two-photon
14. Marshel JH et al (2019) Cortical layer-specific excitation of channelrhodopsin-2 at saturation.
critical dynamics triggering perception. Science Proc Natl Acad Sci U S A 106:15025–15030
(80- ). https://doi.org/10.1126/science. 29. Katona G et al (2012) Fast two-photon in vivo
aaw5202 imaging with three-dimensional random-
15. Klapoetke NC et al (2014) Independent opti- access scanning in large tissue volumes. Nat
cal excitation of distinct neural populations. Methods 9:201–208
Nat Methods 11:338–346 30. Reddy GD, Kelleher K, Fink R, Saggau P
16. Gill JV et al (2020) Precise holographic manip- (2008) Three-dimensional random access mul-
ulation of olfactory circuits reveals coding fea- tiphoton microscopy for functional imaging of
tures determining perceptual detection. neuronal activity. Nat Neurosci 11:713–720
Neuron 1–12. https://doi.org/10.1016/j. 31. Yang W et al (2016) Simultaneous multi-plane
neuron.2020.07.034 imaging of neural circuits. Neuron. https://
17. Dalgleish HWP et al (2020) How many neu- doi.org/10.1016/j.neuron.2015.12.012
rons are sufficient for perception of cortical 32. Piyawattanametha W et al (2006) Fast-
activity? elife 9:1–99 scanning two-photon fluorescence imaging
18. Carrillo-Reid L, Yang W, Bando Y, Peterka DS, based on a microelectromechanical systems
Yuste R (2016) Imprinting and recalling corti- two- dimensional scanning mirror. Opt Lett
cal ensembles. Science (80- ). https://doi.org/ 31:2018–2020
10.1126/science.aaf7560 33. Packer AM, Russell LE, Dalgleish HWP,
19. Chettih SN, Harvey CD (2019) Single-neuron H€ausser M (2014) Simultaneous all-optical
perturbations reveal feature-specific competi- manipulation and recording of neural circuit
tion in V1. Nature 567:334–340 activity with cellular resolution in vivo. Nat
20. Rickgauer JP, Deisseroth K, Tank DW (2014) Methods 12:140
Simultaneous cellular-resolution optical per- 34. Yang W, Carrillo-Reid L, Bando Y, Peterka DS,
turbation and imaging of place cell firing fields. Yuste R (2018) Simultaneous two-photon
Nat Neurosci 17:1816–1824 imaging and two-photon optogenetics of cor-
21. Daie K, Svoboda K, Druckmann S (2021) Tar- tical circuits in three dimensions. elife. https://
geted photostimulation uncovers circuit motifs doi.org/10.7554/eLife.32671
supporting short-term memory. Nat Neurosci 35. Zhang J, Pégard N, Zhong J, Adesnik H, Wal-
24:259–265 ler L (2017) 3D computer-generated hologra-
22. Clancy KB, Schnepel P, Rao AT, Feldman DE phy by non-convex optimization. Optica.
(2015) Structure of a single whisker represen- https://doi.org/10.1364/optica.4.001306
tation in layer 2 of mouse somatosensory. Cor- 36. Eybposh M, Caira N, Atisa M,
tex 35:3946–3958 Chakravarthula P, Pegard N (2020)
23. Sato TR, Gray NW, Mainen ZF, Svoboda K DeepCGH: 3D computer-generated hologra-
(2007) The functional microarchitecture of phy using deep learning. Opt Express. https://
the mouse barrel cortex. PLoS Biol 5: doi.org/10.1364/oe.399624
1440–1452 37. Gerchberg RW, Saxton WO (1972) A practical
24. Ohki K, Chung S, Ch’ng YH, Kara P, Reid RC algorithm for the determination of phase from
(2005) Functional imaging with cellular reso- image and diffraction plane pictures. Optik
lution reveals precise micro-architecture in (Stuttg) 35:237–246
visual cortex. Nature 433:597–603
134 Ian Antón Oldenburg et al.
38. Nikolenko V et al (2008) SLM microscopy: 53. Fenno L, Yizhar O, Deisseroth K (2011) The
scanless two-photon imaging and photostimu- development and application of optogenetics.
lation with spatial light modulators. Front Annu Rev Neurosci 34:389–412
Neural Circuits 2:5 54. Gerchberg RW, Saxton WO (1972) Phase
39. Papagiakoumou E et al (2010) Scanless retrieval by iterated projections. Optik (Stuttg)
two-photon excitation of channelrhodopsin-2. 35:237
Nat Methods 7:848 55. Sinclair G et al (2004) Interactive application in
40. Tal E, Oron D, Silberberg Y (2005) Improved holographic optical tweezers of a multi-plane
depth resolution in video-rate line-scanning Gerchberg-Saxton algorithm for three-
multiphoton microscopy using temporal focus- dimensional light shaping. Opt Express 12:
ing. Opt Lett 30:1686–1688 1665–1670
41. Zhu G, van Howe J, Durst M, Zipfel W, Xu C 56. Shemesh OA et al (2017) Temporally precise
(2005) Simultaneous spatial and temporal single-cell-resolution optogenetics. Nat Neu-
focusing of femtosecond pulses. Opt Express rosci 20:1796–1806
13:2153–2159 57. Chaigneau E et al (2016) Two-photon holo-
42. Spesyvtsev R, Rendall HA, Dholakia K (2015) graphic stimulation of ReaChR. Front Cell
Wide-field three-dimensional optical imaging Neurosci. https://doi.org/10.3389/fncel.
using temporal focusing for holographically 2016.00234
trapped microparticles. Opt Lett 40: 58. Forli A et al (2018) Two-photon bidirectional
4847–4850 control and imaging of neuronal excitability
43. Durst ME, Zhu G, Xu C (2006) Simultaneous with high spatial resolution in vivo. Cell Rep.
spatial and temporal focusing for axial scan- https://doi.org/10.1016/j.celrep.2018.
ning. Opt Express 14:12243–12254 02.063
44. Mayblum T, Schejter A, Dana H, Shoham S 59. Stockklausner C, Ludwig J, Ruppersberg JP,
et al (2015) SPIE BiOS 932928 Klöcker N (2001) A sequence motif responsi-
45. Prevedel R et al (2016) Fast volumetric calcium ble for ER export and surface expression of
imaging across multiple cortical layers using Kir2.0 inward rectifier K+ channels. FEBS
sculpted light. Nat Methods 13(12): Lett. https://doi.org/10.1016/S0014-5793
1021–1028 (01)02286-4
46. Andrasfalvy BK, Zemelman BV, Tang J, Vaziri 60. Lim ST, Antonucci DE, Scannevin RH, Trim-
A (2010) Two-photon single-cell optogenetic mer JS (2000) A novel targeting signal for
control of neuronal activity by sculpted light. proximal clustering of the Kv2.1 K+ channel
Proc Natl Acad Sci U S A 107:11981–11986 in hippocampal neurons. Neuron 25:385–397
47. McCabe DJ et al (2011) Spatio-temporal 61. Sridharan S et al (2022) High-performance
focusing of an ultrafast pulse through a multi- microbial opsins for spatially and temporally
ply scattering medium. Nat Commun 2:447 precise perturbations of large neuronal net-
48. Dana H, Shoham S (2011) Numerical evalua- works. Neuron 110(7):1139–1155.e6
tion of temporal focusing characteristics in 62. Kishi KE et al (2022) Structural basis for chan-
transparent and scattering media. Opt Express nel conduction in the pump-like channelrho-
19:4937–4948 dopsin ChRmine. Cell 185:672–689.e23
49. Therrien OD, Aubé B, Pagès S, De Koninck P, 63. Ronzitti E et al (2017) Submillisecond opto-
Côté D (2011) Wide-field multiphoton imag- genetic control of neuronal firing with
ing of cellular dynamics in thick tissue by tem- two-photon holographic photoactivation of
poral focusing and patterned illumination. chronos. J Neurosci. https://doi.org/10.
Biomed Opt Express 2:696–704 1523/JNEUROSCI.1246-17.2017
50. Bègue A et al (2013) Two-photon excitation in 64. Schneider F, Grimm C, Hegemann P (2015)
scattering media by spatiotemporally shaped Biophysics of channelrhodopsin. Annu Rev
beams and their application in optogenetic Biophys. https://doi.org/10.1146/annurev-
stimulation. Biomed Opt Express 4: biophys-060414-034014
2869–2879 65. Kuhne J et al (2019) Unifying photocycle
51. Hernandez O et al (2016) Three-dimensional model for light adaptation and temporal evolu-
spatiotemporal focusing of holographic pat- tion of cation conductance in
terns. Nat Commun 7:1–10 channelrhodopsin-2. Proc Natl Acad Sci U S
52. Chen IW et al (2019) In vivo submillisecond A . h t t p s : // d o i . o r g / 1 0 . 1 0 7 3 / p n a s .
two-photon optogenetics with temporally 1818707116
focused patterned light. J Neurosci. https:// 66. Berndt A et al (2011) High-efficiency chan-
doi.org/10.1523/JNEUROSCI.1785-18. nelrhodopsins for fast neuronal stimulation at
2018 low light levels. Proc Natl Acad Sci U S
High-Speed All-Optical Neural Interfaces with 3D Temporally Focused Holography 135
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International
License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing, adaptation, distribution
and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative Commons license,
unless indicated otherwise in a credit line to the material. If material is not included in the chapter’s Creative
Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use,
you will need to obtain permission directly from the copyright holder.
Chapter 5
Abstract
All-optical physiology of neuronal microcircuits requires the integration of optogenetic perturbation and
optical imaging, efficient opsin and indicator co-expression, and tailored illumination schemes. It further-
more demands concepts for system integration and a dedicated analysis pipeline for calcium transients in an
event-related manner. Here, firstly, we put forward a framework for the specific requirements for technical
system integration particularly focusing on temporal precision. Secondly, we devise a step-by-step guide for
the image analysis in the context of an all-optical physiology experiment. Starting with the raw image, we
present concepts for artifact avoidance, the extraction of fluorescence intensity traces on single-neuron
basis, the identification and binarization of putatively action-potential-related calcium transients, and finally
ensemble activity analysis.
Key words All-optical physiology, Functional calcium imaging, System integration, Optogenetics,
Event-related analysis of functional fluorescence traces, Stimulation artifact avoidance, Spectral
independency
1 Introduction
Eirini Papagiakoumou (ed.), All-Optical Methods to Study Neuronal Function, Neuromethods, vol. 191,
https://doi.org/10.1007/978-1-0716-2764-8_5, © The Author(s) 2023
137
138 Hendrik Backhaus et al.
first years upon the advent of optogenetics [3]. Yet, there is one
field of neuroscience, which maybe poses the ultimate challenge for
the integration of optogenetics: optical functional imaging. The
field of optical imaging, particularly of individual neuron function
in the intact tissue, relies on detecting subtle changes in light
intensity [4]. The implementation of two-photon microscopy in
combination with calcium indicators in the early 2000 allowed for
the first time a functional readout of neuronal circuits comprising of
several hundred neurons in the rodent cortex [4–6]. This leads to a
tremendous advance in our understanding on how complex circuit
dysfunction arises, particularly also in the early stages of neurologi-
cal disorders [7–11]. The ability to identify a rather small fraction of
dysregulated neurons in the circuitry makes the true difference in
comparison to single-cell or population readouts. And of course,
also here, the implementation of optogenetics holds tremendous
potential, as a causal manipulation of, e.g., single dysregulated
neurons and the simultaneous readout of circuit function would
truly advance our understanding on cause and effect in network
disorders. And indeed, in 2012 [12, 13] the efficient two-photon
excitation of opsins became a reality, followed by combined
two-photon imaging and two-photon optogenetics [14–17]. In
these early proof of concept studies, it already became apparent,
that all-optical approaches pose unique challenges, both in terms of
system integration, experimental design, and not the least, analysis.
First, conventional line-scanning excitation schemes have proven to
be rather less efficient. Second, only a few opsins seem to be
two-photon excitable, and there is yet no molecular or structural
predictor, which could guide molecular engineering of opsins to
modify their two-photon cross section. Thirdly, as mentioned
above, the light intensity used for two-photon optogenetics ranges
orders of magnitude higher compared to the light intensities used
for the excitation of fluorophores, creating problems in terms of
indicator bleaching and tissue heating. Lastly, also the analysis poses
problems in terms of artifact removal and synchronization. All of
this prevented up to now the broad implementation of all-optical
approaches, despite their promises. Here we will focus on the entire
workflow for an all-optical experiment in circuit neuroscience,
reporting on recent advances, and giving guidance for the unique
requirements of all-optical physiology.
2 Experimental Framework
B
One-photon Gradient-index
LED & Gradient-index Full-field
Miniature lens &
Bandpass filter lens excitation
Microscope CCD Chip
Fig. 1 Principles of functional calcium imaging. (a) Scheme of two-photon-based readout using a femtosecond
laser and raster scanning principle. The laser is focused by the objective and the focal point is guided by a
galvo-resonant scanhead over the field of view. Emitted fluorescence is collected and detected by a PMT. (b)
One-photon Miniature Microscopes used for freely moving animals require a miniaturization of light sources,
achieved by bandpass filtered LEDs. A full-field illumination via a GRIN lens is carried out and a CCD-Chip
collects emitted fluorescence
142 Hendrik Backhaus et al.
4.2 One-photon In the field of in vivo calcium imaging, a recently emerging method
Miniature Microscope are miniature microscopes, mounted on an implanted baseplate on
Full-Field Imaging via the head of the animal. In this chapter, we solely refer to
GRIN Lens one-photon miniature microscopes, whereas Chap. 7 gives a
detailed introduction to two-photon miniature microscopes.
Recently, the first three-photon miniature microscope was pro-
posed by the group of Jason Kerr [45]. Miniature microscopes
paved the way for the combination of functional imaging with
behavioral assays in which the animal moves freely. What is more,
brain regions not accessible by two-photon microscopy due to its
limited penetration depth can be targeted. First developed by the
group of Mark Schnitzer [46], nowadays several manufacturers
offer ready-to-use systems (INSCOPIX, Palo Alto, CA, USA
[47]) but also open-source systems are being provided by the
community, e.g., the UCLA Miniscope [48] or the
FinchScope [49].
Retaining the image information also in deeper brain regions is
afforded by a gradient-index optic (GRIN lens, Fig. 1b). GRIN
lenses have a cylindrical shape with a radially decreasing refraction
index, and planar surfaces for optimizing optic interfacing. These
two factors, a gradient of refractory index and the planar surface,
retain, at least to some extent, the image information throughout
the passage of light through the GRIN lens. It is important to note
that there is a deterioration of image quality with increasing the
length of the GRIN lens. As a light source, a LED in combination
with spectral bandpass filters is used [47], and the image is typically
recorded by a CCD chip. The sampling rates for functional calcium
imaging differ depending on the chosen model, ranging from
15 Hz with a field of view of 1440 by 1080 pixels (nVoke, Inscopix,
Palo Alto, California) up to 60 Hz at a field of view of 752 by
480 pixels (Miniscope-v4, UCLA, Los Angeles, California). The
following issue needs to be considered: the reliable recording of
neuronal layers depends on the location of the GRIN lenses tip,
neurons located 100–300 μm below the tip can be resolved by
electronically adjusting a focusing lens [47]. However, changing
the x-y-position of the recorded field of view is not possible and is
predefined by the implantation site of the GRIN lens.
All-Optical Physiology Pipeline 143
B
One-photon Gradient-index
Full-field
Miniature lens
excitation
Microscope (one-photon)
Fig. 2 Schematic overview of system combinations for opsin excitation. (a) Two-photon functional calcium
imaging can be combined with one-photon or two-photon raster scan methods to target opsin-expressing
neurons. The method of Computer Generated Holography is a scanless approach and enables the experi-
menter to excite several neurons at the same time under a two-photon regime using spatial light modulators
(SLMs). (b) Miniature microscopes equipped with a second LED for opsin excitation are based on one-photon
full-field illumination delivered via an implanted GRIN lens
Fig. 3 Mechanisms for opsin excitation. (a) One-photon excitation by a focused beam is carried out by raster
scanning the target neurons. Neurons located above or below the imaging plane are exposed to the excitation
beam cone as well. (b) One-photon full-field paradigms are used in miniature microscopes and do not allow
for neuron-specific excitation but illuminate the entire FOV at once, eventually exciting neurons below the
recorded layers. (c) Two-photon raster scanning approaches overcome this disadvantage due to the physical
principle of two-photon excitation, reaching the necessary photon density only in the focal spot. (d) The
scanless paradigm also enables experimenters to simultaneously excite several target neurons under a
two-photon regime
5.3 Two-photon As described in Subsection 5.2, if more than two neurons need to
Scanless Opsin be efficiently activated within the time window of a typical imaging
Excitation via CGH frame, a two-photon raster scanning approach is reaching its limits.
Parallel excitation methods in principle allow for an infinite number
of simultaneously activated regions of interest (ROIs), limited only
by the laser power and the spatial resolution afforded by the spatial
light modulator used [59]. Parallel methods most commonly use
Computer Generated Holography (CGH) for the generation of the
excitation pattern [60] (Fig. 2a). In CGH, liquid crystal spatial
light modulators (LC-SLMs) are integrated into the beam path to
modulate the phase of the electric field of the laser beam, and
typically a low-repetition rate, high-energy Ytterbium laser is used
as a light source [57]. A precise control of the spatial specificity of
the calculated phase hologram is achieved by applying a Fourier
transform-based iterative algorithm on previously acquired fluores-
cence images of the region of interest [61]. A detailed description
of the technical aspects of this method can be found in Chaps. 4
and 11.
With this approach, an effective, simultaneous excitation of
several neurons is amenable. The illumination targets do not need
to be located in the same z plane, neurons above or below the
imaging plane can be interrogated as well, provided that the loca-
tion of all targets is known (Fig. 3d). Depending on the respective
microscope and illumination concept, neurons can be excited
which are located several hundreds of μm distant from the current
z plane. Yet, upgrading an existing microscope setup with this
technique can be cumbersome and expensive, as major changes in
software and hardware need to be done. However, it might be an
interesting option for scientists experienced in two-photon micros-
copy to upgrade a microscope with this technology. A further
interesting option represent hybrid solutions, combining CGH
and scanning [12]. These hybrid solutions guide the illumination
patterns generated by the SLM on a galvanometric mirror. These
patterns comprise multiple typically rather small focal points, but
with high light intensity. This cloud of focal points is then scanned
by the galvanometric mirror, resulting in each focal point to cover
an entire given neuron either using line or spiral scanning
approaches. Thereby, a limitation of CGH-only approaches in
terms of the ever-decreasing light intensity when increasing the
number of neuron-sized patterns is circumvented. For a constant
laser power, the number of neurons which can be effectively acti-
vated by a two-photon regime, as discussed in the previous section,
and in [12] is consequently higher with hybrid solutions (see also
Chap. 11).
148 Hendrik Backhaus et al.
Munk model. Note that the lateral spread close to the tip of the
fiber may be underestimated by the Kubelka-Munk model [54, 62],
at least for highly scattering shorter wavelengths.
6 Everything You Always Wanted to Know About All-Optical Data Processing But
Were Afraid to Ask
6.1 Roadmap for Prior to image data acquisition, the synchronicity of all necessary
Processing All-Optical subsystems needs to be guaranteed (Fig. 4a). This can be achieved
Data by copying all relevant trigger and stimulation pulses to a multidata
acquisition interface. These signals include the frame trigger of the
6.1.1 System Integration microscope, the logic level controlling the light source for opsin
excitation, and biomonitoring such as the breathing rate and body
temperature. Upon digitalization, the signals are being centrally
displayed by a control software, and a master file for each experi-
ment is generated. Reading out the master files allows for a post hoc
exact assignment of each raw image to a given time and the status
of, e.g., the stimulation regime.
6.1.2 Image Data The initial step for analyzing the data is to reduce the impact of
Acquisition movement artifacts (Fig. 4b). We differentiate between two types of
artifacts: displacements induced by movement in the x-y plane, and
in z plane. While changes in the x-y plane can be corrected retro-
spectively by algorithms based on Hidden-Markov-Model [65] or
discrete Fourier transformation-based image alignment [66],
provided that the field of view is sufficiently large, movements in
the z plane are more complicated to correct. Note that the possibil-
ity for x-y movement correction represents the main reason why
random-access scanning is not recommended, and the field is
150 Hendrik Backhaus et al.
Fig. 4 Workflow of a pipeline for an event-related analysis of all-optical data. Raw data is segmented into ROIs
and intensity traces are extracted. A binarization of transients is carried out and a temporal and spatial
classification of the recorded microcircuit is applied
6.1.3 Segmentation The raw images comprising typically 512 512 pixels contain
several biological compartments: neuronal somata, axons, den-
drites, blood vessels, and certainly a multitude of non-neuronal
cells, such as astrocytes. While the one critical advantage of micros-
copy using fluorescent indicators represents the reduction of image
complexity, i.e., ideally only the features of interest express the
fluorophore, still, only a fraction of the image contains the signal
of interest. What is more, in the context of neuronal microcircuit
imaging, it is of advantage to integrate several pixels which reflect
the same functional compartment: when the experimenter aims for
identification of the suprathreshold activity of individual neurons in
the microcircuit, each neuron can be identified as functional unit.
Consequently, each neuronal soma is defined and segmented as
ROI (Fig. 4c). In recent years, the range of applications that sup-
port the experimenter in ROI segmentation grew rapidly: mathe-
matical models to perform automatic segmentation based on deep
learning algorithms drastically shorten the time-consuming step to
manually identify neurons in the recorded images [67]. While there
are deep learning-based methods that process the average of all
images, other techniques integrate the temporal information of
neuronal activity by using subsets of all images for the segmentation
of active ROIs [67, 68]. The approach proposed by Soltanian-
Zadeh et al. is based on a 3D technique: subsets of the acquired
images are created and used to predict 2D probability maps for
active neurons utilizing a neural network. Upon applying a thresh-
old to exclude low-probability regions from each probability map,
individual somata are extracted from high-probability regions
[67]. Ultimately, all somata positions from each image from the
image sequence are combined to acquire a final output of active
somata areas. However, manual segmentation of functional calcium
imaging data by marking neuronal somata with polygon shapes still
prevails due to the applicability on datasets of varying quality, e.g.,
datasets of a particularly low signal-to-noise ratio, when deep
learning algorithms reach their limits. Depending on the strength
of GCaMP6 expression and the overall SNR ratio, if targeting
somatic changes in calcium concentrations, a decontamination of
the ROIs containing the neuronal soma from neuropil signal might
be useful. The decontamination is carried out by expanding a given
ROI in cardinal and diagonal directions, beyond the region of the
soma, which are then separated into several neuropil regions sur-
rounding the initial ROI. The size of each neuropil subregion
should equal the size of the initial ROI. Under the assumption
that the intensity trace of a given ROI is generated by a mixture of
different underlying signal components, but certainly one of the
components exhibits the somatic signal of interest, non-negative
matrix factorization or independent component analysis are
employed to perform a blind source separation, resulting in a
separation into the underlying signal components. By weighting
152 Hendrik Backhaus et al.
6.1.4 Trace Extraction After the locations of somata constituting a ROI have been identi-
fied within the image sequence, the signal over time of each ROI is
extracted (Fig. 4d). The intensity values of all pixels in a given ROI
are averaged for every image of the temporal sequence, resulting in
an intensity trace for every ROI, with a temporal resolution deter-
mined by the frame rate. It has to be noted that the absolute level of
intensity is based on multiple factors such as autofluorescence, the
expression levels of the GECI, or stray light entering the objective.
Therefore, it is recommended to calculate the relative change of
fluorescence, as these dynamic changes, depending on their tem-
poral dynamics, are most likely due to changes in intracellular
calcium levels, and can therefore be termed calcium transients.
Note that there is inevitably a drift of the baseline levels due to
bleaching. These drifts can be compensated for, as long as linear
operations are being used (see Note 4). There is no convention on
how to perform baseline correction. An option used by us and
others represents the definition of a sufficiently long period of
quiescence, i.e., stable baseline non-interrupted by any transients,
and defines this period as baseline (F0), separately for each neuron
[7, 9]. The relative change of fluorescence is then calculated by
relating the intensity of each time point (F) to the baseline fluores-
cence ΔF ¼ F FF 0 [70].
0
6.1.5 Artifact Removal Prior to binarizing the extracted intensity traces, potential photo-
stimulation artifacts superimposing the signal need to be identified
and, if possible, corrected (Fig. 4e). Characteristic properties of
photostimulation artifacts can serve as a basis for reliable identifica-
tion and subsequent correction. Firstly, looking only at individual
responses to an optogenetic stimulus might lead to the notion of a
physiological signal. Yet maybe the most decisive difference
between an artifact and a physiological response is the inherent
variability of the physiological signal. Artifacts, with rare excep-
tions, are rather consistent. Overlaying and averaging the individual
responses therefore gives important cues on the probability of a
physiological origin. For that, the temporal section of a trace
upon the photostimulation is assessed by a nm matrix M, with
n ¼ stimulus intervall, m ¼ total frames/n. Figure 5b shows an
overlay of the intensity traces of all photostimulation periods of
Fig. 5a (gray lines).
Photostimulation artifacts exhibit several typical features:
Firstly, a photostimulation artifact will show both a sharp onset,
and a sharp offset. Functional calcium transients of physiological
All-Optical Physiology Pipeline 153
Fig. 5 Evaluation of photostimulation artifacts. (a) Artifacts caused by photostimulation do not represent a
physiological signal. Onsets of photostimulation are indicated by black triangles. (b) By overlaying (gray) and
averaging (red) periods of photostimulation, an estimation on the impact of the artifact can be made. Here,
100 sample points prior to and 200 sample points after the photostimulation onset, indicated by the black
triangle, are considered. The averaged signal waveform can be used in an algorithm to minimize the
photostimulation artifact. (c) Subtracting the averaged waveform depicted in (b) at each stimulation onset
can reduce the intensity of the photostimulation artifact. Note that signal components containing the
physiological signal of interest can be altered by the algorithm as well and, in the worst-case scenario, will
be completely eliminated
origin might also display a rather sharp onset, due to the high
affinity of calcium to the indicator represented by a small dissocia-
tion constant KD and the high temporal gradient of calcium influx.
But it will be characterized by a slow decay, an inevitable conse-
quence of the high affinity of the indicator to calcium. The off
kinetics does not mirror the true time course of the decrease of
somatic calcium concentration. If a researcher would be interested
in assessing the re-uptake of calcium, an indicator with a high KD/
low affinity would be advantageous, but these indicators typically
display a lower SNR [71]. Any trace deflection with a sharp decay
when using a low KD indicator is therefore almost certainly not
associated with an AP-related somatic calcium response. Secondly,
if the duration of the intensity deflection equals the duration of
light administration, a physiological origin is highly unlikely. As
depicted in Fig. 5b, the intensity traces during photostimulation
exhibit both of the mentioned features.
What is more, but this very much depends on the imaging
setup used, the latency between the onset of the stimulation pulse
and the putative response is critical: If the latency ranges <3 ms, it
cannot be considered as a physiological response, due to the time
needed for the AP initiation and the influx of calcium into the
cytosol. Please note that this criterion can only be used if the
sampling rate of the GCaMP emission channel is high enough to
resolve durations less than 3 ms. Lastly, a physiological response
might be subject to adaptation, and often times does not occur
upon every stimulus, e.g., due to changes in local inhibition
[54]. Consequently, while a signal deflection might surely be the
result of a physiological response if it occurs upon each stimulus,
nonetheless, if a response is drastically changing its amplitude, or
sometimes is not present, it is likely to be of physiological origin. To
154 Hendrik Backhaus et al.
test for that, it might be useful to modulate the light intensity of the
optogenetic stimulus: if a decrease of stimulation leads to the
sudden disappearance of a transient, it is again likely that an
AP-related origin can be assumed, as a technical artifact would
simply scale with the light intensity, even though also non-linear
correlations can occur. Certainly, a control experiment comprising
of indicator-only expressing cells should be conducted in any case.
The subsequent correction of photostimulation artifacts in the
intensity trace is a delicate task that needs to be conducted carefully.
A possible approach is to employ non-negative matrix factorization
[72] to identify the background noise of a given ROI containing
the stimulation artifact. In a second step, the identified noise com-
ponent is subtracted from the ROIs signal [73]. Here, for illustra-
tion purposes we obtained the averaged intensity trace of all time
periods of photostimulation (Fig. 5b) and subtracted the given
value from the raw intensity trace (Fig. 5a). This results in an
intensity trace devoid of at least the majority of the
non-physiological signal sources (Fig. 5c). Please note that the
processed intensity trace has to be treated with caution: we can
still observe fluctuations in the signal that could be misinterpreted
as functional calcium transients. Furthermore, using methods
based on complex mathematical models often act as a black-box,
making it difficult to grasp the underlying methods for
non-experts. We strongly emphasize to design the experimental
paradigm in a manner to avoid any stimulation artifact during the
measurement and to minimize the need for post-processing of raw
data as much as possible. Characteristic temporal dynamics and
latencies of the used GECI, as put forward before, can serve as a
sanity check, and, following up with an event-based identification
of AP-related calcium transients is mandatory.
6.2 Toward Cross- While the strong light intensities needed for an effective excitation
Talk-Free of an opsin pose the aforementioned problems for simultaneous
Experimental Designs functional calcium imaging, we need to also consider a putative
impact on the constant excitation of the calcium indicator in terms
6.2.1 Assessing the of unwanted activation of opsins (see also Chap. 2). The activation
Impact of Continuous of an opsin requires a distinct quantal energy; in case of the ChR2-
Illumination for Calcium
based opsins or the cis-trans isomerization of retinal, this threshold
Imaging on Opsin ranges from 0.1 to 1 mW mm2. All-optical one-photon experi-
Excitation ments, even using the same wavelength for imaging and opsin
activation, is therefore possible, as long as the excitation intensity
ranges below this threshold [32, 47, 54]. However, for two-photon
line scanning conditions, the effective light intensity per pixel can
well exceed this barrier. Nevertheless, fortunately, there is also a
temporal barrier. It so seems that any pixel dwell times below 3.2 μs
may not suffice for efficient opsin excitation at least for opsins such
as C1V1 [12]. Pixel dwell times of regular resonant scanning range
well below that number. However, novel generations of opsins
designed for more efficient two-photon excitation might allow
shorter pixel dwell times.
6.2.2 Increasing the Avoiding cross-talk in the GECI emission channel is an important
Spectral Separation criterion in the design of all-optical experiments. Both for
Between Opsin and one-photon and two-photon regimes, the excitation light can be
Indicator to Minimize rather easily prevented from entering the PMTs, e.g., by notch
Optogenetic Stimulation filters. But nonetheless, the strong light pulse might lead to a
Artifacts on the Imaging broadband increase in autofluorescence, also in the emission band
Data of the fluorescence calcium indicator, leading to a decrease in SNR.
All-Optical Physiology Pipeline 157
7 Outlook
8 Notes
References
1. Deisseroth K (2015) Optogenetics: 10 years of 10. Iaccarino HF et al (2016) Gamma frequency
microbial opsins in neuroscience. Nat Neurosci entrainment attenuates amyloid load and
18:1213–1225 modifies microglia. Nature 540:230–235
2. Boyden ES, Zhang F, Bamberg E, Nagel G, 11. Rosales Jubal E et al (2021) Acitretin reverses
Deisseroth K (2005) Millisecond-timescale, early functional network degradation in a
genetically targeted optical control of neural mouse model of familial Alzheimer’s disease.
activity. Nat Neurosci 8:1263–1268 Sci Rep 11(1):6649. https://doi.org/10.
3. Dufour S, De Koninck Y (2015) Optrodes for 1038/s41598-021-85912-0
combined optogenetics and electrophysiology 12. Packer AM et al (2012) Two-photon optoge-
in live animals. Neurophotonics 2:031205 netics of dendritic spines and neural circuits.
4. Grienberger C, Konnerth A (2012) Imaging Nat Methods 9:1202–1205
calcium in neurons. Neuron 73:862–885 13. Prakash R et al (2012) Two-photon optoge-
5. Stosiek C, Garaschuk O, Holthoff K, Konnerth netic toolbox for fast inhibition, excitation and
A (2003) In vivo two-photon calcium imaging bistable modulation. Nat Methods 9:
of neuronal networks. Proc Natl Acad Sci U S A 1171–1179
100:7319–7324 14. Rickgauer JP, Deisseroth K, Tank DW (2014)
6. Helmchen F, Konnerth A (2011) In: Yuste R Simultaneous cellular-resolution optical per-
(ed) Imaging in neuroscience: a laboratory turbation and imaging of place cell firing fields.
manual, Imaging series. Cold Spring Harbor Nat Neurosci 17:1816–1824
Laboratory Press 15. Packer AM, Russell LE, Dalgleish HW, Haus-
7. Arnoux I et al (2018) Metformin reverses early ser M (2015) Simultaneous all-optical manipu-
cortical network dysfunction and behavior lation and recording of neural circuit activity
changes in Huntington’s disease. elife 7: with cellular resolution in vivo. Nat Methods
e38744 12:140–146
8. Busche MA et al (2008) Clusters of hyperactive 16. Emiliani V, Cohen AE, Deisseroth K, Hausser
neurons near amyloid plaques in a mouse M (2015) All-optical interrogation of neural
model of Alzheimer’s disease. Science 321: circuits. J Neurosci 35:13917–13926
1686–1689 17. Carrillo-Reid L, Yang W, Bando Y, Peterka DS,
9. Ellwardt E et al (2018) Maladaptive cortical Yuste R (2016) Imprinting and recalling corti-
hyperactivity upon recovery from experimental cal ensembles. Science 353:691–694
autoimmune encephalomyelitis. Nat Neurosci
21:1392–1403
All-Optical Physiology Pipeline 161
18. Kim DH et al (2017) Pan-neuronal calcium 34. Fu T et al (2021) Exploring two-photon opto-
imaging with cellular resolution in freely swim- genetics beyond 1100 nm for specific and
ming zebrafish. Nat Methods 14:1107–1114 effective all-optical physiology. iScience 24:
19. Zagha E, Casale AE, Sachdev RN, McGinley 102184
MJ, McCormick DA (2013) Motor cortex 35. Park D, Dunlap K (1998) Dynamic regulation
feedback influences sensory processing by of calcium influx by G-proteins, action poten-
modulating network state. Neuron 79: tial waveform, and neuronal firing frequency. J
567–578 Neurosci 18:6757–6766
20. Crochet S, Petersen CC (2006) Correlating 36. Bito H (1998) The role of calcium in activity-
whisker behavior with membrane potential in dependent neuronal gene regulation. Cell Cal-
barrel cortex of awake mice. Nat Neurosci 9: cium 23:143–150
608–610 37. Berridge MJ (1998) Neuronal calcium signal-
21. Petersen CC, Crochet S (2013) Synaptic com- ing. Neuron 21:13–26
putation and sensory processing in neocortical 38. Palmer AE, Tsien RY (2006) Measuring cal-
layer 2/3. Neuron 78:28–48 cium signaling using genetically targetable
22. Francis NA et al (2018) Small networks encode fluorescent indicators. Nat Protoc 1:
decision-making in primary auditory cortex. 1057–1065
Neuron 97:885–897 e886 39. Helmchen F, Denk W (2005) Deep tissue
23. Gire DH, Whitesell JD, Doucette W, Restrepo two-photon microscopy. Nat Methods 2:
D (2013) Information for decision-making and 932–940
stimulus identification is multiplexed in sensory 40. Svoboda K, Denk W, Kleinfeld D, Tank DW
cortex. Nat Neurosci 16:991–993 (1997) In vivo dendritic calcium dynamics in
24. Chen TW, Li N, Daie K, Svoboda K (2017) A neocortical pyramidal neurons. Nature 385:
map of anticipatory activity in mouse motor 161–165
cortex. Neuron 94:866–879 e864 41. Song L, Hennink EJ, Young IT, Tanke HJ
25. Redinbaugh MJ et al (2020) Thalamus modu- (1995) Photobleaching kinetics of fluorescein
lates consciousness via layer-specific control of in quantitative fluorescence microscopy. Bio-
cortex. Neuron 106:66–75 e12 phys J 68:2588–2600
26. van Vugt B et al (2018) The threshold for 42. Denk W, Strickler JH, Webb WW (1990)
conscious report: signal loss and response bias Two-photon laser scanning fluorescence
in visual and frontal cortex. Science 360: microscopy. Science 248:73–76
537–542 43. Moneron G, Hell SW (2009) Two-photon
27. Chen TW et al (2013) Ultrasensitive fluores- excitation STED microscopy. Opt Express 17:
cent proteins for imaging neuronal activity. 14567–14573
Nature 499:295–300 44. H€anninen PE, Soini E, Hell SW (1994) Con-
28. Ghanbari L et al (2019) Cortex-wide neural tinuous wave excitation two-photon fluores-
interfacing via transparent polymer skulls. Nat cence microscopy. J Microsc 176:222–225
Commun 10:1500 45. Klioutchnikov A et al (2020) Three-photon
29. Ghanbari L et al (2019) Craniobot: a computer head-mounted microscope for imaging deep
numerical controlled robot for cranial micro- cortical layers in freely moving rats. Nat Meth-
surgeries. Sci Rep 9:1023 ods 17:509–513
30. Boiroux D, Oke Y, Miwakeichi F, Oku Y 46. Ghosh KK et al (2011) Miniaturized integra-
(2014) Pixel timing correction in time-lapsed tion of a fluorescence microscope. Nat Meth-
calcium imaging using point scanning micros- ods 8:871–878
copy. J Neurosci Methods 237:60–68 47. Stamatakis AM et al (2018) Simultaneous
31. Chen X et al (2012) LOTOS-based two-pho- optogenetics and cellular resolution calcium
ton calcium imaging of dendritic spines in vivo. imaging during active behavior using a minia-
Nat Protoc 7:1818–1829 turized microscope. Front Neurosci 12:496
32. Adelsberger H, Grienberger C, Stroh A, Kon- 48. Cai DJ et al (2016) A shared neural ensemble
nerth A (2014) In vivo calcium recordings and links distinct contextual memories encoded
channelrhodopsin-2 activation through an close in time. Nature 534:115–118
optical fiber. Cold Spring Harb Protoc 2014: 49. Liberti WA 3rd et al (2016) Unstable neurons
pdbprot084145 underlie a stable learned behavior. Nat Neu-
33. Döring J, Fu T, Arnoux I, Stroh A (2018) rosci 19:1665–1671
Optogenetics: a roadmap. Springer Protocols
162 Hendrik Backhaus et al.
50. Fois C, Prouvot PH, Stroh A (2014) A road- neural activity with cellular resolution in awake,
map to applying optogenetics in neuroscience. mobile mice. Neuron 56:43–57
Methods Mol Biol 1148:129–147 66. Kaifosh P, Zaremba JD, Danielson NB,
51. Yang JW et al (2017) Optogenetic modulation Losonczy A (2014) SIMA: python software
of a minor fraction of parvalbumin-positive for analysis of dynamic fluorescence imaging
interneurons specifically affects spatiotemporal data. Front Neuroinform 8:80
dynamics of spontaneous and sensory-evoked 67. Soltanian-Zadeh S, Sahingur K, Blau S,
activity in mouse somatosensory cortex in vivo. Gong Y, Farsiu S (2019) Fast and robust active
Cereb Cortex 27:5784–5803 neuron segmentation in two-photon calcium
52. Gradinaru V et al (2010) Molecular and cellular imaging using spatiotemporal deep learning.
approaches for diversifying and extending Proc Natl Acad Sci U S A 116:8554–8563
optogenetics. Cell 141:154–165 68. Klibisz A, Rose D, Eicholtz M, Blundon J,
53. Nagel G et al (2003) Channelrhodopsin-2, a Zakharenko S (2017) Deep learning in medical
directly light-gated cation-selective membrane image analysis and multimodal learning for
channel. Proc Natl Acad Sci U S A 100: clinical decision support, vol 10553.
13940–13945 Springer, Cham
54. Stroh A et al (2013) Making waves: initiation 69. Keemink SW et al (2018) FISSA: a neuropil
and propagation of corticothalamic Ca2+ decontamination toolbox for calcium imaging
waves in vivo. Neuron 77:1136–1150 signals. Sci Rep 8:3493
55. Aravanis AM et al (2007) An optical neural 70. Hendel T et al (2008) Fluorescence changes of
interface: in vivo control of rodent motor cor- genetic calcium indicators and OGB-1 corre-
tex with integrated fiberoptic and optogenetic lated with neural activity and calcium in vivo
technology. J Neural Eng 4:S143–S156 and in vitro. J Neurosci 28:7399–7411
56. Kong C et al (2017) Compact fs ytterbium 71. Paredes RM, Etzler JC, Watts LT, Zheng W,
fiber laser at 1010 nm for biomedical applica- Lechleiter JD (2008) Chemical calcium indica-
tions. Biomed Opt Express 8:4921–4932 tors. Methods 46:143–151
57. Chaigneau E et al (2016) Two-photon holo- 72. Pnevmatikakis EA et al (2016) Simultaneous
graphic stimulation of ReaChR. Front Cell denoising, deconvolution, and demixing of cal-
Neurosci 10:234 cium imaging data. Neuron 89:285–299
58. Forli A et al (2018) Two-photon bidirectional 73. Yang W, Carrillo-Reid L, Bando Y, Peterka DS,
control and imaging of neuronal excitability Yuste R (2018) Simultaneous two-photon
with high spatial resolution in vivo. Cell Rep imaging and two-photon optogenetics of cor-
22:3087–3098 tical circuits in three dimensions. elife 7:
59. Ronzitti E et al (2017) Recent advances in e32671
patterned photostimulation for optogenetics. 74. Mank M, Griesbeck O (2008) Genetically
J Opt 19:113001 encoded calcium indicators. Chem Rev 108:
60. Haist T, Schönleber M, Tiziani HJ (1997) 1550–1564
Computer-generated holograms from 75. Friedrich J, Zhou P, Paninski L (2017) Fast
3D-objects written on twisted-nematic liquid online deconvolution of calcium imaging
crystal displays. Opt Commun 140:299–308 data. PLoS Comput Biol 13:e1005423
61. Lutz C et al (2008) Holographic photolysis of 76. Kerr JN, Greenberg D, Helmchen F (2005)
caged neurotransmitters. Nat Methods 5: Imaging input and output of neocortical net-
821–827 works in vivo. Proc Natl Acad Sci U S A 102:
62. Schmid F et al (2016) Assessing sensory versus 14063–14068
optogenetic network activation by combining 77. Pachitariu M et al (2017) Suite2p: beyond
(o)fMRI with optical Ca2+ recordings. J Cereb 10, 000 neurons with standard two-photon
Blood Flow Metab 36:1885–1900 microscopy. BioRxiv
63. Yang JW, Prouvot PH, Stroh A, Luhmann HJ 78. Seth AK, Barrett AB, Barnett L (2015)
(2018) Combining optogenetics with MEA, Granger causality analysis in neuroscience and
depth-resolved LFPs and assessing the scope neuroimaging. J Neurosci 35:3293–3297
of optogenetic network modulation. Neuro- 79. Chen IW et al (2019) In vivo submillisecond
Methods 133:133–152 two-photon optogenetics with temporally
64. Vo-Dinh T (2003) Biomedical photonics focused patterned light. J Neurosci 39:
handbook. CRC Press 3484–3497
65. Dombeck DA, Khabbaz AN, Collman F, Adel-
man TL, Tank DW (2007) Imaging large-scale
All-Optical Physiology Pipeline 163
80. Marshel JH et al (2019) Cortical layer-specific 84. Jin D et al (2018) Analysis of activity states of
critical dynamics triggering perception. Science local neuronal microcircuits in mouse brain. In:
365:eaaw5202 26th European signal processing conference
81. Mardinly AR et al (2018) Precise multimodal (EUSIPCO)
optical control of neural ensemble activity. Nat 85. Richards BA et al (2019) A deep learning
Neurosci 21:881–893 framework for neuroscience. Nat Neurosci 22:
82. Bovetti S et al (2017) Simultaneous high-speed 1761–1770
imaging and optogenetic inhibition in the 86. Pnevmatikakis EA (2019) Analysis pipelines for
intact mouse brain. Sci Rep 7:40041 calcium imaging data. Curr Opin Neurobiol
83. Zaer H et al (2021) An intracortical implant- 55:15–21
able brain-computer interface for telemetric 87. Schwalm M et al (2017) Cortex-wide BOLD
real-time recording and manipulation of neu- fMRI activity reflects locally-recorded slow
ronal circuits for closed-loop intervention. oscillation-associated calcium waves. eLife 6:
Front Hum Neurosci 15:618626. https:// e27602. https://doi.org/10.7554/eLife.
doi.org/10.3389/fnhum.2021.618626 27602
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International
License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing, adaptation, distribution
and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative Commons license,
unless indicated otherwise in a credit line to the material. If material is not included in the chapter’s Creative
Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use,
you will need to obtain permission directly from the copyright holder.
Chapter 6
Abstract
All-optical experiments promise neuroscientists an unprecedented possibility to manipulate and measure
neuronal circuits with single-cell resolution. They rely on highly fine-tuned microscopes with complex
optical designs. Of similar importance are genetically encoded optical actuators and indicators that also have
to be optimized for such experiments. A particular challenge in these experiments is the detection of natural
firing patterns via genetically encoded indicators while avoiding optical cross-activation of neurons that are
photon-sensitized to allow optical replay of these patterns. Most optogenetic tools are sensitive in a broad
spectral range within the visible spectrum, which impedes artifact-free read-and-write access to neuronal
circuits. Nonetheless, carefully matching biophysical properties of actuators and indicators can permit
unambiguous excitation with a single wavelength in a so-called single-beam all-optical experiment.
In this chapter, we evaluate the current understanding of these biological probes and describe the
possibilities and limitations of those tools in the context of the all-optical single-beam experiment.
Furthermore, we review new insights into the photophysical properties of actuators, and propose a new
strategy for a single-beam two-photon excitation experiment to monitor activity minimizing cross-
activation with the actuators. Finally, we will highlight aspects for future developments of these tools.
1 Introduction
Over the last few years, the development of advanced optical stim-
ulation technologies in combination with novel optogenetic probes
foretold a bright future for the interrogation of brain function in
behaving animals. In particular, optically imprinting naturalistic
firing patterns onto neuronal circuits while simultaneously obtain-
ing activity readouts at a single-cell level is considered the holy grail
of a holistic understanding of neuronal circuits [1–3]. In this chap-
ter, we refer to such an experiment as an all-optical experiment and
offer a review on the possibilities and pitfalls of two-photon stimu-
lation of photon-sensitized single neurons while optically
Eirini Papagiakoumou (ed.), All-Optical Methods to Study Neuronal Function, Neuromethods, vol. 191,
https://doi.org/10.1007/978-1-0716-2764-8_6, © The Author(s) 2023
165
166 Damaris Holder and Matthias Prigge
1.1 Optogenetic Several advances are being made in tool development as new and
Tools for All-Optical better indicators and actuators are constantly being designed.
Experiments However, the most commonly used green-absorbing calcium indi-
cators GCaMP6 through 8 are still superior to their red-absorbing
counterparts jRGECO or K-GECO1 [23, 24]. It is believed that
only the pairing of red indicators and blue-absorbing actuators will
completely prevent cross-interference between imaging and stimu-
lation. Yet, for the more commonly used combination of a blue-
light absorbing indicator in combination with a red-absorbing
opsin the actuator is always cross-activated during imaging, since
retinal isomerization is triggered via the hypsochromic shifted exci-
tation light (shorter wavelength relative to maximum absorption
see Fig. 1a). This can elicit electronic transitions to higher elec-
tronic states (e.g., S0 ! S2), which are less efficient than those to a
lower energy state (S0 ! S1) [25, 26]. To some extent, lower
imaging light powers can mitigate such effects, but deeper imaging
and in vivo applications are often antagonistic to low power
imaging.
Despite ongoing developments toward better red-light absorb-
ing indicators, artifacts from photophysical processes induced by
blue light imaging excitation in these red absorbing indicators still
constitute a major obstacle for their widespread use in all-optical
experiments.
168 Damaris Holder and Matthias Prigge
Activation
lowest
lowest energy
higher energy transition
electronic transition
state transitions
Wavelength Wavelength
Fig. 1 Spectral combinations of indicators and actuators in an all-optical experiment, and their respective
cross-activation. (a) illustrates the overlap between the absorption spectra of GCaMP with the red-light
absorbing opsin ChRimson. Here, during hypsochromic imaging of the indicator, higher electronic state
transitions in the all-trans retinal can evoke photoactivation of ChRimson. While in (b) the absorption spectra
of RCaMP and the action spectra of ChR2 are shown. Here, both spectra are sufficiently separated to avoid
cross-activation of the blue-light absorbing opsin via bathochromic imaging, while contamination of RCaMP
emission during opsin photoactivation is minimal due to low absorption of the red indicator for shorter
wavelengths
soma
activation axon
activation ROI
1
reg ROI
ion 2
1
reg
ion
2
Fig. 2 Fundamental approaches for all-optical experiments. In (a) the expression of actuator and indicator are
separated through space. In this anatomical approach, full-field one-photon illumination of axons still triggers
the activation of actuators and hinders simultaneous imaging of indicators, but low power imaging of the
indicator can reduce cross-activation of the axon drastically. Moreover, if expression of the actuator is
restricted to the soma, photo-stimulation can be entirely spatially separated, rendering an artifact-free
simultaneous all-optical experiment possible. (b) Illustration of a cell-type-defined approach in which different
cell types express either an indicator or actuator, so that imaging and optogenetic stimulation are segmented
into small subregions containing only specific cell types. The approach in (c) is a variation of (b): while using a
distinct expression of indicators and actuators, here, technology to define arbitrary paths allows for the
selection of neurons with indicator expression in the entire field of view. In (d), a typical dual-color all-optical
experiment is illustrated. Here, both indicator and actuator are expressed in the same cell. In this case, the
choice of indicator/actuator combination is crucial. Usually green-absorbing indicators are employed along-
side red actuators to minimize optical crosstalk
170 Damaris Holder and Matthias Prigge
d d 1.4 μm
Ton_cell < 3 μs
rxy = 275 nm =Tline Toff_cell
- = 63 μs - 59.82 μs
Toff_cell Ton_cell
Fig. 3 Illustrations of scanning parameters and actuator stimulation patterns in an all-optical dual-color
configuration experiment. (a) For a given x-y plane, the scanning mirrors move a laser spot across the field of
view (FOV) with a given lateral rxy and axial resolution rz [33, 34]. (b) As an example, we assume a system with
a 16x objective with NA ¼ 0.8, yielding a FOV of 700 700 μm (dFOV); an 8-kHz resonant scanner, scanning
with a resolution of 512 512 lines; and a laser with frepetition ¼ 80 MHz, and a single pulse width of
Tpulse ¼ 140 fs, where we scan a single neuron with a dimension Dsoma ¼ 30 μm (in teal an idealized neuron,
which was approximated to a rectangular scan area (gray shaded area) of 30 30 μm for estimating
parameters). Scanning the entire FOV takes Tframe with a single line scan of the duration Tline. We define the
distance don-cell as the distance scanned by the laser on the cell and doff-cell as the distance the laser scans
over non-opsin expressing parts of the FOV. The distance d is defined in y-direction as the distance between
two scan lines. Depending on the resolution, the laser spot runs over a neuron expressing an actuator several
times per frame. Ton_cell is the time the laser spends scanning the cell for a given line. Depending on the
position of the neuron within the FOV, the subsequent line can give rise to another illumination period, i.e., cell
is located on the border of the FOV so that doff-cell is very small on one end of FOV. Within a single given frame,
a neuron in our example will be exposed to Nsoma ¼ 22 lines with each of them having Ton_cell < 3 μs
DA
N
H
R
τ
rec
N
R DA
LA
DA LA
H
O1 + O2 τ off
O1 + O2
H
des
τ rec
N
H
N R
O1 Time
R
all-trans 15-syn
I peak
13-cis 15-anti
Fig. 4 Unified photocycle for Channelrhodopsin2 in relation to its electrophysiological parameters [44, 45]. (a)
is a schematic depiction of the unifying photocycle where a single two-photon process can either trigger an
anti-photocycle from the dark-adapted (DA) ground state or the transition to the light-adapted (LA) state. The
LA state thermally relaxes back to the DA state or a second two-photon process can trigger a syn-photocycle.
After relaxation to the open state(s), O1 or O2 the molecule transitions back to the closed ground state of the
respective cycle. (b) Illustration of a two-pulse experiment in which two light pulses are given with increasing
delay times (Δt) while monitoring the recovery of peak photocurrents during the second light pulse. The
recovery of the peak photocurrent Trec after varying Δt obeys a monoexponential increase referring to the
transition of LA to DA (dotted red line)
174 Damaris Holder and Matthias Prigge
EE-RR
Soma (88aa) C-terminal fusion
Fig. 5 Overview of cellular localization of different target sequences. The neuron is divided into four segments:
axon initial segment, soma, somatodendritic and soma and proximal dendrites. Here, an overview is given
over the different molecular targeting strategies which are employed depending on which of these segments
are supposed to be expressing the opsin
Spatial and Temporal Considerations of Optogenetic Tools 177
5 Summary
Fig. 6 Comparison of different stimulation approaches for a single-beam experiment. (a) outlines the spiral
stimulation approach, while (b) elucidates the holographic approach, both combined with fast scanning
indicator imaging. The grey inserts outline the methods’ respective advantages and drawbacks
Acknowledgments
References
1. Marshel JH, Kim YS, Machado TA, Quirin S, 9. Ronzitti E, Ventalon C, Canepari M, Forget
Benson B, Kadmon J, Raja C, BC, Papagiakoumou E, Emiliani V (2017)
Chibukhchyan A, Ramakrishnan C, Inoue M, Recent advances in patterned photostimulation
Shane JC, McKnight DJ, Yoshizawa S, Kato for optogenetics. J Opt 19:113001. https://
HE, Ganguli S, Deisseroth K (2019) Cortical doi.org/10.1088/2040-8986/aa8299
layer–specific critical dynamics triggering per- 10. Nadella KMNS, Roš H, Baragli C, Griffiths VA,
ception. Science 365:eaaw5202. https://doi. Konstantinou G, Koimtzis T, Evans GJ, Kirkby
org/10.1126/science.aaw5202 PA, Silver RA (2016) Random-access scanning
2. Packer AM, Russell LE, Dalgleish HWP, microscopy for 3D imaging in awake behaving
H€ausser M (2015) Simultaneous all-optical animals. Nat Methods 13:1001–1004. https://
manipulation and recording of neural circuit doi.org/10.1038/nmeth.4033
activity with cellular resolution in vivo. Nat 11. Maschio MD, Stasi AMD, Benfenati F, Fellin T
Methods 12:140–146. https://doi.org/10. (2011) Three-dimensional in vivo scanning
1038/nmeth.3217 microscopy with inertia-free focus control.
3. Rickgauer JP, Deisseroth K, Tank DW (2014) Opt Lett 36:3503–3505. https://doi.org/10.
Simultaneous cellular-resolution optical per- 1364/OL.36.003503
turbation and imaging of place cell firing fields. 12. Accanto N, Molinier C, Tanese D, Ronzitti E,
Nat Neurosci 17:1816–1824. https://doi. Newman ZL, Wyart C, Isacoff E,
org/10.1038/nn.3866 Papagiakoumou E, Emiliani V (2018) Multi-
4. Akerboom J, Carreras Calderón N, Tian L, plexed temporally focused light shaping for
Wabnig S, Prigge M, Tolö J, Gordus A, high-resolution multi-cell targeting. Optica 5:
Orger M, Severi K, Macklin J, Patel R, 1478–1491. https://doi.org/10.1364/
Pulver S, Wardill T, Fischer E, Schüler C, OPTICA.5.001478
Chen T-W, Sarkisyan K, Marvin J, 13. Foust AJ, Zampini V, Tanese D,
Bargmann C, Kim D, Kügler S, Lagnado L, Papagiakoumou E, Emiliani V (2015)
Hegemann P, Gottschalk A, Schreiter E, Loo- Computer-generated holography enhances
ger L (2013) Genetically encoded calcium indi- voltage dye fluorescence discrimination in adja-
cators for multi-color neural activity imaging cent neuronal structures. Neurophotonics 2:
and combination with optogenetics. Front Mol 021007. https://doi.org/10.1117/1.NPh.2.
Neurosci 6:2. https://doi.org/10.3389/ 2.021007
fnmol.2013.00002 14. Dal Maschio M, Difato F, Beltramo R, Blau A,
5. Forli A, Vecchia D, Binini N, Succol F, Benfenati F, Fellin T (2010) Simultaneous
Bovetti S, Moretti C, Nespoli F, Mahn M, two-photon imaging and photo-stimulation
Baker CA, Bolton MM, Yizhar O, Fellin T with structured light illumination. Opt Express
(2018) Two-photon bidirectional control and 18:18720. https://doi.org/10.1364/OE.18.
imaging of neuronal excitability with high spa- 018720
tial resolution in vivo. Cell Rep 22:3087–3098. 15. Rickgauer JP, Tank DW (2009) Two-photon
https://doi.org/10.1016/j.celrep.2018. excitation of channelrhodopsin-2 at saturation.
02.063 Proc Natl Acad Sci 106:15025–15030.
6. Prigge M, Schneider F, Tsunoda SP, https://doi.org/10.1073/pnas.0907084106
Shilyansky C, Wietek J, Deisseroth K, Hege- 16. Lutz C, Otis TS, DeSars V, Charpak S, DiGre-
mann P (2012) Color-tuned channelrhodop- gorio DA, Emiliani V (2008) Holographic
sins for multiwavelength optogenetics. J Biol photolysis of caged neurotransmitters. Nat
Chem 287:31804–31812. https://doi.org/ Methods 5:821–827. https://doi.org/10.
10.1074/jbc.M112.391185 1038/nmeth.1241
7. Lecoq J, Orlova N, Grewe BF (2019) Wide. 17. Nikolenko V (2008) SLM microscopy: scanless
Fast. Deep: Recent advances in multiphoton two-photon imaging and photostimulation
microscopy of in vivo neuronal activity. J Neu- using spatial light modulators. Front Neural
rosci 39:9042–9052. https://doi.org/10. Circuits 2. https://doi.org/10.3389/neuro.
1523/JNEUROSCI.1527-18.2019 04.005.2008
8. Grewe BF, Helmchen F (2014) High-speed 18. Chaigneau E, Ronzitti E, Gajowa MA, Soler-
two-photon calcium imaging of neuronal pop- Llavina GJ, Tanese D, Brureau AYB,
ulation activity using acousto-optic deflectors. Papagiakoumou E, Zeng H, Emiliani V
Cold Spring Harb Protoc. https://doi.org/10. (2016) Two-photon holographic stimulation
1101/pdb.prot081778
Spatial and Temporal Considerations of Optogenetic Tools 183
of ReaChR. Front Cell Neurosci 10. https:// fluorescent genetically encoded calcium ion
doi.org/10.3389/fncel.2016.00234 indicators. Neuroscience
19. Shemesh OA, Tanese D, Zampini V, Linghu C, 28. Chavarha M, Villette V, Dimov IK, Pradhan L,
Piatkevich K, Ronzitti E, Papagiakoumou E, Evans SW, Shi D, Yang R, Chamberland S,
Boyden ES, Emiliani V (2017) Temporally pre- Bradley J, Mathieu B, St-Pierre F, Schnitzer
cise single-cell-resolution optogenetics. Nat MJ, Bi G, Toth K, Ding J, Dieudonné S, Lin
Neurosci 20:1796–1806. https://doi.org/10. MZ (2018) Fast two-photon volumetric imag-
1038/s41593-017-0018-8 ing of an improved voltage indicator reveals
20. Mardinly AR, Oldenburg IA, Pégard NC, electrical activity in deeply located neurons in
Sridharan S, Lyall EH, Chesnov K, Brohawn the awake brain. Neuroscience
SG, Waller L, Adesnik H (2018) Precise multi- 29. Platisa J, Vasan G, Yang A, Pieribone VA
modal optical control of neural ensemble activ- (2017) Directed evolution of key residues in
ity. Nat Neurosci 21:881–893. https://doi. fluorescent protein inverses the polarity of volt-
org/10.1038/s41593-018-0139-8 age sensitivity in the genetically encoded indi-
21. Pégard NC, Mardinly AR, Oldenburg IA, cator ArcLight. ACS Chem Neurosci 8:
Sridharan S, Waller L, Adesnik H (2017) 5 1 3 – 5 2 3 . h t t p s : // d o i . o r g / 1 0 . 1 0 2 1 /
Three-dimensional scanless holographic opto- acschemneuro.6b00234
genetics with temporal focusing (3D-SHOT). 30. Bando Y, Grimm C, Cornejo VH, Yuste R
Nat Commun 8. https://doi.org/10.1038/ (2019) Genetic voltage indicators. BMC Biol
s41467-017-01031-3 17:71. https://doi.org/10.1186/s12915-
22. Yang W, Carrillo-Reid L, Bando Y, Peterka DS, 019-0682-0
Yuste R (2018) Simultaneous two-photon 31. Gruver KM, Watt AJ (2019) Optimizing opto-
imaging and two-photon optogenetics of cor- genetic activation of purkinje cell axons to
tical circuits in three dimensions. elife 7: investigate the purkinje cell – DCN synapse.
e32671. https://doi.org/10.7554/eLife. Front Synap Neurosci 11:31. https://doi.
32671 org/10.3389/fnsyn.2019.00031
23. Dana H, Mohar B, Sun Y, Narayan S, 32. Yizhar O, Fenno LE, Davidson TJ, Mogri M,
Gordus A, Hasseman JP, Tsegaye G, Holt GT, Deisseroth K (2011) Optogenetics in neural
Hu A, Walpita D, Patel R, Macklin JJ, Barg- systems. Neuron 71:9–34. https://doi.org/
mann CI, Ahrens MB, Schreiter ER, 10.1016/j.neuron.2011.06.004
Jayaraman V, Looger LL, Svoboda K, Kim DS 33. Young MD, Field JJ, Sheetz KE, Bartels RA,
(2016) Sensitive red protein calcium indicators Squier J (2015) A pragmatic guide to multi-
for imaging neural activity. Neuroscience photon microscope design. Adv Opt Photon 7:
24. Shen Y, Dana H, Abdelfattah AS, Patel R, 276–378. https://doi.org/10.1364/AOP.7.
Shea J, Molina RS, Rawal B, Rancic V, Chang 000276
Y-F, Wu L, Chen Y, Qian Y, Wiens MD, 34. Zipfel WR, Williams RM, Webb WW (2003)
Hambleton N, Ballanyi K, Hughes TE, Nonlinear magic: multiphoton microscopy in
Drobizhev M, Kim DS, Koyama M, Schreiter the biosciences. Nat Biotechnol 21:
ER, Campbell RE (2018) A genetically 1369–1377. https://doi.org/10.1038/
encoded Ca2+ indicator based on circularly per- nbt899
mutated sea anemone red fluorescent protein 35. Govardovskii VI, Korenyak DA, Shukolyukov
eqFP578. BMC Biol 16:9. https://doi.org/ SA, Zueva LV (2009) Lateral diffusion of rho-
10.1186/s12915-018-0480-0 dopsin in photoreceptor membrane: a reap-
25. Dokukina I, Weingart O (2015) Spectral prop- praisal. Mol Vis 13
erties and isomerisation path of retinal in C1C2 36. Yang W, Yuste R (2018) Holographic imaging
channelrhodopsin. Phys Chem Chem Phys 17: and photostimulation of neural activity. Curr
25142–25150. https://doi.org/10.1039/ Opin Neurobiol 50:211–221. https://doi.
C5CP02650D org/10.1016/j.conb.2018.03.006
26. Hasson KC, Gai F, Anfinrud PA (1996) The 37. Adamantidis A, Arber S, Bains JS, Bamberg E,
photoisomerization of retinal in bacteriorho- Bonci A, Buzsáki G, Cardin JA, Costa RM,
dopsin: experimental evidence for a three- Dan Y, Goda Y, Graybiel AM, H€ausser M,
state model. Proc Natl Acad Sci 93: Hegemann P, Huguenard JR, Insel TR, Janak
15124–15129. https://doi.org/10.1073/ PH, Johnston D, Josselyn SA, Koch C, Kreitzer
pnas.93.26.15124 AC, Lüscher C, Malenka RC, Miesenböck G,
27. Dalangin R, Drobizhev M, Molina RS, Nagel G, Roska B, Schnitzer MJ, Shenoy KV,
Aggarwal A, Patel R, Abdelfattah AS, Zhao Y, Soltesz I, Sternson SM, Tsien RW, Tsien RY,
Wu J, Podgorski K, Schreiter ER, Hughes TE, Turrigiano GG, Tye KM, Wilson RI (2015)
Campbell RE, Shen Y (2020) Far-red
184 Damaris Holder and Matthias Prigge
Mol Biol 330:553–570. https://doi.org/10. 61. Kim CK, Miri A, Leung LC, Berndt A,
1016/S0022-2836(03)00576-X Mourrain P, Tank DW, Burdine RD (2014)
55. Mattis J, Tye KM, Ferenczi EA, Prolonged, brain-wide expression of nuclear-
Ramakrishnan C, O’Shea DJ, Prakash R, localized GCaMP3 for functional circuit
Gunaydin LA, Hyun M, Fenno LE, mapping. Front Neural Circuits 8. https://
Gradinaru V, Yizhar O, Deisseroth K (2012) doi.org/10.3389/fncir.2014.00138
Principles for applying optogenetic tools 62. Shemesh OA, Linghu C, Piatkevich KD,
derived from direct comparative analysis of Goodwin D, Celiker OT, Gritton HJ, Romano
microbial opsins. Nat Methods 9:159–172. MF, Gao R, Yu C-C, Tseng H-A, Bensussen S,
https://doi.org/10.1038/nmeth.1808 Narayan S, Yang C-T, Freifeld L, Siciliano CA,
56. Katz E, Stoler O, Scheller A, Khrapunsky Y, Gupta I, Wang J, Pak N, Yoon Y-G, JFP U,
Goebbels S, Kirchhoff F, Gutnick MJ, Wolf F, Guner-Ataman B, Noamany H, Sheinkopf ZR,
Fleidervish IA (2018) Role of sodium channel Park WM, Asano S, Keating AE, Trimmer JS,
subtype in action potential generation by neo- Reimer J, Tolias AS, Bear MF, Tye KM, Han X,
cortical pyramidal neurons. Proc Natl Acad Sci Ahrens MB, Boyden ES (2020) Precision cal-
115:E7184–E7192. https://doi.org/10. cium imaging of dense neural populations via a
1073/pnas.1720493115 cell-body-targeted calcium indicator. Neuron
57. Grubb MS, Burrone J (2010) Activity- 107:470–486.e11. https://doi.org/10.1016/
dependent relocation of the axon initial seg- j.neuron.2020.05.029
ment fine-tunes neuronal excitability. Nature 63. Perez-Alvarez A, Fearey BC, O’Toole RJ,
465:1070–1074. https://doi.org/10.1038/ Yang W, Arganda-Carreras I, Lamothe-Molina
nature09160 PJ, Moeyaert B, Mohr MA, Panzera LC,
58. Wu C, Ivanova E, Cui J, Lu Q, Pan Z-H Schulze C, Schreiter ER, Wiegert JS, Gee CE,
(2011) Action potential generation at an axon Hoppa MB, Oertner TG (2020) Freeze-frame
initial segment-like process in the axonless reti- imaging of synaptic activity using SynTagMA.
nal aii amacrine cell. J Neurosci 31: Nat Commun 11:2464. https://doi.org/10.
14654–14659. https://doi.org/10.1523/ 1038/s41467-020-16315-4
JNEUROSCI.1861-11.2011 64. Ernst OP, Murcia PAS, Daldrop P, Tsunoda SP,
59. Zhang Z, Feng J, Wu C, Lu Q, Pan Z-H Kateriya S, Hegemann P (2008) Photoactiva-
(2015) Targeted expression of tion of channelrhodopsin. J Biol Chem 283:
channelrhodopsin-2 to the axon initial seg- 1637–1643. https://doi.org/10.1074/jbc.
ment alters the temporal firing properties of M708039200
retinal ganglion cells. PLoS ONE 10: 65. Nikolic K, Grossman N, Grubb MS, Burrone J,
e0142052. https://doi.org/10.1371/journal. Toumazou C, Degenaar P (2009) Photocycles
pone.0142052 of channelrhodopsin-2. Photochem Photobiol
60. Mahn M, Gibor L, Patil P, Cohen-Kashi 85:400–411. https://doi.org/10.1111/j.
Malina K, Oring S, Printz Y, Levy R, Lampl I, 1751-1097.2008.00460.x
Yizhar O (2018) High-efficiency optogenetic 66. Lin JY (2011) A user’s guide to channelrho-
silencing with soma-targeted anion-conduct- dopsin variants: features, limitations and future
ing channelrhodopsins. Nat Commun 9: developments: a user’s guide to channelrho-
4125. https://doi.org/10.1038/s41467- dopsin variants. Exp Physiol 96:19–25.
018-06511-8 https://doi.org/10.1113/expphysiol.2009.
051961
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International
License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing, adaptation, distribution
and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative Commons license,
unless indicated otherwise in a credit line to the material. If material is not included in the chapter’s Creative
Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use,
you will need to obtain permission directly from the copyright holder.
Chapter 7
Abstract
Miniaturized head-mounted microscopes for in vivo recording of neural activity have gained much recog-
nition within the past decade of neuroscience research. In combination with fluorescent reporters, these
miniature microscopes allow researchers to record the neural activity that underlies behavior, cognition, and
perception in freely moving animals. Single-photon miniature microscopes are convenient for widefield
recording but lack the increased penetration depth and optical sectioning capabilities of multiphoton
imaging. Here we discuss the current state of head-mounted multiphoton miniature microscopes and
introduce a miniature head-mounted two-photon fiber-coupled microscope (2P-FCM) for neuronal
imaging with active axial focusing enabled using a miniature electrowetting lens. The 2P-FCM enables
three-dimensional two-photon optical recording of structure and activity at multiple focal planes in a freely
moving mouse. Detailed methods are provided in this chapter on the 2P-FCM design, operation, and
software for data analysis.
Key words Two-photon microscopy, Fluorescence microscopy, Lens system design, Fiber optics,
Ultrafast pulse propagation, Neural imaging, Image processing
1 Introduction
Eirini Papagiakoumou (ed.), All-Optical Methods to Study Neuronal Function, Neuromethods, vol. 191,
https://doi.org/10.1007/978-1-0716-2764-8_7, © The Author(s) 2023
187
188 Baris N. Ozbay et al.
are now being developed that can allow the animal to be unre-
stricted with more naturalistic behavior. The first miniature micro-
scopes to be developed for this purpose were one-photon widefield
miniscopes [4–7]. These systems rely on innovations in compact
CMOS imaging sensors and efficient visible wavelength light emit-
ting diodes for the excitation source. Despite the success of the
miniscopes with thousands deployed in laboratories internationally,
there are limitations. One-photon widefield excitation in tissue
results in high levels of light scattering, reducing signal to noise as
the out-of-focus fluorescence is also collected on the imaging
detector. Computational methods are typically used to select out
transient fluorescent signals from the background from the image
data. Detailed structural information of the brain region being
imaged is sometimes not possible to obtain. Recently, modified
miniscopes have been developed that place a phase plate [8] or
microlens array [9] in the optical detection path which provides
additional information for computational reconstruction in three
dimensions (3D). However, light scattering, particularly in densely
labelled samples, ultimately limits the imaging depth in these 3D
miniscopes.
In contrast to one-photon widefield, two-photon microscopy
removes the out-of-focus fluorescence and provides cross-sectional
imaging only at the focus. 2P imaging provides excellent signal to
noise and detailed structural information allowing high-resolution
images of cells and processes. Previous work demonstrated minia-
ture laser-scanning multiphoton fiber-coupled microscopes for
head-mounted neuronal imaging in freely moving rodents using
different methods of laser scanning including miniaturized micro-
electromechanical systems (MEMS) actuated mirror [10, 11] or
piezoelectric scanner [12] within the head mount. However, none
of these designs leverages the optical sectioning capabilities for full
volumetric 3D imaging without introducing cumbersome com-
plexity and weight. Recently, Zong et al. integrated an optical
design using a MEMS mirror for lateral scanning and a tunable
lens for axial scanning in a head-attached miniature microscope
[13, 14]. However, this miniature microscope is complex to align
and involves custom lenses and tunable optics. In contrast, the
design reported here incorporates a miniature, compact electrowet-
ting tunable lens (EWTL) for increasing the versatility of
two-photon capabilities by adding active axial focusing, while
using a coherent imaging fiber bundle for lateral scanning. The
design uses only commercially available components and can be
implemented on any standard bench-top laser scanning
microscope.
Miniature Multiphoton Microscopes 189
2 Background
2.1 Optical Design The objectives used for two-photon excitation microscopy in deep
Considerations for tissues typically have a numerical aperture (NA) greater than 0.8
Miniature 2P [2]. Imaging with high NA objectives reduces the spot size of the
Microscopes laser, thereby increasing the two-photon fluorescence signal for a
given excitation power. However, large NA objectives are a prob-
lem for the manufacture of small imaging systems, because the
system aperture is constrained by the diameter of the physical
optics, which in turn forces the optical design to compromise on
other important parameters, such as working distance and field-of-
view [15]. Assuming a fixed aperture size, larger NA lenses require
a reduction in the working distance. Additionally, the complexity of
a lens design can be viewed as a function of its optical invariant, i.e.,
ability to collect light over a large angle and field-of-view (FOV).
Increasing the NA has the effect of making it more challenging to
design for a large FOV as well. When considering the fluorescence
signal detection path, the fluorescent photons are only emitted at
the focus and detected on a photo-multiplier tube (PMT) such that
the background noise of the image is not degraded by scattered
photon trajectories. Therefore, the excitation NA and the collec-
tion NA, more commonly referred to as Etendue, must be consid-
ered separately. The time-averaged intensity of fluorescence
collected geometrically is:
If ,col ffi If ,gen Ωf ð1Þ
where Ωf is the fractional collection efficiency of an imaging system,
summarized as:
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2
11 1 NAcol
n1
Ωf ¼ ð2Þ
2
When imaging into scattering tissue, the excitation NA is not as
important as the collection NA. This is because high-angle rays
have longer optical paths, with higher probability of scattering,
and are affected by aberrations. Additionally, there are diminishing
returns as the focal volume decreases with higher NA. Therefore, a
good efficiency two-photon imaging system can be designed that
has a relatively low NA (~0.4–0.5) excitation path if the collection
NA path can be increased independently. Such systems are often
employed by using light collection paths that involve large-area
detectors, which can be replicated with large effective-area optical
fibers in miniaturized microscopes [10, 16].
190 Baris N. Ozbay et al.
2.2 Optical Design Unlike methods of axial scanning that physically move the sample
Considerations for or the objective lens, the use of a tunable lens to adjust the axial
Axial Scanning with a focus changes the way light propagates through the optical system.
Tunable Lens More specifically, the curvature at the back focal plane of the
objective lens is modified. When propagated through a lens, this
curvature of the wavefront gets transformed into an axial displace-
ment of the paraxial focus spot. This is analogous to how a tilted
wavefront becomes transformed by a lens into a transverse displace-
ment of the focus.
One important consideration when designing such a system is
the magnitude of curvature that is needed to achieve a desired axial
scan range. An EWTL placed in the back focal plane of a lens
directly shapes the wavefront entering the lens.
The wavefront phase term ϕ can be written as a function of the
pupil radius ρ [17]:
NA2 2
ϕðρÞ ¼ kZc ρ ð3Þ
2M2
where M is the objective magnification and Zc is the distance at
which the focus is formed with a given quadratic phase. Rearran-
ging this in terms of the amount of axial change for a given change
in wavefront phase curvature gives:
ϕ ρ2
¼ kNA 2 ð4Þ
Zc 2M2
This last expression shows that the ability of a tunable lens to
change the axial focus of a lens system is dependent inversely on the
square of the objective magnification. This means that designing
for high magnification with the purpose of increasing resolution or
NA will result in a decrease in axial scan range for a given EWTL.
Another consideration is that any imaging system that produces
magnification cannot perfectly represent both the axial and trans-
verse focus transformations simultaneously. This effect is illustrated
in Fig. 1. Botcherby et al. describe the consequences of this on
imaging properties [18]. In lens design, optical engineers typically
optimize for either transverse or axial focusing. Most commonly
transverse imaging is preserved because imaging lenses are expected
to have uniform performance throughout the field-of-view (FOV).
The consequence is an accumulation of spherical aberrations along
the axial range of the lens, shown in Fig. 1c. One can also design for
the Herschel condition, which instead forms perfect focus spots
along the optical axis but accumulates aberrations in the transverse
dimension.
The FCM imaging system designs presented here are opti-
mized primarily for the Herschel condition, by constraining the
optimization parameters in Zemax optical design software to
Miniature Multiphoton Microscopes 191
Fig. 1 Axial focusing by wavefront curvature shaping at the objective back focal plane. (a) An input converging
or diverging wavefront is transformed into an axial translation away from the designed focal length of the lens.
A lens producing an ideal diffraction limited focus at the design working distance (b) will not produce a
diffraction limited focus when the input wavefront is divergent (c) due to spherical aberrations as a result of
the Abbe sine condition
2.3 Cranial Windows Cranial windows provide optical access to surface-level brain struc-
tures with traditional microscope objectives. Typically, two-photon
laser scanning microscopy (2P-LSM) is performed on a head-fixed
mouse with a coverslip implanted in a small craniotomy above the
brain region of interest, normally 3 mm2 or less in size. There are
several variations on windows that show the flexibility of this tech-
nique [19–21]. A similar procedure is skull-thinning and polishing,
which can be healthier and stable for the brain tissue, but is time-
consuming and may not be as optically clear [22]. Recently, Wang
et al. demonstrated that application of index-matching epoxy to the
bone can reduce scattering and enhance optical access [23]. Remov-
able windows and permeable windows have also been
described [24].
Some examples of common regions accessed by cranial win-
dows are olfactory bulb [25], barrel cortex [26], and visual cortex
[27]. These regions display high levels of behavior-dependent neu-
ronal activity relatively close to the surface. The extended depth-
limit of 2P-LSM, compared with widefield microscopy or confocal
laser scanning microscopy, allows access to neurons about 500 μm
below the brain surface. In terms of cortical layers in mouse,
500 μm is approximately the depth of the cell bodies found in layers
4/5 of structures such as the motor and somatosensory cortex.
Three-photon excitation microscopy has been recently demon-
strated to reach structures such as the hippocampus, located around
1 mm below the surface in the mouse brain. However, for deeper
structures, it is necessary to excavate the tissue above the target
structure or to implant relay optics, such as GRIN lenses [28].
2.4 Gradient- GRIN lenses have been classically used for simple and space-
Refractive Index efficient collimation of the divergent light exiting a single fiber
(GRIN) Lenses core. In the past decade, GRIN lenses have been re-purposed for
miniaturized imaging applications [29, 30]. The radial profile of
the refractive index, ng, in a GRIN lens varies according to [31]:
g 2r 2
ng ðr Þ ¼ ng:0 1 ð5Þ
2
where g is the gradient constant, r is the radial distance from the
optical axis, and ng.0 is the index of refraction at r ¼ 0. Rays
launched into one end of the GRIN lens will be focused based on
the axial length of the lens, defined by the pitch length, pL:
2π
pL ¼ ð6Þ
g
Miniature Multiphoton Microscopes 193
Fig. 2 GRIN lens illustrations. Top: Single GRIN lens with pitch ~0.5 with a
magnification of 1. Bottom: Two-element GRIN lens with low NA relay at pitch
~0.75 and high NA objective with pitch ~0.25
2.5 Fiber-Coupling of For 2P-FCM imaging a laser-source and point-scanning system are
Excitation Laser required. For the laser source, high-intensity laser light must be
brought into the miniature microscope enclosure using an optical
fiber. Fibers can be selected for the most efficient propagation of
the laser light, generally such that the fiber-core supports only the
fundamental transverse electromagnetic (TEM00) mode, in which
case it is known as a single-mode fiber. In the special case of
ultrashort laser pulses required for 2P-LSM, the maintenance of
the peak pulse power is dependent on chromatic dispersion, modal
dispersion, polarization dispersion, and non-linear effects
[35]. Pre-conditioning the pulses can help to cancel these different
effects of propagation through the fiber to maintain high peak
powers [36–38]. Alternatively, it is possible to use hollow-core
photonic bandgap fibers [39], in which the electric field exists
mostly in an air-filled core, minimizing chromatic dispersion and
nonlinearities. The choice of fiber should also depend on its weight,
flexibility, and cost for the practical purposes of mouse imaging.
Regardless of the choice, a single fiber core must be focused onto
the sample and actively scanned to form an image.
Fiber-coupled scanning microscopy is a relatively well-
developed field, owed to the progress in endoscopic surgical and
diagnostic clinical tools that use miniaturized imaging heads cou-
pled with fiber-optics to a detection system [40]. There are many
techniques for accomplishing this, two of the most common being
piezoelectric resonance scanning of the fiber-tip [12, 41–44] and
integrated microelectromechanical systems (MEMS) actuated mir-
ror scanning [10, 11, 45].
An extra consideration for these single-fiber delivery systems is
that single-mode fiber-cores are inefficient for the collection of
fluorescence emission. Solutions include the use of large mode-area
fibers flanking the excitation core [16], miniature single-element
detectors mounted on the microscope head [46], and a separate
emission path for large-area collection fibers [10, 11, 47, 48].
Overall, these single-fiber-delivery techniques allow access to
high-efficiency laser-transmission, which is especially useful for
2P-LSM because of the difficulty required in maintaining ultrashort
pulse integrity at the focus. The cost of these techniques is the
added mechanical complexity required to implement miniaturized
laser-scanning. These may result in challenges in the optical design
that limit imaging performance. Frequently, it is difficult to achieve
large scan ranges or maintain large beam apertures using small
mirrors. So far, these systems have been discussed only in the
context of transverse scanning. Miniature fiber-scanning micro-
scopes that implement a third axial scanning dimension can become
overly cumbersome in their complexity.
Miniature Multiphoton Microscopes 195
Fig. 3 Properties of coherent imaging fiber bundles. Left: Illustration of proximal-to-distal image coherence of
CIFB, as well as the pixelation of an image formed through the bundle because of the discrete fiber cores.
Right: Fujikura CIFB fluorescence image of a uniform target showing core distribution. Image was acquired
using a laser scanning microscope focused on the proximal surface of the CIFB and placing a uniform
fluorescent sample at the distal end
Fig. 4 Illustration of material dispersion. The initial pulse out of the laser is short with all spectral frequencies
in phase. Inside the 1 m glass fiber, the different frequencies travel with different velocities, resulting in a
pulse that is stretched out in time. The resulting pulse is positively chirped, where the lower frequencies (red)
travel faster than the higher frequencies (blue)
Fig. 5 Imaging heterogeneity in CIFB. Images taken using the 2P-FCM showing the raw image of the proximal
end of the CIFB focused on a NIR phosphorescent detector card to demonstrate one-photon excitation and a
uniform green fluorescent test slide to demonstrate two-photon excitation
2.8 Tunable Focus by Liquid lenses are a common variety of electrically tunable lens
Liquid Lens (ETL) that use one or more liquids as a shape-changing refractive
Technology interface to allow for focus adjustment with minimal mechanical
motion [67]. There are several common liquid ETL types that may
be used for performing high-speed optical focusing for laser-
scanning microscopy [68]. Large-aperture liquid ETLs are particu-
larly suited for bench-top systems because of their large focal range,
high speed, and repeatability. An example of a commercially avail-
able 10-mm aperture, 30-mm diameter ETL (EL-10-30-C-MV,
Optotune AG, Switzerland) functions by rapidly changing the
volume of liquid in a container with a flexible polymer surface,
which results in a focal length shift. Large aperture liquid ETLs
have enabled rapid axial focusing, up to 100 Hz, in compact laser
scanning confocal microscopes (LSCM), 2P-LSM, and selective
plane illumination microscopy (SPIM) systems without mechanical
movement of the objective or sample [69–72]. However, shape-
changing polymer ETLs are not suitable for miniaturized head-
attached microscopes because of their large size and susceptibility
to orientation and vibration.
Electrowetting tunable lenses (EWTLs) are another type of
liquid ETL that has applications in microscopy [73]. An example
is the Arctic 316, made by Varioptic, France, with outer diameter
7.8-mm and 2.5-mm aperture. Electrowetting is a method for
changing the wettability of a liquid on a dielectric surface by apply-
ing a voltage across the interface, effectively changing the contact
200 Baris N. Ozbay et al.
angle of the liquid with the surface. To form a lens, two fluids with
dissimilar refractive-indices and hydrophobicity are placed in a
container with dielectric sidewalls. The hydrophilic polar liquid
has dissolved impurities to allow it to react to an external electric
field. The contact angle of the liquid interface to the sidewalls of the
lens can be changed by applying an electric field across the lens.
Eventually an equilibrium is reached between the electrostatic
forces acting on the polar liquid and the surface tension of the
system to create a lens surface with a stable curvature.
Carefully matching the density of the liquids at these scales
makes EWTLs very resistant to the effects of gravity due to orien-
tation and vibrations, which makes it ideally suited for a miniature
head-mounted system. Commercial EWTLs have also demon-
strated very large tuning ranges, on the order of 50 diopters. Fur-
ther, the responses of EWTLs to input voltages are well described
by simple oscillators. Although EWTL focusing speed is not as
rapid as some other ETLs, by engineering the voltage input func-
tion, the lens response time can be brought to less than 20 ms,
which is within the range of what is required for an active scanning
system [74]. EWTLs also have the potential to perform extended
optical functions, such as beam steering and wavefront shaping
[75, 76]. Because of the small aperture size, these lenses are not
frequently used in bench-top microscopy applications. However,
EWTLs are good candidates for robust axial focusing solution for
miniature microscopes.
3 Methods
3.1 Overall System The experimental setup for imaging with the 2P-FCM is shown in
Design Fig. 6. The distal optics of the 2P-FCM are housed in a two-part
3D-printed enclosure, which can be repeatedly attached to a base-
plate affixed to the animal’s head. A CIFB couples the distal imag-
ing optics to a custom 2P-LSM system to relay the excitation laser
and collect emitted fluorescence from the sample. Laser-scanning
over the CIFB forms an image of the sample while the fluorescence
Fig. 6 2P-FCM imaging system. Pulses from a Ti:Sapphire laser source are spectrally broadened through
polarization maintaining fiber (PMF) and chirped using a grating-pair pulse stretcher. Output pulses are focused
and scanned onto the surface of the CIFB through scanning mirrors, scan/tube lens relay, and 10x Objective.
Fluorescence emission is collected by the CIFB and directed to a photon counting PMT through a dichroic filter.
The collected signal is amplified and transformed to logic levels to be detected by the DAQ counter
202 Baris N. Ozbay et al.
3.1.2 Spectral and In order to obtain the shortest pulses at the sample, the laser is first
Temporal Pulse Pre- sent through a pre-compensation setup optimized for propagating
compensation through a 1.0 m length CIFB (FIGH-15-600 N, Fujikura). The
output from the laser is focused into a 0.5 m length polarization
maintaining (PM) fiber (PM780-HP, Thorlabs) that causes spectral
broadening of the pulse. The output of the PMF is collimated with
a fixed fiber-collimation lens (F220APC-780, Thorlabs). The
1.0 mm beam is then expanded through a 3.75 beam expander
to reduce the irradiance on the gratings. The beam is reflected off of
the two gratings at near Littrow angle separated by 265 mm. It is
double passed through the grating pair using a retroreflector with
the beam exiting at a different height. The gratings are reflective
ruled gold with a density of 300 grooves/mm (49-572, Edmund
Optics). The grating-pair stretcher applies ~66,000 fs2 of negative
GDD to the laser pulse to compensate for the positive dispersion
from the 0.5 m length PMF, 1.0 m length CIFB, and additional
non-linear dispersion at the power levels used in the experiments.
The precise grating separation distance was determined empirically
by maximizing the fluorescence signal while imaging a fluorescent
test slide (Chroma Technology) through the 2P-FCM. After
traveling through the grating pair, the beam size is reduced by
2.5 with a reverse beam expander to optimize the beam diameter
for coupling into the CIFB.
3.1.3 2P-LSM Bench-Top The beam is scanned through a galvanometric mirror scanning
System system (6215H, Cambridge Technologies), relayed through a
50 mm FL scan lens and 180 mm FL tube lens in an Olympus
IX71 microscope. The beam is focused and laterally scanned on the
surface of the CIFB using a 10/0.4 NA Olympus UPLANSApo
objective lens. A XYZ-translation stage (CXYZ05, Thorlabs) is used
to accurately align the fiber to the focus of the objective lens. The
beam propagates through the CIFB and then through the distal
miniature optics and focused onto the sample. The fluorescence
emission is collected by the same optics, back through the CIFB.
Because the miniature optics are not chromatically corrected, green
emission light is projected onto a ~50 μm diameter area on the
CIFB surface. Power transmission through the CIFB was measured
Miniature Multiphoton Microscopes 203
using a 532 nm light source and was found to have a ~ 67% effi-
ciency through the CIFB when collected through multiple fiber-
cores. The 10 objective collects the fluorescence emission from
the CIFB and is separated from the excitation laser path by a
primary dichroic mirror (T670LPXR, Chroma Technology). A
second dichroic mirror (FF562-Di02, Semrock) splits the red and
green fluorescence that is focused using an achromatic doublet lens
(LB1761-A, Thorlabs) on separate large-area photon counting
photo-multiplier tubes (PMT) (H7422P-40, Hamamatsu). The
output pulses from the PMTs are amplified by a high-bandwidth
amplifier (ACA-4-35db, Becker & Hickl GmbH) and converted to
logic-level pulses by a timing discriminator (6915, Phillips Scien-
tific). The pulses are counted by a data-acquisition card (PCIe-
6259, National Instruments) at a rate of 20 MHz. The counts are
sampled and binned by pixels and converted into an image in
custom software in Labview (National Instruments) that also con-
trols the EWTL driver.
3.1.4 2P-FCM Miniature The imaging system for the 2P-FCM was designed in Zemax
Optical System Design optical design software. Models for the stock lenses and for the
EWTL were obtained from the manufacturers (Edmund Optics,
Thorlabs, and Varioptic). The CIFB (Fujikura Ltd. FIGH-15-
600 N) has an outer diameter of 700 μm, an active image diameter
of 550 μm, and a length of 1.0 m. There are ~15,000 cores with
core-to-core spacing of 4.5 μm and core diameter of 3.2 μm, as
previously reported by Chen et al. measured with scanning electron
microscopy [55].
The miniature optics contained in the head-mounted 2P-FCM
imaging system are shown in Fig. 7. The fiber-coupling lens is an
asphere (FL: 6.2-mm, diameter: 4.7-mm, Edmund Optics 83-710)
to collimate the light diverging from the fiber bundle. The electro-
wetting lens is placed in the collimated beam and the light is
refocused onto the sample by an objective lens consisting of a
plano-convex lens (FL: 7.5-mm, diameter: 3.0-mm, Edmund
Optics 49-177), and an aspheric lens (FL: 2.0 mm, diameter:
3.0 mm, Thorlabs 355151-B). The nominal magnification of this
imaging system is 0.4 and field-of-view (FOV) is ~220-μm,
corresponding to the de-magnified CIFB active imaging diameter.
Similarly, the lateral sampling resolution is the de-magnified core
spacing, which is ~1.8 μm.
A commercially available EWTL (Varioptic Arctic 316) is used
to control the axial focusing of the 2P-FCM imaging system. The
predicted working distance and other imaging properties from
Zemax at these three different voltage settings are summarized in
Table 1. The optical power range of the EWTL is specified as 16
to +36 diopters. The optical system is optimized through the full
focal range of the EWTL to minimize magnification change, maxi-
mize the axial scan range, and maximize the working distance of the
2P-FCM.
204 Baris N. Ozbay et al.
Fig. 7 Optics of the 2P-FCM miniature microscope head that focuses excitation light from CIFB cores onto the
tissue. The CIFB-coupling asphere collects the light from the cores of the CIFB, which are then passed through
the aperture of the EWTL. The plano-convex lens and the objective asphere focus the light onto the tissue
through a #1 coverglass with 0.15-mm thickness
Table 1
2P-FCM optical parameters from Zemax through a range of focal lengths
EWTL control (VRMS) EWTL optical power (m1) Working distance (μm) Magnification NA
60 35 450 0.41 0.43
42 0 570 0.40 0.44
25 -16 690 0.39 0.45
3.1.5 3D-Printed The enclosure for the 2P-FCM optics is designed in Solidworks 3D
Miniature Head-Mount CAD software (Dassault Systemes). The packaging is split into
Design three sections: top, bottom, and baseplate, shown in Fig. 9. The
top-section contains the CIFB ferrule, held in place by two
set-screws, and the fiber-collimating asphere. The bottom-section
contains the objective lenses. The unmounted lenses are held in by
friction in precisely sized openings. The top section has two curved
tabs that interface with slots in the bottom section, which help to
ensure reproducible alignment. The EWTL and the electrode are
sandwiched between the bottom section and the top section with
an O-ring that ensures good electrical contact. The flat-flex
Miniature Multiphoton Microscopes 205
Fig. 8 Numerical aperture comparison of forward excitation light at 910 nm (top) and backward emission light
at 532 nm (bottom). Note that the blue lines indicate the optical rays entering the system on axis, while the
green and red are off-axis rays. Largest emission and excitation field positions are matched at 220-μm FOV.
Note that the backward emission is defocused on the fiber bundle end so that multiple fibers are used collect
the fluorescent emission
Fig. 9 3D-printed enclosure. (a) A two-part 2P-FCM snaps together to secure the EWTL and electrode. (b)
Photo of 3D-printed parts, top: before any processing with supports still attached and bottom: assembled
2P-FCM
206 Baris N. Ozbay et al.
electrode cable exits the enclosure through a small slot between the
sections. The top-section tabs have a single thread at the end, which
interfaces with the baseplate as shown in Fig. 9a. In this way, the
baseplate is pulled up against the bottom section by the top section.
This greatly improves rigidity when attached to a moving animal.
The baseplate is designed with ridges and holes to improve the
adhesion of the cement for attachment to the animal skull. The
entire enclosure is 3D-printed using a high-resolution projection-
based resin printer (Kudo3D Titan 1), with a resolution of 50 μm.
The material used is a photo-curable resin (3DM-XGreen) dyed
with 0.5% molybdenum disulfide to decrease light scattering and
thus increase feature resolution. A photo of the top and bottom
sections immediately after printing is shown in Fig. 9b. 3D printing
allows optimization of the prototype and easily enables design
changes, such as the inclusion of GRIN lenses for deep-brain
imaging.
3.2 Test Sample Resolution and axial scan range measurements were performed by
Preparation imaging fluorescent beads embedded in agarose (Sigma-Aldrich
A9414) and a USAF 1951 resolution target (Edmund Optics
38-257). 2-μm yellow-green fluorescent beads (Invitrogen
F8853) were used to measure axial scanning extent as well as lateral
and axial resolution.
Low-melting-point agarose was prepared at a concentration of
0.5% in water. The 2-μm diameter fluorescent yellow-green beads
were diluted in the agarose to a concentration of ~2.0 107 beads/
mL. Approximately 2.0 mL of solution was placed on a #1 cover-
glass and allowed to set at room temperature. The beads were
imaged in sequential axial planes by a 20x 0.75 NA Olympus
UPLANSApo objective with a motorized stage and separately by
the 2P-FCM by changing the voltage applied to the EWTL.
3.3 Mouse Imaging All experiments were approved and conducted in accordance with
Setup the Institutional Animal Care and Use Committee of the University
of Colorado Anschutz Medical Campus. Three-month-old male
C57BL/6 mice (neocortex recordings, Jackson Laboratories
stock No 000664) or Nst1-Cre (piriform recordings, MMRRC
stock No 030648-UCD) were anesthetized by intraperitoneal
ketamine-xylazine injection. The skin above the target site was
numbed by lidocaine injection and retracted to expose the skull.
For cranial window recordings in neocortex, the mouse was
injected with an adeno-associated virus driving the expression of
GCaMP6s under the synapsin promoter (AAV5.Syn.GCaMP6s),
similar to procedures in [78]. The coordinates of the injection
targeted the hindlimb somatosensory cortex, 0.2 mm posterior to
bregma and 1.5 mm lateral to the midline, at a depth of 300 μm
Miniature Multiphoton Microscopes 207
Fig. 10 Mouse attachment. (a) The CIFB is attached to a coupling objective on the proximal end and the
2P-FCM on the distal end. (b) 2P-FCM is attached to the permanent baseplate on the mouse with a quarter-
turn
3.4 Image The images from the 2P-FCM show a honeycomb pixelation pat-
Processing tern due to the packing of the cores of the CIFB. Several methods
have been described to de-pixelate images from CIFBs [53, 80,
81]. The simplest methods involve low-pass filtering with either a
blurring function [82] or masking the image in the frequency
domain [83]. However, two-photon imaging through a CIFB has
the additional complication of the non-uniformity of the fiber
cores. Each core is assumed to have a unique sensitivity, due to
the variability in diameter, shape, NA, and amount of cladding
between adjacent cores. This manifests as discrete variations in
image intensity across the FOV.
This was addressed by programmatically dividing out the sensi-
tivity of each core and interpolating the core values to remove the
pixelation pattern. A flat map of the full field CIFB fiber-cores was
taken by imaging a fluorescent test slide (Chroma Technologies)
with the 2P-FCM, with an example shown in Fig. 11a. The flat-
map stores the centroid coordinates of the cores and their
corresponding sensitivity. The processing was performed with cus-
tom software (Matlab, Mathworks). Each image to be analyzed was
registered to the flat-map, which allowed the identification of the
cores. An example of a raw pixelated image is shown in Fig. 11b.
The relative sensitivity of each core was compensated by dividing by
the flat map values. The honeycomb pattern was eliminated by
using the nearest neighbor interpolation method [84]. A Savitsky-
Golay filter was used to reduce the added single-pixel noise intro-
duced by the core multiplication factor during flat normalization.
Miniature Multiphoton Microscopes 209
Fig. 11 Image processing of two-photon imaging through CIFB. (a) Flat-field map showing non-uniformity due
to differences in sensitivity in individual cores in the CIFB. (b) Unprocessed image of cells in mouse cortex with
fiber-pixelation. (c) Post-processed image after fiber-cores were corrected with flat-field mask and re-gridded
into a typical square pattern. Grid lines added to emphasize pixels
3.5 Testing and The lateral and axial resolution and the axial focusing range of the
Calibration of 2P-FCM can be tested by imaging 2-μm diameter yellow-green
Resolution, fluorescent micro-beads embedded in clear agarose. The lateral
Magnification, and resolution is fundamentally limited by the average spacing of the
Axial Scan Range fiber cores in the CIFB. With the core-to-core spacing of ~4.5 μm
and 2P-FCM magnification factor of 0.4, the theoretical lateral
sampling at the object is ~1.8 μm.
The beads are imaged with the 2P-FCM at sequential focal
planes by tuning the EWTL focus in discrete steps to obtain a
Z-stack. The images are processed to remove the fiber-pixelation
pattern as described in the previous section. Figure 12 shows a
comparison of processed images of the beads imaged with a 20
0.75 NA objective and the 2P-FCM. The average lateral and axial
line profiles of 5 beads measured from different focus positions are
fit to a Gaussian function. The axial bead size measured by the 20
objective is 4.5-μm FWHM and with the 2P-FCM it is 9.9-μm
FWHM. The lateral bead size measured by the 20 objective is
1.7-μm FWHM (dotted grey line) and with the 2P-FCM it is 2.6-μ
210 Baris N. Ozbay et al.
Fig. 12 Axial and lateral resolution tested by imaging fluorescent beads with a 20x Olympus objective (dashed
lines) or with the 2P-FCM (solid lines). Reprinted from [92]
Fig. 13 Measurement of axial scan range. (a) Side-projection of ~2-μm diameter fluorescent beads
suspended in clear agarose and imaged with a 20x 0.8 NA Olympus objective (green) using 910 nm excitation
light and a motorized stage or with the 2P-FCM while varying the EWTL power (red). (b) Predicted scan range
as the EWTL optical power is changed modeled in Zemax (grey line) and Z-positions of measured beads (black
circles)
Table 2
Measured magnification variation through focus
4 Discussion
Fig. 15 3D imaging with 2P-FCM of fixed tissue with oligodendrocytes expressing eGFP. (a) 3D volume
acquired by the 2P-FCM (220 μm dia. Lateral 180 μm axial) with over 200 cells in the image. (b) Processed
image of a single slice in the stack after filtering to remove pixelation pattern. (Reprinted from [92])
4.2 2P-FCM Imaging For in vivo 2P-FCM imaging a 3D-printed baseplate is permanently
In Vivo attached to the mouse. During each imaging session, the 2P-FCM
is attached to the baseplate with little force on the skull. A photo of
a baseplate attached to a mouse with a cranial window is shown in
Fig. 18a. Figure 18b shows a photo of a mouse with the 2P-FCM
attached. The CIFB and electrode wire for the EWTL are draped
passively over a horizontal post to reduce the weight on the mouse.
When attached, the mouse is able to move around freely in a small
area (about 1200 square). The movement is somewhat restricted by
the length of the CIFB (1.0 m) and torsional resistance of the
214 Baris N. Ozbay et al.
Fig. 16 Tilted-field imaging enabled by rapid focusing of the EWTL. (a) Maximum intensity projection of fixed
mouse brain tissue expressing GCaMP6s in neurons, acquired by the 2P-FCM. Arrows indicate cell bodies
retained in fields. (b) Side projection of the volume. The same cell bodies are indicated by the arrows. The
planes for the horizontal and angled scan limits are indicated. (c) Tilted field images taken from selected
planes at angles (30,0,30) degrees as indicated in (b). (Reprinted from [92])
CIFB, which prevents the rotation of more than about 180 . It was
found that, after short acclimation (~30 min), mice could traverse
the entire behavioral area and habituated to the restrictions. Future
implementation may include a commutator with rotation encoder
for realignment, which has been shown previously [86].
Figure 19 shows a widefield epifluorescence image of the
region for an implanted mouse through the 2P-FCM. The left
image shows the background fluorescence and vasculature on the
day of implantation. The right image was taken 17 days later
showing that the baseplate is stable and only a slight lateral shift
in alignment is observed.
Miniature Multiphoton Microscopes 215
Fig. 17 Multi-color imaging. (a) Maximum intensity projection of a region of brain tissue acquired with a 20x 0.75
NA objective. Yellow cells are oligodendrocytes (Green and Red), while red-only cells are astrocytes and
oligodendrocytes. Two astrocytes are marked with arrows and are easily identified by the characteristic bushy
morphology. (b) Same tissue imaged with the 2P-FCM, using the same detectors, filters, and excitation wavelength.
Green and red cells are visible in the field, with likely astrocytes marked by arrows. (Modified from [92])
Fig. 18 Mouse 2P-FCM attachment photos. (a) Baseplate implanted on mouse with cranial window. (b) Mouse
behaving with 2P-FCM attached
4.2.1 2P-FCM In Vivo In vivo two-photon imaging of neuronal activity through the
Mouse Imaging Through 2P-FCM was performed in an awake and mobile mouse expressing
Cranial Window GCaMP6s in cortical neurons. The mouse was allowed to wander
freely in a 700 by 1100 plastic cage. The cage was filled with sawdust as
well as food and various novel objects, such as raisins and tissue
paper, to motivate the mouse to explore the environment while
attached to the 2P-FCM. A 3D projection of a Z-stack taken by
Miniature Multiphoton Microscopes 217
tuning the EWTL focus, showing the imaging volume was acquired
in Fig. 21b, c. Bright cell bodies and processes were visible down to
160 μm, corresponding to cortical layers 2/3. Time-courses were
acquired sequentially at three focal planes at z ¼ 50, 95, and
140 μm below the cranial window. At the deepest focal plane
(z ¼ 140-μm), the active regions are round objects ~10–20 μm in
diameter, which are likely cell bodies. At the middle depth of
95 μm, there is a mixture of processes and cell bodies, while closer
to the brain surface, activity appears predominately from processes
as shown in Fig. 21d. Figure 21f shows the time courses of the ΔF/
F signal for the five indicated regions of interest at the three
different depths.
The mouse was recorded with a camera during the imaging
session to correlate the imaging results to the motion of the mouse.
Lateral motion artifacts were present during some of the record-
ings. A motion correction algorithm was used to correct for the
laterally shifting field [87]. The results of the motion correction
indicate a time-averaged motion artifact magnitude of <2 μm, but
had peaks up to 10 μm in some cases. The intensity of bright cells
that did not exhibit fluorescence activity did not vary with the
motion artifacts, so we conclude that the extent of the motion in
the axial-dimension is lower than the depth-of-focus for the
2P-FCM (<10 μm). Overall, these motion artifacts were similar
to those reported in head-fixed 2PE imaging studies [88, 89] and
widefield imaging, which benefits from a much larger depth of field
[4]. The 2P-FCM did not become loose or dislodged during the
imaging sessions, which lasted between 1 and 4 h.
4.2.2 2P-FCM In Vivo The 2P-FCM can additionally be used for imaging in deep brain
Mouse Imaging in Deep regions by coupling to a GRIN lens. The baseplate is attached such
Brain Regions Through that the 2P-FCM is aligned at the center of the GRIN lens and with
GRIN Lens the focus of the 2P-FCM positioned at the image working distance
of the GRIN lens. The 1:1 GRIN relay lens used did not change the
magnification of the 2P-FCM image, although aberrations from
imaging through the GRIN lens reduced the field–of-view (FOV)
in comparison to the cranial window imaging. Additionally, actuat-
ing the EWTL changed the axial focal plane imaged by the 2P-FCM
through the GRIN lens as shown in Fig. 22. As an example of
behaviorally relevant imaging of neuronal activity in deep brain
regions, the 2P-FCM was used to image activity in anterior piriform
cortex in a freely moving female Ntsr1-cre mouse exploring a novel
environment with male bedding. GCaMP6s was expressed virally
using AAV_CWSL.hSyn.DIO.Synaptophysin-GCaMP6s.P2A.
mRuby. Imaging data was processed to remove the fiber bundle
artifact using the custom Matlab software and then analyzed by
manually selecting ROIs where ΔF/F transients exceeded 6 stan-
dard deviations threshold. Behavioral video was analyzed with the
218 Baris N. Ozbay et al.
Fig. 21 Two-photon Ca2+-imaging in an awake and freely moving mouse using the 2P-FCM. (a) Widefield
camera image of the epifluorescence taken through the FCM showing vasculature in the FOV. A black
rectangle indicates the acquisition region for the following images. (b) 3D projection of a Z-stack acquired
by sequentially imaging and focusing through the tissue using the EWTL. (c) Side projection of the same
Z-stack indicating the three depths at which time series were acquired: at 50 μm, 95 μm, and 140 μm depth,
recorded at frame rates of 2.5, 1.3, and 2.0 Hz, respectively. (d) Maximum intensity projections of image
frames that coincide with Ca2 + transients, showing structural changes through the focal depths. Scale bar
is 50 μm. (e) Selected ROIs that contain fluorescence transients that exceed the 6 SD threshold. Time traces
for five representative ROIs are selected for each depth. (f) Detailed time-courses of the ΔF/F signal for the
five indicated regions. (g) All identified transients, aligned by the peak ΔF/F signal, shown in gray with
averaged intensity signal in black. (Modified from [92])
Fig. 22 Behaviorally dependent responses from 2P-FCM recording in piriform cortex in a mouse implanted
with a GRIN lens and GCaMP6s expressing neurons. (a) Still frame of video of mouse behavior environment
containing familiar bedding and novel unfamiliar bedding during freely moving behavior with attached
2P-FCM. (b) Example maximum projections at two separate focal planes, shown with relative distance of
focus from GRIN lens. (c) Distance between the female mouse’s snout and the foreign male mouse’s bedding
versus time. Manual correction was made to marker placement when DeepLabCut misplaced or lost markers.
(d) An example frame of the behavioral video labeled by DeepLabCut. The green marker tracks the mouse’s
body, the blue marker tracks the mouse’s snout, and the red marker is placed on the male bedding. (e)
Summary of activity, showing greater activity in some ROIs as the mouse approaches the novel odor. (f) Max
projection showing six manually selected regions used in the analysis
220 Baris N. Ozbay et al.
5 Materials
Pulse Pre-compensator
l 1x Polarization Maintaining Fiber – Thorlabs Custom Fiber
Patch Cable: PM780-HP, FC/APC connectors both ends,
keyed to slow axis, FT800SM-Blue Tubing, 0.5 m in length
l 1x Input Coupler – Thorlabs ZC618APC-B – Variable Magnifi-
cation and Focal Length
l 1x Output Coupler – Thorlabs F810APC-850 – 7.8 mm 1/e2
beam diameter
l 2x Gratings – Edmund 37–127 – 300 line per mm blazed at
760 nm
l 1x D-mirror – Thorlabs PFD10-03-M01 – 1” Protected Gold
D-Shaped Mirror
l 1x Retro reflector – Thorlabs HRS1015-AG – Hollow Roof
Mirror, Ultrafast Enhanced Silver Coating
l 1x linear translator – OptoSigma TSD-65171SUU (Imperial)
+/ 25 mm travel
l 1x dovetail optical rail
l Additional optomechanics – kinematic mirror mounts, grating
holders.
Objective Adapter to Hold CIFB
l 1x - SMA Fiber Adapter – Thorlabs SM1SMA
l 1x - Base Cage Plate – Newport OC1-BT
l 1x - Z Stage – Newport OC1-TZ
l 1x - XY Mount – Newport OC1-LH-XY
l 1x - Holder for Objective – Newport OC1-LH1-TZ
l 4x - Cage Rod – Thorlabs ER4
Miniature Multiphoton Microscopes 221
6 Notes
6.1 Optimizing One of the challenges is setting up the single-mode fiber and
Alignment into Single grating pair compressor in front of the microscope. In particular,
Mode Fiber aligning the beam through the single-mode fiber can be challeng-
ing. Make sure to terminate the fiber with an APC (angled physical
contact) on both fiber ends. APC termination minimizes back
reflection because the fiber face is cut at an angle as opposed to a
flat surface which can reflect back along the same optical path and
cause the ultrafast laser to stop modelocking. For input coupling
into the single mode fiber, we use a fiber collimator (Thorlabs
ZC618APC-B) that allows for the adjustment of beam diameter
and divergence. Before the input coupler, set up two steering
mirrors in x, y kinematic mounts to optimize for the beam input
position and angle. It is ideal for the last mirror before the input
coupler to be as close to the input coupler as possible so that it
mostly controls the input angle while the first mirror controls the
222 Baris N. Ozbay et al.
Fig. 23 Method for alignment of single mode fiber. Visualization of red back-propagating beam from visual
fault locator with forward propagating beam from femtosecond laser on IR card. Femtosecond laser is incident
on back of IR card and seen as a green spot on the front side
Fig. 24 Photo of grating pair compressor. Labels indicate (a) fiber output coupler
on kinematic mount, (b) D-mirror, (c) retro-reflector roof mirror on kinematic
mount and linear stage, (d) output mirror, (e) kinematic mount with iris for
alignment, (f) first grating, and (g) second grating
6.2 Setup and Alignment steps for the grating pair compressor (shown in Fig. 24)
Alignment of the are as follows:
Grating Pair 1. Adjust the output beam collimator (A) until the output beam
Compressor from the fiber is traveling parallel to one of the hole lines on the
optical table and at a constant height.
2. Insert a D-shaped mirror mount (B) and adjust until the beam
is traveling approximately 90 degrees to the fiber output, again
parallel to a hole line. Finely adjust the position of the mirror
until clipping is minimized on the D-shaped mirror and beam is
again traveling at a constant height and parallel to the hole line.
3. Turn the linear stage knob holding the retro-reflector
(C) forward until it intercepts the incoming beam. Adjust the
position until the beam is reflected at a lower height so it passes
below the D-shaped mirror. Adjust the tilt of the retro-reflector
to move the reflected beam laterally until it is again parallel to
the hole line, ensuring that the reflection is close to perfect in
the lateral dimension with only a change in height.
4. Insert a mirror (D) to intercept the reflected beam after passing
beneath the D-shaped mirror to send to the microscope input.
Again, adjust the mount until the beam is perfectly
perpendicular.
5. Using a final mirror mount, re-align the output beam with the
beam path into the microscope.
224 Baris N. Ozbay et al.
6.3 Optimizing The FCM objective holder, used to align the proximal end of the
Imaging Through CIFB to the microscope focus, is shown in Fig. 25.
Coherent Fiber Bundle 1. Screw in the FCM objective holder into the 2P-LSM using the
appropriate adapter. Adjust the micrometer on the linear stage
(OC1-TZ) to position the surface of the CIFB at the focus of
the objective. Visualizing the fiber surface is easiest when using
widefield epifluorescence. Center the fiber using the xy adjust-
ments on the cage adapter (OC1-LH-XY).
2. If there is dirt on the fiber repeat cleaning with methanol/air
puffs.
Miniature Multiphoton Microscopes 225
Fig. 25 Diagram of FCM objective holder to mount the FCM to any two-photon
laser scanning microscope. The holder screws into the objective turret on the
microscope with the appropriate adapter. A linear stage (OC1-TZ) adjusts the
focus of objective to the end face of the coherent imaging fiber bundle (CIFB). A
cage mount with xy translation (OC1-LH-XY) adjusts the lateral position of the
fiber to center it on the objective field-of-view
6.4 Installing the Optimize 2P-FCM imaging as detailed above. Then perform the
Pedestal for the 2P- following steps to install the FCM over the cranial window in an
FCM on the Cranium anesthetized mouse.
1. Hold the 2P-FCM attached to the baseplate perpendicular to
the cranial window. Bring the 2P-FCM close to the cranial
window and zero the Sutter micromanipulator. Focus on the
GCaMP-labeled neurons.
226 Baris N. Ozbay et al.
7 Conclusions
Acknowledgments
References
1. Denk W, Strickler J, Webb W (1990) imaged in freely moving animals. Proc Natl
Two-photon laser scanning fluorescence Acad Sci 106(46):19557–19562
microscopy. Science 248(4951):73–76 13. Zong W, Wu R, Chen S et al (2021) Miniature
2. Helmchen F, Denk W (2005) Deep tissue two-photon microscopy for enlarged field-of-
two-photon microscopy. Nat Methods 2(12): view, multi-plane and long-term brain imaging.
932–940 Nat Methods 18(1):46–49
3. Bjerre A-S, Palmer LM (2020) Probing cortical 14. Zong W, Obenhaus H, Skytøen E et al (2021)
activity during head-fixed behavior. Front Mol Large-scale two-photon calcium imaging in
Neurosci 13(30) freely moving mice. bioRxiv. https://doi.org/
4. Ghosh KK, Burns LD, Cocker ED et al (2011) 10.1101/2021.09.20.461015
Miniaturized integration of a fluorescence 15. Liang C, Sung K-B, Richards-Kortum RR et al
microscope. Nat Methods 8(10):871–878 (2002) Design of a high-numerical-aperture
5. Cai DJ, Aharoni D, Shuman T et al (2016) A miniature microscope objective for an endo-
shared neural ensemble links distinct contex- scopic fiber confocal reflectance microscope.
tual memories encoded close in time. Nature Appl Opt 41(22):4603–4610
534(7605):115–118 16. Wu Y, Xi J, Cobb MJ et al (2009) Scanning
6. Aharoni D, Khakh BS, Silva AJ et al (2019) All fiber-optic nonlinear endomicroscopy with
the light that we can see: a new era in miniatur- miniature aspherical compound lens and mul-
ized microscopy. Nat Methods 16(1):11–13 timode fiber collector. Opt Lett 34(7):
7. de Groot A, van den Boom BJ, van Genderen 953–955
RM et al (2020) NINscope, a versatile minis- 17. Botcherby EJ, Juskaitis R, Booth MJ et al
cope for multi-region circuit investigations. (2007) Aberration-free optical refocusing in
elife 9:e49987 high numerical aperture microscopy. Opt Lett
8. Yanny K, Antipa N, Liberti W et al (2020) 32(14):2007–2009
Miniscope3D: optimized single-shot miniature 18. Botcherby EJ, Juškaitis R, Booth MJ et al
3D fluorescence microscopy. Light Sci Appl 9: (2008) An optical technique for remote focus-
171 ing in microscopy. Opt Commun 281(4):
9. Skocek O, Nobauer T, Weilguny L et al (2018) 880–887
High-speed volumetric imaging of neuronal 19. Grutzendler J, Yang G, Pan F et al (2011,
activity in freely moving rodents. Nat Methods 2011) Transcranial two-photon imaging of
15(6):429–432 the living mouse brain. Cold Spring Harb Pro-
10. Zong W, Wu R, Li M et al (2017) Fast high- toc (9):pdb.prot065474
resolution miniature two-photon microscopy 20. Isshiki M, Okabe S (2014) Evaluation of cra-
for brain imaging in freely behaving mice. Nat nial window types for in vivo two-photon
Methods 14(7):713–719 imaging of brain microstructures. Microscopy
11. Klioutchnikov A, Wallace DJ, Frosz MH et al (Oxf) 63(1):53–63
(2020) Three-photon head-mounted micro- 21. Holtmaat A, Bonhoeffer T, Chow DK et al
scope for imaging deep cortical layers in freely (2009) Long-term, high-resolution imaging
moving rats. Nat Methods 17(5):509–513 in the mouse neocortex through a chronic cra-
12. Sawinski J, Wallace DJ, Greenberg DS et al nial window. Nat Protoc 4(8):1128–1144
(2009) Visually evoked activity in cortical cells
228 Baris N. Ozbay et al.
22. Shih AY, Mateo C, Drew PJ et al (2012) A 36. Agrawal GP, Potasek MJ (1986) Effect of fre-
polished and reinforced thinned-skull window quency chirping on the performance of optical
for long-term imaging of the mouse brain. communication systems. Opt Lett 11(5):
JoVE 61:3742 318–320
23. Wang T, Ouzounov DG, Wu C et al (2018) 37. Shen X, Kahn JM, Horowitz MA (2005) Com-
Three-photon imaging of mouse brain struc- pensation for multimode fiber dispersion by
ture and function through the intact skull. Nat adaptive optics. Opt Lett 30(22):2985–2987
Methods 15(10):789–792 38. Clark SW, Ilday FÖ, Wise FW (2001) Fiber
24. Goldey GJ, Roumis DK, Glickfeld LL et al delivery of femtosecond pulses from a Ti:sap-
(2014) Removable cranial windows for long- phire laser. Opt Lett 26(17):1320–1322
term imaging in awake mice. Nat Protoc 9(11): 39. Bouwmans G, Luan F, Knight JC et al (2003)
2515–2538 Properties of a hollow-core photonic bandgap
25. Wachowiak M, Economo MN, Diaz-Quesada fiber at 850 nm wavelength. Opt Express
M et al (2013) Optical dissection of odor infor- 11(14):1613–1620
mation processing in vivo using GCaMPs 40. Jabbour JM, Saldua MA, Bixler JN et al (2012)
expressed in specified cell types of the olfactory Confocal Endomicroscopy: instrumentation
bulb. J Neurosci 33(12):5285–5300 and medical applications. Ann Biomed Eng
26. Peron Simon P, Freeman J, Iyer V et al (2015) 40(2):378–397
A cellular resolution map of barrel cortex activ- 41. Engelbrecht CJ, Johnston RS, Seibel EJ et al
ity during tactile behavior. Neuron 86(3): (2008) Ultra-compact fiber-optic two-photon
783–799 microscope for functional fluorescence imag-
27. Andermann ML, Kerlin AM, Reid C (2010) ing in vivo. Opt Express 16(8):5556–5564
Chronic cellular imaging of mouse visual cor- 42. Myaing MT, MacDonald DJ, Li X (2006)
tex during operant behavior and passive view- Fiber-optic scanning two-photon fluorescence
ing. Front Cell Neurosci 4(3) endoscope. Opt Lett 31(8):1076–1078
28. Murray TA, Levene MJ (2012) Singlet gradient 43. Flusberg BA, Jung JC, Cocker ED et al (2005)
index lens for deep in vivo multiphoton micros- In vivo brain imaging using a portable 3.9
copy. J Biomed Opt 17(2):021106 gram two-photon fluorescence microendo-
29. Jung JC, Schnitzer MJ (2003) Multiphoton scope. Opt Lett 30(17):2272–2274
endoscopy. Opt Lett 28(11):902–904 44. Rivera DR, Brown CM, Ouzounov DG et al
30. Knittel J, Schnieder L, Buess G et al (2001) (2011) Compact and flexible raster scanning
Endoscope-compatible confocal microscope multiphoton endoscope capable of imaging
using a gradient index-lens system. Opt Com- unstained tissue. Proc Natl Acad Sci 108(43):
mun 188(5–6):267–273 17598–17603
31. Jung JC, Mehta AD, Aksay E et al (2004) In 45. Qiu Z, Piyawattanamatha W (2017) New
vivo mammalian brain imaging using one- and endoscopic imaging technology based on
two-photon fluorescence microendoscopy. J MEMS sensors and actuators. Micromachines
Neurophysiol 92(5):3121–3133 8(7):210
32. Wang C, Ji N (2013) Characterization and 46. Helmchen F, Fee MS, Tank DW et al (2001) A
improvement of three-dimensional imaging miniature head-mounted two-photon micro-
performance of GRIN-lens-based two-photon scope. Neuron 31(6):903–912
fluorescence endomicroscopes with adaptive 47. Helmchen F, Denk W, Kerr JN (2013) Minia-
optics. Opt Express 21(22):27142–27154 turization of two-photon microscopy for imag-
33. Antonini A, Sattin A, Moroni M et al (2020) ing in freely moving animals. Cold Spring Harb
Extended field-of-view ultrathin microendo- Protoc 2013(10):904–913
scopes for high-resolution two-photon imag- 48. Zhao Y, Nakamura H, Gordon RJ (2010)
ing with minimal invasiveness. elife 9:e58882 Development of a versatile two-photon endo-
34. Barretto RPJ, Schnitzer MJ (2012, 2012) In scope for biological imaging. Biomed Opt
vivo microendoscopy of the hippocampus. Express 1(4):1159–1172
Cold Spring Harb Protoc (10):pdb. 49. Wood HAC, Harrington K, Stone JM et al
prot071472 (2017) Quantitative characterization of endo-
35. Agrawal GP (2000) Nonlinear fiber optics. In: scopic imaging fibers. Opt Express 25(3):
Christiansen PL, Sørensen MP, Scott AC (eds) 1985–1992
Nonlinear science at the Dawn of the 21st 50. Pierce M, Yu D, Richards-Kortum R (2011)
century, Springer, vol 542. Berlin Heidelberg, High-resolution fiber-optic microendoscopy
Berlin, Heidelberg, pp 195–211 for in situ cellular imaging. JoVE 47:2306
Miniature Multiphoton Microscopes 229
51. Cheon GW, Cha J, Kang JU (2014) Spatial optimized dispersion control by reflection
compound imaging for fiber-bundle optic grisms at 800 nm. Opt Express 20(23):
microscopy. In: SPIE BiOS (ed) Optical fibers 25624–25635
and sensors for medical diagnostics and treat- 65. Udovich JA, Kirkpatrick ND, Kano A et al
ment applications, vol XIV, p 893811. https:// (2008) Spectral background and transmission
doi.org/10.1117/12.2038690 characteristics of fiber optic imaging bundles.
52. Han J-H, Yoon SM (2011) Depixelation of Appl Opt 47(25):4560–4568
coherent fiber bundle endoscopy based on 66. Garofalakis A, Kruglik SG, Mansuryan T et al
learning patterns of image prior. Opt Lett (2019) Characterization of a multicore fiber
36(16):3212–3214 image guide for nonlinear endoscopic imaging
53. Shinde A, Matham MV (2014) Pixelate using two-photon fluorescence and second-
removal in an image fiber probe endoscope harmonic generation. J Biomed Opt 24(10):
incorporating comb structure removal meth- 1–12
ods. J Med Imaging Hlth Inform 4(2): 67. Chiu C-P, Chiang T-J, Chen J-K et al (2012)
203–211 Liquid lenses and driving mechanisms: a
54. Steelman ZA, Kim S, Jelly ET et al (2018) review. J Adhes Sci Technol 26(12–17):
Comparison of imaging fiber bundles for 1773–1788
coherence-domain imaging. Appl Opt 57(6): 68. Blum M, Büeler M, Gr€a tzel C et al (2011)
1455–1462 Compact optical design solutions using focus
55. Chen X, Reichenbach KL, Xu C (2008) Exper- tunable lenses. In: SPIE optical systems design,
imental and theoretical analysis of core-to-core Marseille, France, p 81670W. https://doi.org/
coupling on fiber bundle imaging. Opt Express 10.1117/12.897608
16(26):21598–21607 69. Jabbour JM, Malik BH, Olsovsky C et al
56. Xu C, Webb WW (1996) Measurement of (2014) Optical axial scanning in confocal
two-photon excitation cross sections of molec- microscopy using an electrically tunable lens.
ular fluorophores with data from 690 to 1050 Biomed Opt Express 5(2):645–652
nm. J Opt Soc Am B 13(3):481–491 70. Jiang J, Zhang D, Walker S et al (2015) Fast
57. Diaspro A, Chirico G, Collini M (2005) 3-D temporal focusing microscopy using an
Two-photon fluorescence excitation and electrically tunable lens. Opt Express 23(19):
related techniques in biological microscopy. Q 24362–24368
Rev Biophys 38(2):97–166 71. Yang W, Carrillo-Reid L, Bando Y et al (2018)
58. Lefort C, Mansuryan T, Louradour F et al Simultaneous two-photon imaging and
(2011) Pulse compression and fiber delivery two-photon optogenetics of cortical circuits in
of 45 fs Fourier transform limited pulses at three dimensions. elife 7:e32671
830 nm. Opt Lett 36(2):292 72. Ryan DP, Gould EA, Seedorf GJ et al (2017)
59. Treacy E (1969) Optical pulse compression Automatic and adaptive heterogeneous refrac-
with diffraction gratings. IEEE J Quantum tive index compensation for light-sheet micros-
Electron 5(9):454–458 copy. Nat Commun 8(1):612
60. Oberthaler M, Höpfel RA (1993) Special nar- 73. Zhao Y-P, Wang Y (2013) Fundamentals and
rowing of ultrashort laser pulses by self-phase applications of Electrowetting. Rev Adhes
modulation in optical fibers. Appl Phys Lett Adhesives 1(1):114–174
63(8):1017–1019 74. Supekar OD, Zohrabi M, Gopinath JT et al
61. Lelek M, Suran E, Louradour F et al (2007) (2017) Enhanced response time of Electrowet-
Coherent femtosecond pulse shaping for the ting lenses with shaped input voltage functions.
optimization of a non-linear micro-endoscope. Langmuir 33(19):4863–4869
Opt Express 15(16):10154–10162 75. Smith NR, Abeysinghe DC, Haus JW et al
62. Lefort C, Kalashyan M, Ducourthial G et al (2006) Agile wide-angle beam steering with
(2014) Sub-30-fs pulse compression and electrowetting microprisms. Opt Express
pulse shaping at the output of a 2-m-long opti- 14(14):6557–6563
cal fiber in the near-infrared range. J Opt Soc 76. Supekar OD, Ozbay BN, Zohrabi M et al
Am B 31(10):2317–2324 (2017) Two-photon laser scanning microscopy
63. Kim D, Choi H, Yazdanfar S et al (2008) Ultra- with electrowetting-based prism scanning.
fast optical pulse delivery with fibers for non- Biomed Opt Express 8(12):5412–5426
linear microscopy. Microsc Res Tech 71(12): 77. Berge B, Peseux J (2000) Variable focal lens
887–896 controlled by an external voltage: an applica-
64. Kalashyan M, Lefort C, Martı́nez-León L et al tion of electrowetting. Eur Phys J E 3(2):
(2012) Ultrashort pulse fiber delivery with 159–163
230 Baris N. Ozbay et al.
78. Chen T-W, Wardill TJ, Sun Y et al (2013) 85. Ozden I, Lee HM, Sullivan MR et al (2008)
Ultrasensitive fluorescent proteins for imaging Identification and clustering of event patterns
neuronal activity. Nature 499(7458):295–300 from in vivo multiphoton optical recordings of
79. Franklin KBJ, Paxinos G (2013) Paxinos and neuronal ensembles. J Neurophysiol 100(1):
Franklin’s the mouse brain in stereotaxic coor- 495–503
dinates, 4th edn. Academic Press, an imprint of 86. Flusberg BA, Nimmerjahn A, Cocker ED et al
Elsevier, Amsterdam (2008) High-speed, miniaturized fluorescence
80. Lee C-Y, Han J-H (2013) Elimination of hon- microscopy in freely moving mice. Nat Meth-
eycomb patterns in fiber bundle imaging by a ods 5(11):935–938
superimposition method. Opt Lett 38(12): 87. Dubbs A, Guevara J, Yuste R (2016) Moco:
2023–2025 fast motion correction for calcium imaging.
81. Lee C-Y, Han J-H (2013) Integrated spatio- Front Neuroinform 10:6
spectral method for efficiently suppressing 88. Dombeck D, Tank D (2014) Two-photon
honeycomb pattern artifact in imaging fiber imaging of neural activity in awake mobile
bundle microscopy. Opt Commun 306:67–73 mice. Cold Spring Harb Protoc 2014(7):pdb.
82. Göbel W, Kerr JND, Nimmerjahn A et al top081810
(2004) Miniaturized two-photon microscope 89. Dombeck DA, Khabbaz AN, Collman F et al
based on a flexible coherent fiber bundle and (2007) Imaging large-scale neural activity with
a gradient-index lens objective. Opt Lett cellular resolution in awake, Mobile mice. Neu-
29(21):2521–2523 ron 56(1):43–57
83. Winter C, Rupp S, Elter M et al (2006) Auto- 90. Mathis A, Mamidanna P, Cury KM et al (2018)
matic adaptive enhancement for images DeepLabCut: markerless pose estimation of
obtained with fiberscopic endoscopes. IEEE user-defined body parts with deep learning.
Trans Biomed Eng 53(10):2035–2046 Nat Neurosci 21(9):1281–1289
84. Rupp S, Winter C, Elter M Evaluation of 91. Ziv Y, Burns LD, Cocker ED et al (2013)
spatial interpolation strategies for the removal Long-term dynamics of CA1 hippocampal
of comb-structure in fiber-optic images. In: place codes. Nat Neurosci 16(3):264–266
2009 annual international conference of the 92. Ozbay BN, Futia GL, Ma M et al (2018) Three
IEEE Engineering in Medicine and Biology dimensional two-photon brain imaging in
Society, 3–6 Sept. 2009, pp 3677–3680. freely moving mice using a miniature fiber cou-
https://doi.org/10.1109/IEMBS.2009. pled microscope with active axial-scanning. Sci
5334719 Rep 8(1):8108
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International
License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing, adaptation, distribution
and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative Commons license,
unless indicated otherwise in a credit line to the material. If material is not included in the chapter’s Creative
Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use,
you will need to obtain permission directly from the copyright holder.
Chapter 8
Abstract
Light-sheet microscopy is a powerful method for imaging small translucent samples in vivo, owing to its
unique combination of fast imaging speeds, large field of view, and low phototoxicity. This chapter briefly
reviews state-of-the-art technology for variations of light-sheet microscopy. We review recent examples of
optogenetics in combination with light-sheet microscopy and discuss some current bottlenecks and
horizons of light sheet in all-optical physiology. We describe how 3-dimensional optogenetics can be
added to an home-built light-sheet microscope, including technical notes about choices in microscope
configuration to consider depending on the time and length scales of interest.
Larval fish, flies, and worms are popular model organisms in devel-
opmental biology [1] and, increasingly, in systems neuroscience
[2–4]. Optical translucency make these organisms well-suited to
visualize physiological functions using high-resolution fluorescence
imaging with sub-cellular spatial resolution. Small size and their
ability to thrive when immersed in water make it possible even to
image embryonic development and behaviors in toto over hours or
days [5].
Light-sheet microscopy, also known as Selective Plane Illumi-
nation Microscopy (SPIM), has emerged as the method of choice
for imaging smaller organisms, offering a number of advantages
over point-scanning microscopy in speed, accessible volume, and
phototoxicity. The light-sheet revival over the last two decades is
tightly associated with important milestones in live tissue imaging,
including the iconic example of whole-brain imaging in larval zeb-
rafish (Danio rerio). Launched by early examples of calcium imag-
ing of fictive activity [6], several research groups worldwide now
Eirini Papagiakoumou (ed.), All-Optical Methods to Study Neuronal Function, Neuromethods, vol. 191,
https://doi.org/10.1007/978-1-0716-2764-8_8, © The Author(s) 2023
231
232 Laura Maddalena et al.
routinely record calcium activity from the nearly 105 neurons of the
young awake, behaving zebrafish. The resulting avalanche of data is
beginning to lead to new insights about the communication
between different brain areas (see [3, 7] for recent reviews).
A natural extension to such imaging studies is the integration of
optical methods for perturbation, such as optogenetics [8, 9],
optopharmacology [10], and cell ablation [11, 12]. The expanding
toolkit of molecular probes offers many optogenetic actuators to
remotely activate or inactivate cellular processes. In the context of
controlling neural activity, this progress in engineering molecular
probes, together with the development of suitable optical methods
[13], makes possible to photostimulate action potentials in cells
expressing photosensitive channels and read out the affected neural
activity using fluorescent reporters for calcium [14] or voltage
[15]. Many of these tools have been developed into transgenic
animal strains [16, 17]. Early demonstrations of optogenetic
manipulations in combination with light-sheet microscopy include
optogenetic control over a variety of physiological phenomena,
especially in larval zebrafish, from the beating of the heart [18] to
cellular control of reflexive behaviors [19].
In this chapter, we describe how optogenetics can be added to a
home-built microscope inspired by the open-source project Open-
Spim [20]. As an example, we describe in detail a microscope
configuration suitable for cellular or sub-cellular optogenetics in
larval zebrafish. The method involves adding two-photon photo-
stimulation shaped by computer-generated holography (2P-CGH).
The stimulation module exploits the high numerical aperture
(NA) detection objective of the light-sheet microscope to simulta-
neously excite multifocal points targeted either to sub-cellular
regions or to multiple somata. The light-sheet module provides
flexibility to readout neural activity in tiny organisms from small
volumes to whole brain. Volumetric imaging is achieved with an
electrically tunable lens, allowing independent control of imaging
depth without moving the detection objective and consequently
the axial location of the stimulation foci.
We provide technical notes on optical alignment, alternative
configurations for different applications, and limitations and chal-
lenges of combining optogenetics with light-sheet microscopy.
Finally, we offer some perspectives on extending all-optical physiol-
ogy to higher spatial resolution in vivo.
Fig. 1 Selection of light-sheet microscope configurations used for whole-brain imaging in larval zebrafish.
(a) Fluorescence is collected along a sheet of light formed by side-ways illumination. (b) In variations of multi-
view light-sheet microscopy, additional objectives illuminate or collect fluorescence simultaneously allowing
for either greater uniformity of illumination or multiplexing image formation from different angles. (c) In
variations of single-objective LS, a tilted sheet is swept laterally across the sample while collecting a tilted
epifluorescence image. (d) Selective volume illumination microscopy is a hybrid of light-sheet illumination and
extended depth of field detection
Multiview Illumination
In the case of multiple illumination sheets, each sheet needs to
cover only half of the total FOV, so a lower NA sheet can be used
to better preserve optical sectioning. For instance, the IsoView [28]
light-sheet microscope employs two different DSLM geometries to
simultaneously illuminate the sample from two opposite sides,
collecting the view with two cameras (Fig. 1b). In this method,
combining the overlapping images from multiple angles, it is possi-
ble to achieve isotropic spatial resolution [28]. This parallel excita-
tion also represents a robust solution against sample opacity.
Moreover, employing two sets of galvanometric mirrors in each
illumination arm (one for scanning, one for correcting incidence
angle on the sample) allows to employ online optimization algo-
rithms [29] to partially correct for low-order sample-induced aber-
rations. These improvements come at a cost in terms of both
hardware (the number of parts and alignment difficulty) and soft-
ware complexity. Additionally, because every image is collected
twice, the amount of data collected necessitates high-end storage
capabilities and lengthy analysis pipelines in order to properly fuse
the views into a final volume. An added benefit provided by this
geometry is the sample that can be left stationary, while the scan-
ning is performed by the galvanometric mirrors (to move the
excitation profile in 3D) coupled with piezo motors that keep the
detection objective focused on the illumination plane.
Table 1
Beam properties. Bessel beam characteristics depend on the geometry of
an annular mask, placed in a plane conjugated with the back aperture of the
excitation objective. Parameters: e = pixel size; NA = objective numerical
aperture; M = magnification; n = refractive index; λ = wavelength; w =
annulus width; J1 = Bessel function of the first kind; α, α′ = constants
proportional to outer and inner annulus radius, respectively
1.2.2 Type of The experimenter has many options to add photostimulation optics
Photostimulation to a light-sheet microscope, including the variety of methods dis-
cussed in Chapters 1, 3–5, and 11 of this book. Both scanning and
parallel approaches to photostimulation can be applied to small,
translucent samples. In the first case, resonant scanners or galvano-
metric mirrors steer a focused beam across multiple regions of
interest (ROIs), whereas, in the latter case, all the ROIs are illumi-
nated simultaneously by using computer-generated holograms
(CGHs) projected through spatial light modulators (SLMs).
The same excitation strategies applied in living animals require
increased optical sectioning and penetration depth, both provided
by two-photon (2P) illumination. For example, Dal Maschio et al.
[48] have integrated a 2P-CGH module with a two-photon-scan-
ning microscope, generating an instrument capable of identifying
behavior-related neural circuits in living zebrafish larvae. The stim-
ulation is targeted to single soma with a diameter of 6 μm and an
axial resolution of 9 μm over a volume of 160×80×32 μm.
Comparable lateral and axial resolutions for circuit optogenetics are
achieved by McRaven and colleagues [49], in their 2P-CGH setup
coupled to a 2P scanning microscope with remote focusing, to
discover cellular-level motifs in awake zebrafish embryos. On the
other hand, De Medeiros et al. [12] have combined a scanning unit
with a multiview light-sheet microscope. This is a flexible
Optogenetics and SPIM 239
2 Materials
Fig. 2 Example configuration of light-sheet microscope with optogenetics (a) The system consists of a light-
sheet imaging module (left, blue lines) and an optogenetics module (right, red lines) that share a common light
path through the high-NA detection objective (DO). In the imaging module, a continuous-wave visible
wavelength laser is shaped in a Bessel beam by an axicon (A) and scanned onto the sample through galvo
mirrors G1 and G2 and excitation objective (EO). Fluorescence collected through the detection objective (DO) is
transmitted through a low-pass dichroic mirror (DM) and imaged onto a sCMOS camera by an electrically
tunable lens (ETL).The fluorescent signal is recorded with the confocal slit detection approach schematically
shown in panel b. In the 2P-DH module, a Ti:Sa pulsed laser beam is magnified through a telescope (L1, L2),
impinges on the SLM, placed in a plane where the wavefront is flat. Lenses (L3, L4) relay the wavefront from
242 Laura Maddalena et al.
Fig. 2 (continued) the SLM to the pupil of the DO. In the focal plane of lens L3, the inverse pinhole (IP) blocks
the zero-order diffraction spot. (b) Left: Scheme of confocal slit detection. An active window (yellow line) on
the sCMOS camera is rolled synchronously with the scanning of the Bessel beam illumination (cyan line). In
this way, we minimize the detection of fluorescence excited by the side lobes of the Bessel beam (light cyan
halo). Right: Bessel beam projected into uniform fluorescent solution. The zoom in on the middle region of FOV
shows the central lobe of the Bessel beam as it appears with confocal slit detection
Optogenetics and SPIM 243
to the pupil of the DO: here, the Fourier lens L3 (focal 250 mm)
and lens L4 (focal 500 mm) magnify the beam to fill the back
aperture of the objective. In the focal plane of lens L3, an inverse
pinhole (IP, diameter 1.4 mm) blocks the zero-order
diffraction spot.
The light paths of the light-sheet microscope and the 2P-CGH
module join at a dichroic mirror (DM; Semrock, FF01-720/SP-
25), with a cutoff wavelength at 720 nm. This mirror reflects the
NIR stimulation light to the sample and transmits the fluorescent
readout to the sCMOS. Two-photon-excited (TPE) epifluores-
cence can be recorded, and this is useful to characterize the photo-
stimulation beam as described in Subheading 3.1.
The choice of focal lengths of lenses L3 and L4 is important in
the design of the experiment. If the SLM image at the objective
back aperture is smaller than the aperture itself, the NA of the
objective is not fully exploited. This might be an intentional choice.
Otherwise, if the SLM image is larger, some stimulation light gets
lost. The best option is to choose a telescope magnification that
matches the dimension of the SLM image at the back aperture to
the aperture itself. In this case, the lateral resolution is only limited
by the objective NA.
The pixelated structure of the SLM chip introduces in the
reconstructed hologram a zero-order diffraction spot that appears
as a bright spot in the center next to the hologram. To separate the
hologram from the zero-order illumination, we can either proceed
algorithmically, or we can block the zero-order illumination physi-
cally. Since the result given by the second method can strongly
degrade the quality of the image, it is preferable to implement a
combination of the two approaches to preserve the image quality.
The hologram can be displaced from the zero order algorithmically
by introducing a constant defocus. Once CGH and zero order lay at
different depths, the zero-order beam is blocked by means of an
inverse pinhole without impairing the quality of the CGH.
In high-resolution applications, the spatial accuracy of photo-
stimulation is paramount. As reported in [57], the precision to
address a CGH to a specific target depends on the number of pixels
in the SLM and gray scale values available. Several companies
produce SLM with over 1000 pixels in the shorter axis, which
provide the spatial accuracy required for sub-cellular manipulations
and also a high number of degrees of freedom if the SLM is used as
an adaptive element to correct for high-order aberrations. The
accessible lateral and axial FOV for a CGH is inversely proportional
to the SLM pixel size. However, fairly large pixel size (in the order
of 10 μm) is preferred because, assuming a constant inter-pixel gap,
the fill factor increases with the pixel size. A larger pixel size also
reduces the cross-talk between pixels. Cross-talk acts as a low-pass
filter on the CGH, and it is due to fringing field effect that causes
gradual voltage changes across the border of neighboring pixels and
244 Laura Maddalena et al.
2.3 Sample In SPIM, samples are usually mounted in tubes made of fluorinated
Preparation ethylene propylene (FEP), a plastic with refractive index similar to
water. The tube is filled with a solution of water and low-melting-
temperature agarose (1.5–2%) for short-term imaging (1-3
hours). For larval zebrafish, these conditions ensure stability of
the sample while maintaining good physiological conditions given
the time frame of the experiment. However, as reported in [59] for
longer experiments (over 1 day), it is recommended to use lower
agarose percentage (0.1%) or methylcellulose solutions (3%) to
ensure stability but also proper growth of the sample especially
between 24–72 h post-fertilization.
3 Methods
Fig. 3 Main steps of system alignment. (a) Example of good planar alignment. This image shows an overlay of
the Bessel profile at different positions across the planar FOV. Inset shows a shallow Bessel beam with its side
lobes. (b) Example of misaligned beam across the planar FOV. (c) Affine transformation between light-sheet
and 2P-CGH module: (1) 3D plot of the coordinates fed into the algorithm to generate a point cloud CGH of
random points spanning over a 3D volume. Those coordinates are defined in the coordinate system x′y′z′ of
the 2P-CGH module. (2) 3D plot of the coordinates measured on the TPE image of the CGH. Those coordinates
are defined in the coordinate system xyz of the light-sheet module. (3) Result of the affine transformation
between 2P-CGH coordinates and light-sheet coordinates
246 Laura Maddalena et al.
3.1.2 Align the 2P-CGH The 2P-CGH module is aligned by projecting the zero-order dif-
Module fraction beam into the same fluorescent solution used for the light-
sheet alignment:
1. Align the NIR beam to impinge on the SLM with a small angle
and to uniformly illuminate the whole SLM chip.
2. Check that the reflection of the zero-order beam from the SLM
is well-separated from the incoming beam and that these beams
have the same height.
3. Center the zero-order diffracted beam onto the irises placed in
front of the downstream optics.
4. Check the position of the beam at the sample using the sCMOS
camera. The beam must appear as a bright spot in the middle of
the camera FOV. If this spot appears far away from the center of
the camera, adjustments in the upstream optics are needed.
5. Install the inverse pinhole in the optical path and visualize a
CGH projected into the sample.
6. Calibrate the 2P-CGH module with the ETL voltages. For
this, load a series of single-point CGHs with different
z positions onto the SLM. For each CGH, adjust the voltage
on the ETL and acquire an image of the TPE fluorescence
produced by the CGH. Record the second moment of the
spot intensity as a function of the voltage on each image series.
The voltage corresponding to the minimum second moment is
considered the right voltage to have the point in focus. This
procedure links the z position of a CGH to the optimal voltage
needed for an image in focus.
3.1.3 Registration To obtain good fidelity between the stimulation target and the
Between Light-Sheet holographic illumination, it is critical that the light-sheet module
Module and 2P-CGH and 2P-CGH module share a common coordinate system. This is
Module accomplished by first calibrating the scan volume of the LS and
Optogenetics and SPIM 247
3.2 Workflow of The general workflow for all-optical physiology experiment using
Light-Sheet the system described above would be similar to other methods
Optogenetics described in this volume. After alignment of the system, it is typical
Experiment to first acquire a baseline fluorescence image or a 3D movie to
characterize anatomical structure and possibly baseline activity.
Then, the ROIs to be stimulated are selected. The criteria to choose
the ROIs from the baseline image are set by the experimenter and
fed into an algorithm for targeting light, e.g., CGH calculation.
Subsequently, we program a pulse sequence encoding for the CGH
where the stimulation light is temporally gated by an external
shutter. Afterward, the fluorescence signal is recorded over time.
3.2.1 Imaging Larval After microscope alignment, mount a zebrafish in FEP tubing. Use
Zebrafish brightfield illumination to orient and position the sample appropri-
ately. Obtain a baseline fluorescence image. A high-resolution vol-
ume is recommended to help with brain registration later in image
analysis. Based on anatomical or functional criteria, select regions of
248 Laura Maddalena et al.
3.2.2 Analysis of Large- Figure 4 shows the analysis of a representative data set measuring
Scale Ca2+ Data Set spontaneous neural activity in a 5 days past fertilization (dpf)
zebrafish larvae expressing nuclear localized GCaMP6s in neurons,
as acquired by volumetric calcium imaging on the described Bessel-
beam light-sheet microscope. Fluorescence was captured in 20 vol-
ume sections of the forebrain, spaced 8 μm apart, with an acquisi-
tion rate of 0.67 Hz. Motion correction and the extraction of
activity traces from individual neurons were implemented with the
open-source calcium image analysis package CaImAn [60]. Motion
artifacts were corrected through piecewise-rigid registration. Active
neurons were then detected in the motion-corrected data through
constrained non-negative matrix factorization (CNMF). To initial-
ize CNMF, the images were first filtered with a Gaussian kernel.
Thereafter, the Pearson correlation with neighboring pixels and the
peak-to-noise ratio was calculated for each pixel (Fig. 4a). Local
maxima in the pointwise product of the peak-to-noise ratio image
and correlation image were used as initialization positions. With
CNMF, the contours of found neurons (Fig. 4b) and activity traces
(Fig. 4c) were then extracted. In total, 2077 active neurons were
detected in the imaged brain volume (Fig. 4d,e).
3.3 Choice of CGH The 2P-CGH module shown in Fig. 2a is based on Fourier holog-
Algorithm raphy and requires an appropriate algorithm to calculate the desired
phase hologram. The aim of CGH algorithms is to retrieve the
phase mask, namely the phase of the hologram field Uh, to address
the SLM by knowing the field Uo of the target at the image plane,
where these fields are one and the Fourier transform of the other.
To generate a hologram of ROIs distributed in 3D, we compute a
3D hologram corresponding to a field Uo defined at different
depths z as schematically described in Fig. 5a. Optically, the Fourier
transform of the field Uh is realized through the objective lens when
the SLM is conjugated to the objective back aperture through a
telescope (L3 and L4 in Fig. 2a).
The choice of the algorithm to calculate CGHs depends on
several considerations. Ideally, the algorithm provides high diffrac-
tion efficiency, uniformity over the volume of interest, and accuracy
between the target and the reconstructed hologram. The algorithm
should also be fast given the real-time nature of optogenetics
experiments. The shape and dimensions of the target photostimu-
lation pattern also influence the choice of CGH algorithm. When
the target object has a complex and extended lateral shape (>
1 μm), image-based algorithms are employed. They enable the
generation of extremely complex illumination patterns in very
short times; however, they are limited to illumination light focused
Optogenetics and SPIM 249
Fig. 4 Analysis of light sheet calcium imaging data from the larval zebrafish forebrain with the open-source
calcium image analysis package CaImAn. (a) Images were obtained from the forebrain of a larval zebrafish
(5 dpf) expressing nuclear localized GCaMP6s. A total of 20 volume sections were imaged with Bessel beam
light-sheet microscopy at an acquisition rate of 0.67 Hz. To detect neurons, CNMF was initialized by filtering
the motion corrected timeseries data with a Gaussian kernel and calculating the peak-to-noise ratio and
Pearson correlation with 4 nearest neighboring pixels. Local maxima in the point-wise product of the peak-to-
noise ratio image and correlation image were then used as initialization positions. (b) Contours of neurons
250 Laura Maddalena et al.
Fig. 4 (continued) extracted with CNMF in a single section. (c) The fluorescence (△F/F) signals of 10 randomly
selected neurons. (d) Maximum intensity projection of the measured volume. (e) Activity map of 2077
fluorescence traces from neurons found in all sections clustered by correlation coefficient
Optogenetics and SPIM 251
Fig. 5 CGH approach. (a) Optical transformation needed to project a 3D digital hologram. Uh represents the
spatial Fourier transform of the desired pattern Uozi defined at different depths. f is the focal length of the lens
(middle element). (b) Scheme of algorithms for extended CGH. Phase masks producing different features at
focal planes I, II, III are combined by superposition principle at the SLM. Each of the three phase masks is
calculated based on the corresponding image at the focal plane. (c) Scheme of algorithms for point-cloud
CGH. Phase masks producing different features at focal planes I, II, III are combined by superposition principle
at the SLM. Each of the three phase masks is calculated based on the coordinates xyz of the spots at the focal
plane. (d) 3D view of image-based CGH of a grid 160 × 160 × 100 μm. (e) 3D view of point-cloud CGH of the
same grid
3.4 Effect of All algorithms previously described compute CGHs based on the
Aberrations on CGH assumption that the sample has a uniform index of refraction. This
assumption is rarely true in neuroscience since brain tissue is gener-
ally optically turbid and non-uniform. Optical inhomogeneities of
samples cause a distortion of the stimulation pattern known as
252 Laura Maddalena et al.
Table 2
Summary of computer-generated holography algorithms
4 Summary
Fig. 6 Effect of coma on CGH: (a) TPE fluorescence induced onto a rhodamine slide by a one-point hologram
with progressively increasing horizontal coma coefficient. Subscripts on the images indicate the
corresponding peak-to-valley (PV) coefficient in units of the wavelength of horizontal coma. The additional
coma phase introduces a dramatic drop in the fluorescence intensity compared to the case without aberra-
tions (0 μm PV coefficient). (b) Bar plot showing the maximum intensity of each aliased fluorescent spot in
a. Intensities are normalized to the maximum intensity of the aberration-free spot (0 μm PV coefficient). The
maximum intensity is reduced below 50% compared to the intensity of the aberration-free spot. (c) Same
fluorescence images showed in (a) where the aberrated spots intensity is enhanced by a constant factor. (d)
Maximum intensity projection of the fluorescence images in (b) showing the position of each point across the
camera chip. The aberration besides the loss of intensity introduces a loss of the CGH fidelity
254 Laura Maddalena et al.
5 Notes
Note 1 Linear affine transforms will account for all linear transfor-
mations between coordinates, including lateral and axial shifts,
magnification mismatches (independently for all axes), rigid rota-
tions, and coordinate shearing. Non-linear distortions, such as
barrel or pincushion distortion, would require a more complex
non-linear transform. Given that the FOV is limited to hundreds
of microns, non-linear effects rarely appear in a well-designed sys-
tem, and a linear transform provides sufficient accuracy even for
photostimulation of sub-micron structures. The linear affine trans-
form matrix between homogeneous coordinates is represented by a
4x4 matrix [67]; however, it is determined by only 12 values as the
last row in the matrix is always defined as {0, 0, 0, 1}. Hence, the
matrix can be determined through least square minimization
between vectors Xi of the image coordinates and Xi vectors of
12 SLM coordinates. The accuracy of the matrix estimation is
improved using more than 12 coordinates see Subheading 3.1.3.
Problem Description
The mathematical approach to calculate a CGH differs based on
whether we deal with image-based or point-cloud CGHs. To target
Optogenetics and SPIM 255
complex and extended shapes spanning over few focal planes, the
electric field of the desired pattern per each focal plane is modeled
as
U o ðξ,η,zÞ = uo ðξ,ηÞe iϕz , ð2Þ
where uo is the square root of the desired intensity at depth z and ϕz
is a random phase. The hologram field is defined as the fast Fourier
transform (FFT) of the field Uo at each focal plane. Then the phase
delay between the fields at different depths is optimized to get an
interference pattern at the SLM plane as close as possible to the
target illumination pattern. On the other hand for diffraction-
limited spots arbitrarily distributed in the 3D FOV, the hologram
field is defined as the wavefront generated by the coordinates x, y,
z of the single spots in the focal planes. Hence, the field at the SLM
is calculated as the superposition of the fields generating each
single spot:
P
N - ið2π 0 0 zn π 02 02
λf ðx n x þy n y Þþ 2 2 ðx þy Þþϕn Þ
Uh = uo,n e λ f , ð3Þ
n=1
where uo,n is the square root of the desired intensity of each spot at
the focal plane, f is the focal length of the optical system, λ is the
wavelength of the light source, xn, yn, zn are the coordinates of the
n desired spots in 3D space, x′ and y′ are the coordinates of each
SLM pixel, and ϕn is a constant term. This field can be related to the
field at the focal plane via a discrete Fourier transform (DFT) of the
single spots. Also in this case, the aim is to optimize the phase delay
between the fields generating each single spot.
• Gerchberg–Saxton (GS)
The electric field Uo of the desired pattern is modeled as in
the previous algorithm. Then the hologram is calculated as the
inverse Fourier transform of the object field Uo. In the resulting
hologram field, the phase is retained, while the amplitude is
substituted with a Gaussian amplitude. This new field undergoes
another Fourier transform to produce the object field at the
focal plane. The pattern produced at the focal plane is similar
to the desired one, but it has a decreased efficiency; hence, the
phase is kept, while the amplitude is replaced with the one of the
desired patterns. At this stage, the inverse Fourier transform of
the field is calculated to retrieve the phase mask at the SLM
plane. This procedure is repeated iteratively until the algorithm
converges to phase mask that produces the best intensity pat-
tern. This algorithm can be implemented for 3D holograms
including in the calculation the propagation of the fields to
different depth levels according to the diffraction theory [68].
Recent Algorithms
• DeepCGH
The DeepCGH algorithm [65] uses convolutional neural
networks (CNNs) with unsupervised learning to compute 3D
holograms with the image-based approach. The distribution of
amplitudes A(x, y, z) of the 3D target pattern at the focal planes
is the input of the CNN. The CNN provides in output an
estimate of the complex field P(x, y, z = 0) at depth z = 0. This
field backpropagated to the SLM plane via 2D inverse Fourier
258 Laura Maddalena et al.
Acknowledgements
References
1. Lemon WC, Pulver SR, Höckendorf B, 8. Fenno L, Yizhar O, Deisseroth K (2011) The
McDole K, Branson K, Freeman J, Keller PJ development and application of optogenetics.
(2015) Whole-central nervous system func- Annu Rev Neurosci 34:389–412
tional imaging in larval Drosophila. Nature 9. Johnson HE, Toettcher JE (2018) Illuminat-
Communications 6(Article number 7924) ing developmental biology with cellular opto-
2. Vanwalleghem GC, Ahrens MB, Scott EK genetics. Curr Opin Biotechnol 52:42–48
(2018) Integrative whole-brain neuroscience 10. Kramer RH, Mourot A, Adesnik H (2013)
in larval zebrafish. Curr Opin Neurobiol 50: Optogenetic pharmacology for control of
136–145 native neuronal signaling proteins. Nature
3. Loring MD, Thomson EE, Naumann EA Neuroscience 16(7):816–823
(2020) Whole-brain interactions underlying 11. Hill RA, Damisah EC, Chen F, Kwan AC,
zebrafish behavior. Curr Opin Neurobiol 65: Grutzendler J (2017) Targeted two-photon
88–99 chemical apoptotic ablation of defined cell
4. Eschbach C, Zlatic M (2020) Useful types in vivo. Nature Communications 8:
road maps: studying Drosophila larva’s central 15837
nervous system with the help of connectomics. 12. de Medeiros G, Kromm D, Balazs B, Norlin N,
Curr Opin Neurobiol 65:129–137 Günther S, Izquierdo E, Ronchi P, Komoto S,
5. Keller PJ, Ahrens MB (2015) Visualizing Krzic U, Schwab Y, Peri F, de Renzis S,
whole-brain activity and development at the Leptin M, Rauzi M, Hufnagel L (2020) Cell
single-cell level using light-sheet microscopy. and tissue manipulation with ultrashort infra-
Neuron 85:462–483 red laser pulses in light-sheet microscopy. Sci-
6. Ahrens MB, Orger MB, Robson DN, Li JM, entific Reports 10:1–12
Keller PJ (2013) Whole-brain functional imag- 13. Emiliani V, Cohen AE, Deisseroth K, H€ausser
ing at cellular resolution using light-sheet M (2015) Symposium all-optical interrogation
microscopy. Nature Methods 10(5):413–420 of neural circuits. J Neurosci 35(41):
7. Lovett-Barron M (2021) Learning-dependent 13917–13926
neuronal activity across the larval zebrafish 14. Knöpfel T (2012) Genetically encoded optical
brain. Curr Opin Neurobiol 67:42–49 indicators for the analysis of neuronal circuits.
Nat Rev Neurosci 13:687–700
Optogenetics and SPIM 259
15. St-Pierre F, Marshall JD, Yang Y, Gong Y, isotropic spatial resolution. Nature Methods
Schnitzer MJ, Lin MZ (2014) High-fidelity 12:1171–1178
optical reporting of neuronal electrical activity 29. Royer LA, Lemon WC, Chhetri RK, Keller PJ
with an ultrafast fluorescent voltage sensor. (2018) A practical guide to adaptive light-sheet
Nature Neuroscience 17(6):884–889 microscopy. Nature Protocols 13(11):
16. Asakawa K, Kawakami K (2008) Targeted gene 2462–2500
expression by the Gal4-UAS system in zebra- 30. Dunsby C (2008) Optically sectioned imaging
fish. Dev Growth Differ 50:391–399 by oblique plane microscopy. Optics Express
17. Antinucci P, Dumitrescu A, Deleuze C, Morley 16(25):20306–20316
HJ, Leung K, Hagley T, Kubo F, Baier H, 31. Bouchard MB, Voleti V, Mendes CS,
Bianco IH, Wyart C (2020) A calibrated opto- Lacefield C, Grueber WB, Mann RS, Bruno
genetic toolbox of stable zebrafish opsin lines. RM, Hillman EM (2015) Swept confocally-
eLife 9 aligned planar excitation (SCAPE) microscopy
18. Rost BR, Schneider-Warme F, Schmitz D, for high-speed volumetric imaging of behaving
Hegemann P (2017) Optogenetic tools for organisms. Nature Photonics 9:113–119
subcellular applications in neuroscience. Neu- 32. Hillman EM, Voleti V, Li W, Yu H (2019)
ron 96:572–603 Light-sheet microscopy in neuroscience. Annu
19. Pantoja C, Hoagland A, Carroll EC, Karalis V, Rev Neurosci 42:295–313
Conner A, Isacoff EY (2016) Neuromodula- 33. Strack R (2021) Single-objective light sheet
tory regulation of behavioral individuality in microscopy. Nature Methods 18:28
zebrafish. Neuron 91(3):587–601 34. Galland R, Grenci G, Aravind A, Viasnoff V,
20. https://openspim.org/ Studer V, Sibarita JB (2015) 3D high-and
21. Wan Y, McDole K, Keller PJ (2019) Light- super-resolution imaging using single-
sheet microscopy and its potential for under- objective SPIM. Nature Methods 12(7):
standing developmental processes. Annu Rev 641–644
Cell Dev Biol 35:655–681 35. Meddens MBM, Liu S, Finnegan PS, Edwards
22. Power RM, Huisken J (2017) A guide to light- TL, James CD, Lidke KA (2016) Single objec-
sheet fluorescence microscopy for multiscale tive light-sheet microscopy for high-speed
imaging. Nature Methods 14(4):360–373 whole-cell 3D super-resolution. Biomedical
23. Huisken J, Swoger J, Del Bene F, Wittbrodt J, Optics Express 7(6):2219–2236
Stelzer EH (2004) Optical sectioning deep 36. Tomer R, Lovett-Barron M, Kauvar I,
inside live embryos by selective plane illumina- Andalman A, Burns VM, Sankaran S,
tion microscopy. Science 305(5686):56–86 Grosenick L, Broxton M, Yang S, Deisseroth
24. Verveer PJ, Swoger J, Pampaloni F, Greger K, K (2015) SPED light sheet microscopy: fast
Marcello M, Stelzer EH (2007) High- mapping of biological system structure and
resolution three-dimensional imaging of large function. Cell 163(7):1796–1806
specimens with light sheet-based microscopy. 37. Quirin S, Vladimirov N, Yang C-T, Peterka DS,
Nature Methods 4(4):311–313 Yuste R, Ahrens MB (2016) Calcium imaging
25. Keller PJ, Schmidt AD, Wittbrodt J, Stelzer of neural circuits with extended depth-of-field
EH (2008) Reconstruction of zebrafish early light-sheet microscopy. Optics Letters 41(5):
embryonic development by scanned light sheet 855–858
microscopy. Science 322(5904):1065-9 38. Truong TV, Holland DB, Madaan S,
26. Gao L (2015) Extend the field of view of selec- Andreev A, Keomanee-Dizon K, Troll JV,
tive plan illumination microscopy by tiling the Koo DE, McFall-Ngai MJ, Fraser SE (2020)
excitation light sheet. Optics Express 23(5): High-contrast, synchronous volumetric imag-
6102–6111 ing with selective volume illumination micros-
27. Vladimirov N, Wang C, Höckendorf B, copy. Communications Biology 3(1):74
Pujala A, Tanimoto M, Mu Y, Yang CT, Wit- 39. Sapoznik E, Chang BJ, Ju RJ, Welf ES,
tenbach JD, Freeman J, Preibisch S, Broadbent D, Carisey AF, Stehbens SJ, Lee
Koyama M, Keller PJ, Ahrens MB (2018) KM, Marı́n A, Hanker AB, Schmidt JC,
Brain-wide circuit interrogation at the cellular Arteaga CL, Yang B, Kruithoff R, Shepherd
level guided by online analysis of neuronal DP, Millett-Sikking A, York AG, Dean KM,
function. Nature Methods 15:1117–1125 Fiolka R (2020) A single-objective light-sheet
28. Chhetri RK, Amat F, Wan Y, Höckendorf B, microscope with 200 nm-scale resolution.
Lemon WC, Keller PJ (2015) Whole-animal bioRxiv
functional and developmental imaging with
260 Laura Maddalena et al.
40. Mertz J (2011) Optical sectioning microscopy 49. McRaven C, Tanese D, Zhang L, Yang C-T,
with planar or structured illumination. Nature Ahrens MB, Emiliani V, Koyama M, High-
Methods 8:811–819 throughput cellular-resolution synaptic con-
41. Jemielita M, Taormina MJ, Delaurier A, Kim- nectivity mapping in vivo with concurrent
mel CB, Parthasarathy R (2013) Comparing two-photon optogenetics and volumetric
phototoxicity during the development of a zeb- Ca2+ imaging. bioRxiv. https://doi.
rafish craniofacial bone using confocal and light org/10.1101/2020.02.21.959650
sheet fluorescence microscopy techniques. J 50. Ko K, Nig E (2000) Multiphoton microscopy
Biophoton 6:920–928 in life sciences. J Microscopy 200(2):83–104
42. Planchon TA, Gao L, Milkie DE, Davidson 51. Supatto W, Truong TV, Débarre D, Beaure-
MW, Galbraith JA, Galbraith CG, Betzig E paire E (2011) Advances in multiphoton
(2011) Rapid three-dimensional isotropic microscopy for imaging embryos. Curr Opin
imaging of living cells using Bessel beam plane Genet Dev 21:538–548
illumination. Nature Methods 8(5):417–423 52. Truong TV, Supatto W, Koos DS, Choi JM,
43. Gao L, Shao L, Higgins CD, Poulton JS, Fraser SE (2011) Deep and fast live imaging
Peifer M, Davidson MW, Wu X, Goldstein B, with two-photon scanned light-sheet micros-
Betzig E (2012) Noninvasive imaging beyond copy. Nature Methods 8:757–762
the diffraction limit of 3D dynamics in thickly 53. Wolf S, Supatto W, Debrégeas G, Mahou P,
fluorescent specimens. Cell 151(6): Kruglik SG, Sintes JM, Beaurepaire E, Cande-
1370–1385 lier R (2015) Whole-brain functional imaging
44. Fahrbach FO, Gurchenkov V, Alessandri K, with two-photon light-sheet microscopy.
Nassoy P, Rohrbach A (2013) Light-sheet Nature Methods 12:379–380
microscopy in thick media using scanned Bessel 54. Alberto D, Mauro R (2000) Two-photon exci-
beams and two-photon fluorescence excitation. tation of fluorescence for three-dimensional
Optics Express 21(11):13824–13839 optical imaging of biological structures. J
45. Chen BC, Legant WR, Wang K, Shao L, Milkie Photochem Photobiol B Biol 55(1):1–8
DE, Davidson MW, Janetopoulos C, Wu XS, 55. Picot A, Dominguez S, Liu C, Chen IW,
Hammer JA, Liu Z, English BP, Mimori- Tanese D, Ronzitti E, Berto P,
Kiyosue Y, Romero DP, Ritter AT, Papagiakoumou E, Oron D, Tessier G, Forget
Lippincott-Schwartz J, Fritz-Laylin L, Mullins BC, Emiliani V (2018) Temperature rise under
RD, Mitchell DM, Bembenek JN, Reymann two-photon optogenetic brain stimulation.
AC, Böhme R, Grill SW, Wang JT, Cell Reports 24:1243–1253
Seydoux G, Tulu US, Kiehart DP, Betzig E 56. dal Maschio M, Donovan JC, Helmbrecht TO,
(2014) Lattice light-sheet microscopy: Imag- Baier H (2017) Linking neurons to network
ing molecules to embryos at high spatiotempo- function and behavior by two-photon holo-
ral resolution. Science 346(6208). https://doi. graphic optogenetics and volumetric imaging.
org/10.1126/science.125799 Neuron 94:774–789
46. Liu TL, Upadhyayula S, Milkie DE, Singh V, 57. Papagiakoumou E, Ronzitti E, Chen IW,
Wang K, Swinburne IA, Mosaliganti KR, Col- Gajowa M, Picot A, Emiliani V (2018)
lins ZM, Hiscock TW, Shea J, Kohrman AQ, Two-photon optogenetics by computer-
Medwig TN, Dambournet D, Forster R, generated holography. In: Neuromethods, vol
Cunniff B, Ruan Y, Yashiro H, Scholpp S, 133, pp 175–197. Humana Press Inc.
Meyerowitz EM, Hockemeyer D, Drubin
DG, Martin BL, Matus DQ, Koyama M, 58. Martin Persson DE, Goksör M (2012) Reduc-
Megason SG, Kirchhausen T, Betzig E (2018) ing the effect of pixel crosstalk in phase only
Observing the cell in its native state: Imaging spatial light modulators. Optics Express
subcellular dynamics in multicellular organ- 20(20):22334–22343
isms. Science 360(6386). https://doi.org/10. 59. Kaufmann A, Mickoleit M, Weber M, Huisken
1126/science.aaq13 J (2012) Multilayer mounting enables long-
47. Baumgart E, Kubitscheck U (2012) Scanned term imaging of zebrafish development in a
light sheet microscopy with confocal slit detec- light sheet microscope. Development
tion. Optics Express 20(19):21805–21814 139(17):3242–3247
48. dal Maschio M, Donovan JC, Helmbrecht TO, 60. Giovannucci A, Friedrich J, Gunn P, Kalfon J,
Baier H (2017) Linking Neurons to Network Brown BL, Koay SA, Taxidis J, Najafi F, Gau-
Function and Behavior by Two-Photon Holo- thier JL, Zhou P, Khakh BS, Tank DW,
graphic Optogenetics and Volumetric Imaging. Chklovskii DB, Pnevmatikakis EA (2019) CaI-
Neuron 94:774–789 mAn an open source tool for scalable calcium
Optogenetics and SPIM 261
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International
License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing, adaptation, distribution
and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative Commons license,
unless indicated otherwise in a credit line to the material. If material is not included in the chapter’s Creative
Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use,
you will need to obtain permission directly from the copyright holder.
Chapter 9
Abstract
Optical imaging has the potential to reveal high-resolution information with minimal photodamage. The
recent renaissance of super-resolution, widefield, ultrafast, and computational imaging methods has broad-
ened its horizons even further. However, a remaining grand challenge is imaging at depth over a widefield
and with a high spatiotemporal resolution. This achievement would enable the observation of fast collective
biological processes, particularly those underpinning neuroscience and developmental biology. Multipho-
ton imaging at depth, combining temporal focusing and single-pixel detection, is an emerging avenue to
address this challenge. The novel physics and computational methods driving this approach offer great
potential for future advances. This chapter articulates the theories of temporal focusing and single-pixel
detection and details the specific approach of TempoRAl Focusing microscopy with single-pIXel detection
(TRAFIX), with a particular focus on its current practical implementation and future prospects.
Key words Imaging at depth, Temporal focusing, Widefield imaging, Multiphoton microscopy,
Single-pixel imaging, Compressive sensing
1 Introduction
Eirini Papagiakoumou (ed.), All-Optical Methods to Study Neuronal Function, Neuromethods, vol. 191,
https://doi.org/10.1007/978-1-0716-2764-8_9, © The Author(s) 2023
263
264 Philip Wijesinghe and Kishan Dholakia
the time domain rather than solely in the spatial domain [11, 12]
has revitalized the concept of widefield multiphoton imaging at
depth [13–15].
The premise that one can focus light in time rather than in
space has emerged rapidly over the past decades. Spatial focusing,
the concentration of the intensity of a light field in space, is ubiqui-
tous in virtually all-optical imaging systems. It is well-known that
spatial frequencies can be focused in space via the Fourier trans-
forming action of a lens [16]. Similarly, spectral frequencies can be
focused in time, with much of the same equivalence in their Fourier
transform properties, by introducing phase modulation and spatial
dispersion [17]. This has formed the foundation of pulse compres-
sion and has led to chirped pulse amplification [18] that won
Strickland and Mourou a share of the Nobel Prize in Physics in
2018. Defocusing or time stretching, on the other hand, has
enabled the recording of ultrashort phenomena and spectral con-
tent in the time domain [19].
More recently, the simultaneous use of spatial and temporal
focusing was presented in 2005 together by Oron et al. [11] and
Zhu et al. [12], demonstrating an interesting phenomenon wherein
a time-compressed pulse can be made to exist only at the focus of a
lens. Away from the focus, the pulse broadens in space and in time,
with a concomitant rapid reduction in its peak intensity. This
restriction of the pulse to the focal region enables axial confinement
of non-linear optical excitation in a scanless, widefield illumination
scheme. This development, termed temporal focusing (TF), has had
profound impact on multiphoton microscopy, where previously
axial confinement could only be achieved by point scanning a highly
focused beam across the image plane, which has severe limits on
temporal resolution. The initial work was followed by a flurry of
demonstrations of TF in widefield imaging [20–25], excitation
[26, 27], harmonic generation [28, 29], super-resolution [30],
micromachining [31–35], remote focusing [36, 37], tissue ablation
[38], and trapping [39], among others. Unsurprisingly, the capac-
ity for ultrafast widefield excitation has particularly flourished in
optogenetics and neuroimaging [40–49]. This is because the
absence of scanning has readily allowed for the simultaneous exci-
tation and measurement of neuronal firing events on the millisec-
ond timescale and over wide fields of view, previously unattainable
by point-scanning approaches.
While a direct mathematical correspondence can be made
between focusing in space and in time [19], it has become evident
that TF behaves differently to spatial focusing in the presence of
wavefront aberrations and through scattering media [50–55] due
to the added angular diversity of the illumination spectra at the
focus. For instance, the addition of TF has demonstrated a substan-
tial improvement in propagation through scattering media and a
reduction in speckle at the focus [51]. This discovery has enabled
Widefield Imaging with Temporal Focusing 265
2 Methods
Fig. 1 Illustrations of the pulse shape in space and time in (a) spatial focusing and (b) temporal focusing.
Temporal focusing realized using a (c) scattering plate, SP, and a (d) diffraction grating, DG. L: lens; Obj:
objective; FP: common Fourier plane; IP: image plane. Adapted from [11]
where
rffiffiffiffiffiffiffiffi
iπf
½1 þ i zzM - 2 ,
1
κ =Ω
zR
k0 α=f
γ = ,
1 þ iz=z M
iz=z B
χ = ,
1 þ iz=z M
4f 2 2z
s 22 = 2 2
þi ,
k0 s k0
1 2f 2 1 2f 2 1 2f 2
zM = , zR = , and z B = :
s 2 k0 s 2 þ α2 Ω2 k0 α2 Ω2 k0
Notably, the spatial profile at the focus is equivalent to that defined
by any one of the monochromatic beams, i.e., a Gaussian width of
2f/k0s. However, the modified Rayleigh range zR is dependent on
both s and αΩ. Typically, αΩ ≪ s for widefield illumination, there-
fore, zR is defined by the extent to which the spectral dispersion fills
the back aperture of the objective. Rayleigh-like coefficients zM and
zB, related to the spatial and temporal distributions, respectively,
additionally modify the phase evolution with time and away from
the focus. Here, z is defined as the distance away from the focus
(z = 0). The temporal evolution of intensity in the last exponent
defines the pulse shape, and the width can be given as [22]
pffiffiffiffiffiffiffiffiffiffiffi 12
2 2 ln 2 zM z2 ð3Þ
τðzÞ = 1þ :
Ω zB z2 þ zM zR
The pulse is shortest at the focus, reaching the minimum
transform-limited pulse width of 1/Ω (1/e2 radius). The important
conclusion of these equations is that, compared to a spatially
focused Gaussian beam, TF decouples axial sectioning (zR) from
the lateral beam shape (s2), which in turn allows precise control of
the multiphoton excitation profile in 3D.
An alternative and more intuitive view of TF was offered by
Durfee et al. [59] by examining the evolution of the phase delay
(chirp) of each frequency with respect to the focus and lateral
position. The formulation is detailed in Note 2. Figure 2a visualizes
the evolution of the phase front of three selected frequencies of the
Widefield Imaging with Temporal Focusing 269
Fig. 2 Profile of a temporally focused pulse in the (a) spectral and (b) time domains. Lines in (a) represent the
phase front delay of different spectral components. The profiles in (b) represent the pulse shapes and their
relative delay with respect to lateral position, x. PFT: pulse front tilt
Fig. 3 Point-spread function of (a,b) temporally focused and (c,d) spatially focused beams with (a,c) no
scattering and (b,d) though 900-μm of a scattering phantom. Scale bar is 20 μm. Reproduced from [13]
Fig. 4 Imaging process in scattering media. Imaging performed (a) in parallel with a camera; (b) sequentially
using point-scanning microscopy; and (c) by multiplexing with single-pixel detection
2.2.1 Compressive Compressive sensing (CS) describes the recovery of a signal from
Sensing far fewer measurements than required by the Nyquist sampling
criterion. The idea that one can accurately reconstruct
N independent values of a signal x from M ≫ N measurements is
counter-intuitive. However, let us consider a similar topic of image
compression, where a one megapixel raw image taking 4MB can
routinely be compressed into a 400-kB JPEG with imperceptible
losses in quality. This process relies on the idea of “sparsity.” Any
image x in the Cartesian coordinate space can be represented in a
different domain Ψ by coefficients s, such that x = Ψ s. Images are
considered sparse when there exists a domain where the coefficient
vector s possesses only a few non-zero values. The number of
non-zero values, K ≫ N, indicates that the image is “K-sparse,”
and it follows that the remainder N - K values are zero. In practice,
images are not perfectly sparse; however, s possesses a few large
coefficients with the rest being close to zero. For example, JPEG
compression selects Ψ to be the discrete cosine transform (DCT);
very briefly, a DCT is performed on x and the largest 10% of values
are stored.
CS, developed independently by Candes et al. [66] and
Donoho [67] in 2006, introduced compression directly to the
measurement process by leveraging the assumption that the signal
to be measured is sparse. This is in contrast to making a full
measurement and performing compression at a later time. If we
consider a case where the imaging process y = Φx is compressed by
performing 10% of the measurements needed to achieve Nyquist
sampling (M = N/10), such that Φ has N columns and M rows. It is
evident that this problem is underdetermined and the solution
space of all possible x that can generate y is infinite. For instance,
in point scanning, this would be equivalent to imaging the first 10%
of the field of view and leaving the other 90% up to interpretation.
CS approaches this problem by a careful design of the measurement
process. We realize that x can be decomposed into a different
domain with a sparse coefficient vector, modifying the imaging
process to y = Φx = ΦΨ s = Θs [68] (Fig. 5). A solution to the CS
problem lies in finding the most sparse s that satisfies y = Θs. The
major challenge in CS is twofold: first, is finding an efficient mini-
mization algorithm that can quantify the “sparsity” of s; and, sec-
ond, is the selection of an appropriate measurement matrix Φ. We
consider these aspects in turn.
In early CS work, a minimization algorithm using an l1-norm
proved to be effective in finding a sparse representation of s and
thus in recovering a compressed signal [66, 67]. Using this
approach, the estimated image x^ is recovered as
274 Philip Wijesinghe and Kishan Dholakia
1
https://statweb.stanford.edu/~candes/software/l1magic/.
Widefield Imaging with Temporal Focusing 275
Fig. 6 TRAFIX setup, illustrating the projection of a sequence of patterns (ϕi) and single-pixel detection of a
series of measurements (yi). BE: beam expander; DLS: dynamic light shaper; RL: relay lens; DG: diffraction
grating; L: lens; EP: entrance pupil; Obj: objective; S: sample; DM: dichroic mirror; SPD: single-pixel detector.
Numbers (1–3) correspond to the image plane on the DG, the common Fourier plane, and the sample image
plane, respectively
276 Philip Wijesinghe and Kishan Dholakia
2.3.2 Imaging Initial work employed the Hadamard matrix as the set of test
patterns (Φ) [13, 14, 75, 76]. The Hadamard matrix [65] (also
termed the Walsh–Hadamard matrix) is formed recursively. It is
orthonormal and symmetrical, with each column being orthogonal
to any other. It is routinely used for single-pixel imaging because it
is easy to generate and it is its own inverse. The first-order Hada-
mard matrix is unity; the second order is given as
" #
1 1
H2 = ; ð5Þ
1 -1
and the 2n order is formed from the n order as
278 Philip Wijesinghe and Kishan Dholakia
Fig. 7 Images recovered using TRAFIX. (a) Reference image of a fluorescent test sample with no scattering, (b)
imaged through a 400-μm brain slice using widefield detection, and (d) TRAFIX. (c) TRAFIX through a 200-μm
brain slice. (e,g) Reference images of mouse-derived astrocytes compared with (f,h) TRAFIX. Scale bar is
20 μm. (a–d) Adapted from [13]
" #
Hn Hn
H 2n = : ð6Þ
Hn -Hn
The sizes of Hadamard matrices are limited to powers of 2, which
set the allowed image sampling sizes. Each row of the Hadamard
matrix represents an individual test pattern.
Figure 7 shows examples of images generated with TRAFIX
using Hadamard test sequences [13]. Figure 7a shows a reference
image of a test target (x) fabricated from a 200-nm spun-coated
layer of super-yellow polymer. The target was imprinted by photo-
bleaching a negative pattern. When the test target is obscured by a
400-μm section of rat brain tissue (mean-free path length, ls = 55
μm), conventional widefield detection using a camera is impossible
due to scattering in the detection (Fig. 7b). Using TRAFIX, the
image may be retrieved when obscured by 200-μm (Fig. 7c) and
400-μm (Fig. 7d) section of rat brain tissue.
Figures 7e–h demonstrate the proof-of-principle of TRAFIX
on primary mouse astrocytes. No scattering was introduced; how-
ever, the results demonstrate the capacity to reconstruct images
from the faint signal in biological samples. Importantly, a long
recording time per test pattern on the order of 0.1–0.5 s was
required due to the low pulse energy density of the fast-repetition-
rate (80 MHz) laser used. However, even with a non-ideal laser,
three-photon signal recovery was demonstrated with TRAFIX
Widefield Imaging with Temporal Focusing 279
Fig. 8 Compressive sensing in microscopy. (a–c) The projected pattern (1), the spectrum in the Fourier plane
(2) clipped by the objective pupil (blue circle), and the resulting pattern in the sample plane (3), for the (a)
Hadamard, (b) random, and (c) Morlet patterns. (d) Performance of compression (M/N) using each pattern set,
compared to the reference image (e). Scale bar is 10 μm. Adapted from [58]
280 Philip Wijesinghe and Kishan Dholakia
Fig. 6. The blue circle indicates the entrance pupil size. The Hada-
mard pattern (Fig. 8a) demonstrates a large structured pattern in
the Fourier plane that exceeds the entrance pupil. The random
pattern (Fig. 8b) shows a broad specular bandwidth. In both
cases, the sampling pixel size of each pattern was set as the diffrac-
tion limit.
The limit in the spatial-frequency bandwidth leads to two
important issues in CS. First, the effective low-pass filtering leads
to a different pattern being projected onto the sample plane to the
one assumed in the CS optimization algorithm. This is clear by
comparing locations (1) and (3) in Fig. 8. In this scenario, the CS
problem becomes more poorly conditioned; the noise in the mea-
surement process may be attributed to larger non-existing frequen-
cies. Second, diffraction leads to a large portion of the pulse energy
to be blocked by the entrance pupil. In the widefield regime, pulse
energy is an important parameter to maximize, directly impacting
the signal-to-noise ratio.
An alternative pattern set was proposed to control and selec-
tively probe the spatial-frequency space, generated from Morlet
wavelets mathematically convolved with randomly generated matri-
ces [58, 77]. These Morlet patterns feature two important proper-
ties. First, the Morlet wavelet, based on real-valued Gabor filters,
reaches the limit set by the uncertainty principle, i.e., an optimal
trade-off between spatial and spatial-frequency localization. Sec-
ond, the convolution with a randomly generated, Gaussian-
distributed matrix satisfies the mutual incoherence property
required by CS.
Figure 8c shows the propagation of a Morlet pattern [58]. It is
evident that the Morlet patterns can be designed to fit the entrance
pupil, and the pattern at the sample plane resembles the pattern
projected by the DLS. In this demonstration, the sampling pixel
size N matches the theoretical diffraction limit. Interestingly, the
CS sampling pixels size sets the image resolution, yet, due to the
Nyquist sampling criterion, a bandwidth of twice the pattern fre-
quency is required to propagate the pattern through a microscopy
system. This corresponds to the observations in Fig. 8a–c (2) and
suggests that an image resolution below twice the diffraction limit
is not achievable with Hadamard or Random measurement matri-
ces. Optionally, this can be overcome with digital microscanning
techniques [78], by taking multiple CS measurements with a series
of patterns, spatially shifted by half the sampling size. Morlet pat-
terns, due to the independent selection of bandwidth and sampling
size, are able to reach the diffraction limit.
CS recovery from microscopy data requires additional consid-
eration. The conventional method of CS recovery using l1-norm
minimization using the basis pursuit algorithm is inefficient for
imaging applications for several reasons. Importantly, the CS
Widefield Imaging with Temporal Focusing 281
2
https://statweb.stanford.edu/~candes/software/nesta/.
282 Philip Wijesinghe and Kishan Dholakia
3 Future Prospects
4 Summary
5 Notes
x τ0 β z τ20 β2 1
ϕ2 ðx,zÞ = ð - Þð :Þ ð10Þ
s 2 ω0 z R 4 1 þ z 2 =z 2R
Notably, ϕ2 is dominated by the quadratic term, which is
negligible at the focus and increases with z, with a linear rela-
tionship when z is well within the Rayleigh range z R = k0 s 22 =2.
This second-order chirp leads to pulse broadening from tem-
poral dispersion away from the focus (illustrated in Fig. 2).
3. Stability of Compressive Sensing
Like many other inversion methods, the efficacy of CS
relies on several closely linked parameters: the mutual incoher-
ence and the condition number of the measurement matrix,
and the noise in the measurement. The interplay between these
parameters dictates the qualities seen in the recovered images.
For instance, a compressed Hadamard matrix is not mutually
incoherent (in fact, it is perfectly coherent, i.e., there will be at
least two columns that are identical); however, it is well-
Widefield Imaging with Temporal Focusing 287
References
1. Escobet-Montalbán A, et al. (Apr 14, 2019) 15. Sun B, et al. (2018) Four-dimensional light
TRAFIX: Imaging at depth with temporal shaping: manipulating ultrafast spatiotemporal
focusing and single-pixel detection. In: Bio- foci in space and time. Light Sci Appl 7:17117–
photonics congress: Optics in the life sciences 17117
congress 2019 (BODA, BRAIN, NTM, OMA, 16. Kammel R, et al. (2014) Enhancing precision
OMP) (2019), paper NT1C.4, NT1C.4 in fs-laser material processing by simultaneous
2. Adam Y, et al. (2019) Voltage imaging and spatial and temporal focusing. Light Sci Appl 3:
optogenetics reveal behaviour dependent e169–e169
changes in hippocampal dynamics. Nature 17. Escobet-Montalbán A, et al. (Feb 22, 2019)
569:413–417 Wide-field multiphoton imaging with
3. Block E, et al. (2013) Simultaneous spatial and TRAFIX. In: Multiphoton microscopy in the
temporal focusing for tissue ablation. Biomed biomedical sciences XIX, vol 10882, 108821G
Opt Exp 4:831–841 18. Dai W, Milenkovic O (2009) Subspace Pursuit
4. Turcotte R, et al. (2019) Dynamic superreso- for Compressive Sensing Signal Reconstruc-
lution structured illumination imaging in the tion. IEEE Trans Inf Theory 55:2230–2249
living brain. Proc Natl Acad Sci 116:9586– 19. Schrödel T, et al. (2013) Brain-wide 3d imag-
9591 ing of neuronal activity in Caenorhabditis ele-
5. Dana H, et al. (2014) Hybrid multiphoton gans with sculpted light. Nature Methods 10:
volumetric functional imaging of large-scale 1013–1020
bioengineered neuronal networks. Nature 20. Becker S, Bobin J, Candès E (2011) NESTA: A
Communications 5:1–7 Fast and Accurate First-Order Method for
6. Papagiakoumou E, et al. (2010) Scanless Sparse Recovery. SIAM J Imag Sci 4:1–39
two-photon excitation of 21. Oron D, Silberberg Y (2005) Spatiotemporal
channelrhodopsin-2. Nature Methods 7:848– coherent control using shaped, temporally
854 focused pulses. Optics Express 13:9903–9908
7. Pégard NC, et al. (2016) Compressive light- 22. Oron D, Silberberg Y (2005) Harmonic gen-
field microscopy for 3D neural activity record- eration with temporally focused ultrashort
ing. Optica 3:517–524 pulses. JOSA B 22:2660–2663
8. Rowlands CJ, et al. (2019) Increasing the pen- 23. Papagiakoumou E, et al. (2008) Patterned
etration depth of temporal focusing multipho- two-photon illumination by spatiotemporal
ton microscopy for neurobiological shaping of ultrashort pulses. Optics Express
applications. J Phys D Appl Phys 52:264001 16:22039–22047
9. Andrasfalvy BK, et al. (2010) Two-photon sin- 24. Donoho DL (2006) Compressed Sensing.
gle-cell optogenetic control of neuronal activ- IEEE Trans Inf Theory 52:1289–1306
ity by sculpted light. Proc Natl Acad Sci 107: 25. Czajkowski KM, Pastuszczak A, Kotyn´ski R
11981–11986 (2018) Single-pixel imaging with Morlet wave-
10. Czajkowski KM, Pastuszczak A, Kotyński R let correlated random patterns. Scientific
(2018) Real-time single-pixel video imaging Reports 8:466
with Fourier domain regularization. Optics 26. Zhang Y, et al. (2019) Overcoming tissue scat-
Express 26:20009–20022 tering in wide-field deep imaging by extended
11. Bègue A, et al. (2013) Two-photon excitation detection and computational reconstruction.
in scattering media by spatiotemporally shaped bioRxiv, 611038
beams and their application in optogenetic 27. Papagiakoumou E, Ronzitti E, Emiliani V
stimulation. Biomed Opt Exp 4:2869–2879 (2020) Scanless two-photon excitation with
12. Zhang S, et al. (2014) Analysis of pulse front temporal focusing. Nature Methods, 1–11
tilt in simultaneous spatial and temporal focus- 28. Dana H, Shoham S (2011) Numerical evalua-
ing. JOSA A 31:2437–2446 tion of temporal focusing characteristics in
13. Vitek DN, et al. (2010) Temporally focused transparent and scattering media. Optics
femtosecond laser pulses for low numerical Express 19:4937–4948
aperture micromachining through optically 29. Papagiakoumou E, et al. (2013) Functional
transparent materials. Optics Express 18: patterned multiphoton excitation deep inside
18086–18094 scattering tissue. Nature Photonics 7:274–278
14. Lu R, et al. (2017) Video-rate volumetric func- 30. Sun M-J, et al. (2016) Improving the signal-to-
tional imaging of the brain at synaptic resolu- noise ratio of single-pixel imaging using digital
tion. Nature Neuroscience 20:620–628
Widefield Imaging with Temporal Focusing 289
microscanning. Optics Express 24:10476– microscopy. Opt Commun Opt Life Sci 281:
10485 1796–1805
31. Rodrı́guez C, et al. (2018) Three-photon fluo- 46. Leshem B, et al. (2014) When can temporally
rescence microscopy with an axially elongated focused excitation be axially shifted by disper-
Bessel focus. Optics Letters 43:1914–1917 sion?. Optics Express 22:7087–7098
32. Vitek DN, et al. (2010) Spatio-temporally 47. Mardinly AR, et al. (2018) Precise multimodal
focused femtosecond laser pulses for nonrecip- optical control of neural ensemble activity.
rocal writing in optically transparent materials. Nature Neuroscience 21:881–893
Optics Express 18:24673–24678 48. Goodman JW (2005) Introduction to Fourier
33. Therrien OD, et al. (2011) Wide-field multi- optics, 520 pp. Roberts and Company
photon imaging of cellular dynamics in thick Publishers
tissue by temporal focusing and patterned illu- 49. Baumert LD, Hall M (1965) Hadamard matri-
mination. Biomed Opt Exp 2:696–704 ces of the Williamson type. Math Comput 19:
34. Spesyvtsev R, Rendall HA, Dholakia K (2015) 442–447
Wide-field three-dimensional optical imaging 50. Du R, et al. (2009) Analysis of fast axial scan-
using temporal focusing for holographically ning scheme using temporal focusing with
trapped microparticles. Optics Letters 40: acousto-optic deflectors. J Modern Opt 56:
4847–4850 81–84
35. Horton NG, et al. (2013) In vivo three-photon 51. Aulbach J, et al. (2011) Control of Light
microscopy of subcortical structures within an Transmission through Opaque Scattering
intact mouse brain. Nature Photonics 7:205– Media in Space and Time. Phys Rev Lett 106:
209 103901
36. Candes EJ, Romberg J, Tao T (2006) Stable 52. Escobet-Montalbán A, et al. (2018) Wide-field
signal recovery from incomplete and inaccurate multiphoton imaging through scattering
measurements. Commun Pure Appl Math 59: media without correction. Science Advances
1207–1223 4:eaau1338
37. Sie YD et al. (2018) Fast and improved bioima- 53. Durst ME, Zhu G, Xu C (2006) Simultaneous
ging via temporal focusing multiphoton excita- spatial and temporal focusing for axial scan-
tion microscopy with binary digital- ning. Optics Express 14 12243–12254
micromirror device holography. J Biomed Opt 54. Wang T, et al. (2018) Three-photon imaging
23:116502 of mouse brain structure and function through
38. Zhu G, et al. (2005) Simultaneous spatial and the intact skull. Nature Methods 15:789–792
temporal focusing of femtosecond pulses. 55. Chen I-W, et al. (2019) In vivo submillisecond
Optics Express 13:2153–2159 two-photon optogenetics with temporally
39. Baraniuk RG (2007) Compressive sensing focused patterned light. J Neurosci 39:3484–
[Lecture Notes]. IEEE Sig Process Mag 24: 3497
118–121 56. McCabe DJ, et al. (2011) Spatio-temporal
40. Kazemipour A, et al. (2019) Kilohertz frame- focusing of an ultrafast pulse through a multi-
rate two-photon tomography. Nature Methods ply scattering medium. Nature Communica-
16:778–786 tions 2:1–5
41. Dana H, et al. (2013) Line temporal focusing 57. Lu R, et al. (2018) 50 Hz volumetric func-
characteristics in transparent and scattering tional imaging with continuously adjustable
media. Optics Express 21:5677–5687 depth of focus. Biomed Opt Exp 9:1964–1976
42. Baraniuk R, et al. (2008) A simple proof of the 58. Abdelfattah AS, et al. (2019) Bright and
restricted isometry property for random matri- photostable chemigenetic indicators for
ces. Constructive Approximation 28:253–263 extended in vivo voltage imaging. Science
43. Candes EJ, Romberg J, Tao T (2006) Robust 365:699–704
uncertainty principles: exact signal reconstruc- 59. Papagiakoumou E, et al. (2009) Temporal
tion from highly incomplete frequency infor- focusing with spatially modulated excitation.
mation. IEEE Trans Inf Theory 52:489–509 Optics Express 17:5391–5401
44. Treacy E (1969) Optical pulse compression 60. Choi H, et al. (2013) Improvement of axial
with diffraction gratings. IEEE J Quantum resolution and contrast in temporally focused
Electron 5:454–458 widefield two-photon microscopy with
45. Durst ME, Zhu G, Xu C (2008) Simultaneous structured light illumination. Biomed Opt
spatial and temporal focusing in nonlinear Exp 4:995–1005
290 Philip Wijesinghe and Kishan Dholakia
61. Straub A, Durst ME, Xu C (2011) High speed 76. Vaziri A, et al. (2008) Multilayer three-
multiphoton axial scanning through an optical dimensional super resolution imaging of thick
fiber in a remotely scanned temporal focusing biological samples. Proc Natl Acad Sci 105:
setup. Biomed Opt Exp 2:80–88 20221–20226
62. Pégard NC, et al. (2017) Three-dimensional 77. Kolner BH, Nazarathy M (1989) Temporal
scanless holographic optogenetics with tempo- imaging with a time lens. Optics Letters 14:
ral focusing (3d-SHOT). Nature Communica- 630–632
tions 8:1–14 78. Figueiredo MAT, Nowak RD, Wright SJ
63. Ji N, Freeman J, Smith SL (2016) Technolo- (2007) Gradient projection for sparse recon-
gies for imaging neural activity in large struction: application to compressed sensing
volumes. Nature Neuroscience 19:1154–1164 and other inverse problems. IEEE J Sel Top
64. Ronzitti E, Emiliani V, Papagiakoumou E Sig Process 1:586–597
(2018) Methods for three-dimensional all-op- 79. Wijesinghe P, et al. (2019) Optimal compres-
tical manipulation of neural circuits. Front Cell sive multiphoton imaging at depth using
Neurosci 12:469 single-pixel detection. Optics Letters 44:4981
65. Wei X, et al. (2020) Real-time frequency- 80. Accanto N, et al. (2018) Multiplexed tempo-
encoded spatiotemporal focusing through scat- rally focused light shaping for high- resolution
tering media using a programmable 2d ultra- multi-cell targeting. Optica 5:1478–1491
fine optical frequency comb. Science Advances 81. Dana H, Shoham S (2012) Remotely scanned
6:eaay1192 multiphoton temporal focusing by axial grism
66. Alemohammad M, et al. (2018) Widefield scanning. Optics Letters 37: 2913–2915
compressive multiphoton microscopy. Optics 82. Yang J, Zhang Y, Yin W (2010) A fast alternat-
Letters 43:2989–2992 ing direction method for TVL1-L2 signal
67. Vaziri A, Shank CV (2010) Ultrafast widefield reconstruction from partial fourier data. IEEE
optical sectioning microscopy by multifocal J Sel Top Sig Process 4:288–297
temporal focusing. Optics Express 18:19645– 83. Hernandez O, et al. (2016) Three-dimensional
19655 spatiotemporal focusing of holographic pat-
68. Mosk AP, et al. (2012) Controlling waves in terns. Nature Communications 7:1–11
space and time for imaging and focusing in 84. Strickland D, Mourou G (1985) Compression
complex media. Nature Photonics 6:283–292 of amplified chirped optical pulses. Optics
69. Wadduwage DN et al. (2019) De-scattering Communications 56:219–221
with excitation patterning (DEEP) enables 85. Wu Y, Shroff H (2018) Faster sharper and
rapid wide-field imaging through scattering deeper: structured illumination microscopy
media. arXiv:1902.10737 [physics] for biological imaging. Nature Methods 15:
70. Tal E, Oron D, Silberberg Y (2005) Improved 1011
depth resolution in video-rate line-scanning 86. Abdeladim L, et al. (2019) Multicolor multi-
multiphoton microscopy using temporal focus- scale brain imaging with chromatic multipho-
ing. Optics Letters 30:1686–1688 ton serial microscopy. Nature Communications
71. Piatkevich KD, et al. (2019) Population imag- 10:1662
ing of neural activity in awake behaving mice. 87. Durfee CG, et al. (2012) Intuitive analysis of
Nature 574:413–417 space-time focusing with double-ABCD calcu-
72. Kim D, So PTC (2010) High-throughput lation. Optics Express 20:14244–14259
three-dimensional lithographic microfabrica- 88. Parot VJ, et al. (2019) Compressed Hadamard
tion. Optics Letters 35:1602–1604 microscopy for high-speed optically sectioned
73. Oron D, Tal E, Silberberg Y (2005) Scanning- neuronal activity recordings. J Phys D Appl
less depth-resolved microscopy. Optics Express Phys 52:144001
13:1468–1476 89. Prevedel R, et al. (2016) Fast volumetric cal-
74. Katz O, et al. (2011) Focusing and compres- cium imaging across multiple cortical layers
sion of ultrashort pulses through scattering using sculpted light. Nature Methods 13:
media. Nature Photonics 5:372–377 1021–1028
75. He F, et al. (2011) Independent control of 90. Villette V, et al. (2019) Ultrafast two-photon
aspect ratios in the axial and lateral cross sec- imaging of a high-gain voltage indicator in
tions of a focal spot for three-dimensional fem- awake behaving mice. Cell 179:1590–
tosecond laser micromachining. New J Phys 1608.e23
13:083014
Widefield Imaging with Temporal Focusing 291
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International
License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing, adaptation, distribution
and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative Commons license,
unless indicated otherwise in a credit line to the material. If material is not included in the chapter’s Creative
Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use,
you will need to obtain permission directly from the copyright holder.
Chapter 10
Abstract
Brain is composed of complex networks of neurons that work in concert to underlie the animal’s cognition
and behavior. Neurons communicate via structures called synapses, which typically require submicron
spatial resolution to visualize. To understand the computation of individual neurons as well as neural
networks, methods that can monitor neuronal morphology and function in vivo at synaptic spatial resolu-
tion and sub-second temporal resolution are required. In this chapter, we discuss the principles and
applications of two enabling optical microscopy methods: two-photon fluorescence microscopy equipped
with Bessel focus scanning technology and widefield fluorescence microscopy with optical sectioning ability,
both of which could be combined with optogenetic stimulation for all optical interrogation of neural
circuits. Details on their design and implementation, as well as example applications, are presented.
Key words Optical microscopy, Neural imaging, Synapse, Bessel beam, Two-photon fluorescence
microscopy, Widefield microscopy, Structured illumination microscopy
1 Introduction
Authors Guanghan Meng and Qinrong Zhang have equally contributed to this chapter.
Eirini Papagiakoumou (ed.), All-Optical Methods to Study Neuronal Function, Neuromethods, vol. 191,
https://doi.org/10.1007/978-1-0716-2764-8_10, © The Author(s) 2023
293
294 Guanghan Meng et al.
Fig. 1 The concept of Bessel focus scanning multiphoton microscopy. (a) 2D scanning of a conventional
Gaussian focus obtains information from a single plane (the plane in yellow). (b) 2D scanning of a Bessel focus
covers a volume (the volume in orange)
2.1.3 Bessel Focus During most brain imaging experiments in vivo, neurons exhibit
Scanning Technology variations in their fluorescence signal associated with their func-
tional activity, but their positions remain unchanged. Therefore,
when the temporal dynamics of neurons is the subject of interest,
one does not need to constantly track their 3D locations, but can
instead obtain axially projected views of 3D volumes via extended
depth-of-field imaging by scanning an axially elongated focus, e.g.,
a Bessel-like beam [14–16, 39–41] (Fig. 1). Since all structures
along the elongated focus are probed simultaneously within a single
2D scan, the 2D frame rate becomes a projected 3D volume rate.
Although axial resolution is reduced substantially during Bessel
focus scanning, the 3D positions of neurons and neuronal struc-
tures can be obtained from a conventional 3D stack by scanning a
Gaussian focus. Because at the same numerical aperture (NA), a
Bessel focus has a higher lateral resolution than a Gaussian focus
[41], synapse-resolving lateral resolution is maintained even for a
0.3-NA Bessel focus [42]. With Bessel focus scanning, imaging
throughput can improve by tens to a hundred times, with the
image data size reduced by the same amount [15]. Bessel focus
scanning technology is compatible with other fast scanning meth-
ods mentioned in Subheading 2.1.2 and, when combined together,
can further boost the volumetric imaging speed of a multiphoton
microscope [15, 38]. The extended depth of field of a Bessel focus
also makes the imaging process insensitive to axial motion, which
eliminates axial motion artifacts. Together with the much reduced
data size, it substantially simplifies image processing
[15, 42]. Although Bessel focus scanning technology can be
incorporated into both two-photon and three-photon fluorescence
High-Speed Neural Imaging with Synaptic Resolution 297
Fig. 2 The Diagram of two Bessel focus modules. (a) A two-photon microscope with an SLM-based Bessel
module (gray box). (b) An axicon-based Bessel module. L2 can be translated along optical axis to change the
axial length and numerical aperture of Bessel beam. D is defined as 0 mm when the mask is at the front focal
plane of L2 and positive when L2 moves away from the mask. Ti:Sa Ti:Sapphire laser, EOM electro-optical
modulator, BE beam expander, M mirror, L lens, Obj objective
2.2 Materials and A typical two-photon fluorescence microscope includes the follow-
Equipment ing optical components (Fig. 2): a femtosecond laser, an excitation
power control unit such as an electro-optical modulator (EOM)
(i.e., a Pockels cell), a pair of scanners/galvos, a scan lens and a tube
lens, a dichroic mirror, an objective (often mounted on a piezoelec-
tric stage to perform axial scanning), detection filters, one or two
photomultipliers (PMTs).
A Bessel focus module includes the following optical compo-
nents (Fig. 2): two mirrors to switch the light path between Gauss-
ian and Bessel modes, an SLM or an axicon, a lens, an annular mask,
and a pair of conjugation lenses.
on the second galvo as well as the objective back focal plane during
scanning, but leads to more power loss due to adding additional
optical elements. In some systems, a resonant scanner is
incorporated in addition to the two raster-scanning galvos, with
the three scanners together enabling high-speed imaging in a small
subfield positioned anywhere inside a large field of view
[36, 38]. After the galvos, a scan lens (L4) and a tube lens
(L5) conjugate the scanners to the back focal plane of the objective.
The emission light path (e.g., the dichroic mirror and the detec-
tors) is not shown in the diagram.
2.3.2 Design and Setup An annular illumination pattern at the objective back focal plane
of a Bessel Focus Module generates a Bessel-like focus. This annular illumination can be
created with a phase mask [14], an SLM [15, 42], or an axicon
[16, 39, 40].
In an SLM-based Bessel focus module (Fig. 2a, rectangle box),
a reflective phase-only SLM is placed at the front focal plane of a
lens (L1). A circular binary phase pattern (alternating 0 and π) on
the SLM diffracts the incident Gaussian beam preferentially into the
1 diffraction orders, which form an annular ring at the back focal
plane of L1. An annular aperture mask is placed at the back focal
plane of L1 to selectively transmit the desired annular electric field,
which is conjugated to the galvos by a pair of lenses L2 and L3, and
then to the objective back focal plane.
An axicon-based Bessel focus module (Fig. 2b) has a similar
configuration to an SLM-based module, except that an axicon is
placed at the front focal plane of L1. The conical surface of the
axicon refracts the light according to Snell’s law, which forms a ring
at the back focal plane of L1. The annular aperture mask at the back
focal plane of L1 is not necessary for an ideal axicon (with infinitely
small conical tip), but is necessary in practice to block the unwanted
light refracted through the tip of an imperfect axicon.
When it comes to choosing between an SLM and an axicon-
based Bessel module, several factors need to be considered. An
SLM-based module offers more flexibility in terms of point spread
function (PSF) engineering [15] and allows the NA and axial length
of the Bessel focus to be adjusted independently [15]. Details on
designing Bessel foci with different PSF profiles are discussed in
Subheading 2.3.3, but a brief overview is presented here. In both
methods, the axial length of the Bessel focus can be altered by
varying the size of the Gaussian beam impinging on the SLM or
the axicon, for example, by adding a beam expander or reducer at
the entrance of the Bessel module. With the beam size fixed, a user
can adjust the NA and axial length of the Bessel focus indepen-
dently by changing the phase pattern on the SLM and the dimen-
sions of the annular mask. Therefore, an SLM-based module is ideal
for systems utilizing multiple objectives with different NAs (e.g., a
1.05-NA objective for neocortical imaging or a 0.5-NA
High-Speed Neural Imaging with Synaptic Resolution 299
2.3.3 Design of a Bessel We use MATLAB® to calculate the PSF profiles of Bessel foci and to
Module guide the design of Bessel module. The underlying physics is
described below, and the MATLAB codes can be found in Refs.
[15, 16].
Generation of Annular Concentric binary grating patterns with phase values alternating
Illumination with an SLM between 0 and π are applied to a phase-only SLM to diffract most of
(Adapted from Ref. [15]) the incident electromagnetic field into the 1 orders (see note 2 in
Subheading 2.4), which after lens L1 forms a ring at the mask
plane. The radius of the ring (ρ), determined by the period of the
grating (d), the focal length of L1 ( f1), and the wavelength of the
light (λ), is calculated from the grating equation as:
f 1λ
ρ¼ :
d
For an annular mask to transmit the ring and block the other
diffraction orders, its inner and outer diameters Di and Do should
satisfy the relation:
D o þ D i ¼ 4ρ:
Combining the two equations above, to generate an annular
illumination pattern that centers on an annular mask with inner and
outer diameters Do and Di, the period of the circular binary grating
on the SLM is:
4f 1 λ
d¼ :
Do þ Di
With the size of the SLM pixel defined as p, the period of the
circular binary grating in units of pixels S is:
d 4f 1 λ
S¼ ¼ :
p pðD o þ D i Þ
300 Guanghan Meng et al.
Generation of Annular For an axicon with an apex angle A, the angle of incidence α on the
Illumination with an Axicon conical surface is: α ¼ πA
2 . The refraction angle is then derived
using Snell’s law: αr ¼ nα, given that α is small and n is the refractive
index of the axicon. Therefore, the angle between the refracted
light and the optical axis is:
θ ¼ αr α ¼ ðn 1Þα:
The radius of the ring at mask plane is:
ρ ¼ θf 1 ¼ ðn 1Þαf 1 :
The annular mask to transmit the ring should again satisfy:
D o þ D i ¼ 4ρ:
Therefore, the apex angle and refractive index of the axicon
should meet:
Do þ Di
ðn 1Þα ¼ :
4f 1
Calculation of Two-Photon Two-photon excitation PSF can be calculated using Richards and
Excitation PSF Wolf integrals [44, 45], from which both the lateral and axial full
width at half maximum (FWHM) of the PSF can be determined.
The information required for PSF calculation includes: the wave-
length of excitation light, the objective information (i.e., NA,
magnification, immersion media), and the electric field distribution
at the objective back focal plane.
The following equations hold for all infinity-corrected
objectives:
FL tube
F L obj ¼ ;
M obj
BPD obj ¼ 2N A obj F L obj
where the FLobj is the focal length of the objective, FLtube is the
focal length of the tube lens, Mobj is the magnification of the
objective, and BPDobj is the back pupil diameter of the objective.
The Bessel annulus at the back focal plane with an outer diameter
Do and inner diameter Di gives rise to an excitation NA as:
Do
N ABessel ¼ N A obj :
BPD obj
Or:
Do
N ABessel ¼ :
2F L obj
High-Speed Neural Imaging with Synaptic Resolution 301
The PSF profile is determined by Do, Di, and the electrical field
distribution within the annulus, which is discussed below.
Design of the Annular The annular mask is designed to be conjugated to the objective
Aperture Mask in an SLM- back focal plane, with a magnification factor M, which is deter-
Based Bessel Module mined jointly by focal lengths of lenses L2–L5. In most cases, the
(Adapted from Ref. [15]) objective, L4, and L5 (scan lens and tube lens) are already selected
and built prior to the design of the Bessel module as an add-on.
Therefore, one only needs to choose L2 and L3 to fit into the
available space and conjugate the mask plane to the galvos. It has
been demonstrated previously that a 0.4-NA Bessel focus works
well for in vivo two-photon fluorescence imaging in the brain [15],
a 0.3-NA Bessel focus has better performance than 0.4-NA one
when combined with two-photon fluorescence microendoscopy
[42] (due to the substantial off-axis aberrations of gradient refrac-
tive index lenses), and a larger NA Bessel beam is more suitable for
three-photon microscopy [37]. With the objective, L2–L5, and the
desired NA selected, the outer diameter of the annular mask is
determined as:
2N A Bessel F L obj
Do ¼ :
M
With NA (i.e. Do) fixed, the axial length of the Bessel beam is
dictated by the thickness of the ring, which is determined by the
focal length of L1 and the beam diameter on SLM, with smaller f1
and larger beamD generating thinner rings and longer Bessel foci.
The axial length of the Bessel focus can be further adjusted by fine-
tuning Di and S.
The electrical field distribution at the mask, including its ampli-
tude and phase, can be visualized via simulation (Fig. 3). The
annular mask selectively passes the electric field within Di < r < Do,
where r is the radial coordinate at the mask plane, and is usually
centered around the largest electric field amplitude (rmax, blue line,
Fig. 3). Two cases are of particular interest. In Case I, Di and Do are
selected to have the annulus fall on the two amplitude peaks closest
to and on either side of rmax (Fig. 4b, red lines in Fig. 3a, c). In Case
II, the edges of the annulus are located at the two zero-amplitude
crossings that are closest to and on either side of rmax (Fig. 4e, green
lines in Fig. 3a, c). Even though Case I has a thicker annulus than
Case II, the resulting Bessel focus has a longer axial FWHM than
that of Case II (Fig. 4c, f). This is because the negative parts of the
electric field (between the green and red lines, Fig. 3a, c) destruc-
tively interfere with the positive parts (between the green lines) and
broaden the axial FWHM. Reducing the thickness further from
Case II (e.g., between the purple lines) increases the axial FWHM
again. As shown in Fig. 4, to ensure that rmax is located at the center
of the annulus, without changing Do (i.e., the NA of the Bessel
beam), the period of the SLM pattern, S, can be concurrently
altered with Di.
302 Guanghan Meng et al.
Fig. 3 Electrical field distribution at the mask plane in an SLM-based Bessel module. (a) Amplitude and (b)
phase of the (c) electric field at different radial positions. 0 is the location of the optical axis. The red and green
dashed lines in (a) and (c) represent the inner and outer diameters of the annular masks for Case I and Case II
discussed in Subheading 2.3.3.4. (This figure is adapted from Ref. [15])
Design of an Axicon-Based When using an axicon for Bessel beam generation, one should start
Bessel Module with selecting a high-quality axicon (e.g., 1-APX-2-H254-P,
ALTECHNA Inc.; XFL25–010-U-B, ASPHERICON). Once the
axicon is selected, the apex angle α is determined (see Subheading
2.3.3.2); thus, the radius of the ring at the mask plane is only
determined by the focal length of L1 (Fig. 2b). From here, one
High-Speed Neural Imaging with Synaptic Resolution 303
Fig. 4 Profiles of Bessel foci using masks with different ring thickness but the same outer diameter. SLM
phase pattern, the dimension of an electric field at annular mask, and a measured profile of a 2 μm diameter
fluorescent bead, corresponding to (a–c) Case I and (d–f) Case II in Fig. 3. (g, h) Simulated axial profiles of
Bessel foci generated by annular masks of different widths. (This figure is adapted from Ref. [15])
304 Guanghan Meng et al.
can simulate the electric field at the mask plane and the 3D PSF
(code can be found in Ref. [16]) to guide the selection of L1, L2,
and L3. Example simulation results are presented in Fig. 5. In this
simulation, lens L2 is translated along the optical axis. The location
where the front focal plane of L2 coincides with the back focal plane
of L1 (i.e., the mask plane) is indicated as D ¼ 0 mm (Fig. 5, second
column). When L2 is moving toward the mask plane (negative D),
more power is allocated to the central region of the objective back
focal plane (Fig. 5, first column), which yields a smaller effective NA
and, together with the phase distribution of the pupil function, a
longer Bessel focus. In contrast, moving the lens away from the
mask plane distributes more power to the edge of the objective
pupil function, leading to a bigger effective NA and a shorter focus
(Fig. 5, third column).
The mean radius of the annular mask becomes well defined
once L1 and axicon are both selected (see Subheading 2.3.3.2),
although one can jointly vary the outer and inner radii of the ring
and obtain different ratios of transmittance and axial profiles. The
mask is intended to eliminate the unwanted light passing through
the imperfect, typically round, tip of the axicon, which if not
blocked can interfere with the rest of the refracted electric field
and cause the measured PSF to deviate from simulation. The thin-
ner the annular mask is, the more effectively it can block the
unwanted light, but more power loss will be introduced to the
system. For the case when limited excitation power is available,
one can use a thicker mask, or even no mask, and obtain
non-theoretical yet still usable PSF profiles [16].
2.3.4 Alignment of The Bessel focus module is located between the laser and the
Bessel Module (Adapted two-photon fluorescence microscope. First, obtain locations of
from Ref. [15]) the front and back focal planes of lenses L1–L3 (e.g., from their
Zemax models or vendor specifications). Second, place lenses so
Installation of the Optical that the back focal plane of L3 is at the first scanner (or between the
Components two scanners, if they are not conjugated), the front focal plane of L3
is superimposed with the back focal plane of L2, and the front focal
plane of L2 is superimposed with the back focal plane of L1 (i.e.,
the mask). At last, place the SLM or axicon at the front focal plane
of L1 (see note 4 in Subheading 2.4). It is helpful to have two
mirrors with tip/tilt mounts between L3 and the first scanner to
co-align Gaussian and Bessel paths. Alternatively, if there is not
enough space, set up one mirror with both tip/tilt and translational
controls.
Gross Alignment of the Set up two irises in the optical path shared by the Gaussian and
Bessel and Gaussian Beam Bessel focus modalities between the alignment mirror
Paths (s) mentioned in Subheading 2.3.4.1 and the first scanner. They
should be sufficiently separated and not conjugated with each
other. Ideally, the first alignment iris in the Bessel module should
High-Speed Neural Imaging with Synaptic Resolution 305
Fig. 5 Engineering Bessel foci by displacing Lens L2 in an axicon-based Bessel module. (a, e, i) 2D and (b, f, j)
1D (along x direction) representations of the amplitude and phase of the electric fields at the objective back
focal plane, (c, g, k) axial, and (d, h, l) lateral two-photon excitation point spread functions for L2
displacements of D ¼ 20 mm, 0 mm, and 20 mm, respectively. (This figure is adapted from Ref. [16])
306 Guanghan Meng et al.
be placed right after the alignment mirror(s) and the second iris as
close to the first scanner as possible to increase the accuracy of the
alignment. Adjust the positions of the irises to center on the (well
aligned) Gaussian beam. For SLM-based Bessel module, apply a
concentric binary grating on the SLM, and the reflected beam
appears as concentric rings. Then adjust the mirror(s) between L3
and the first scanner iteratively so that the Bessel concentric rings
pass through and center on the two irises (see note 5 in
Subheading 2.4).
Fine Alignment For fine alignment, dismount the objective and place a camera (e.g.,
DCC1545M, Thorlabs) at the back focal plane of the objective.
Under the Bessel mode, one should see a ring on the camera. (For a
well conjugated system, where the scanners are all conjugated to
each other, sweeping one or more scanners should not move the
position of the ring on the camera. But if the microscope does not
have its scanners conjugated, place the camera at the plane with
minimal movement.) Adjust the axial positions of L2 and L3 so that
the ring is sharpest on the camera, which indicates that the back
focal plane of the objective is conjugated to the annular mask (see
notes 6–8 in Subheading 2.4).
Placement of annular filter mask: Place the annular filter mask
at the back focal plane of L1. Apply the corresponding concentric
binary grating as calculated in Subheading 2.3.3 and adjust first the
lateral position of the mask until the post-mask ring is symmetric
and then the axial position so that the power passing through the
annular filter is maximized (see note 9 in Subheading 2.4).
2.3.5 Results and Data In this section, representative in vivo demonstrations of Bessel
Analysis focus scanning 2PFM are presented. In an SLM-based Bessel mod-
ule, using the method described in Subheading 2.3.4, the Bessel
beam of different NAs and lengths was employed to image the
cortical neurites of an awake Thy1-YFP line H mouse (Fig. 6).
With the increase of NA or axial length of the Bessel focus, more
energy is distributed to the side rings, which reduces image con-
trast. 0.4-NA Bessel foci provide high signal-to-background ratio
while maintaining the ability to resolve synapses laterally in
two-photon microscopy (Fig. 7). A single 2D scan of a Bessel
focus probes all structures from the entire volume (e.g., a depth
range of 60 μm in Fig. 7), and the calcium activities in individual
dendritic spines can be characterized (Fig. 7c). In contrast, a mini-
mum 36 scans of Gaussian beam (calculated by the ratio of axial
range of structures and Gaussian focus axial FWHM) are required
to cover the same volume (60 2D scans were used to generate
Fig. 7a). By using longer Bessel foci, imaging throughput can be
improved by more than 100 folds, with data size reduced by the
same factor.
High-Speed Neural Imaging with Synaptic Resolution 307
Fig. 6 In vivo images of cortical neurites taken with Bessel foci of different NA and axial lengths, in a head-
fixed awake Thy1-YFP mouse. An SLM-based Bessel module was used. (a) Images obtained by 2D scans of
Bessel foci and their axial point spread functions. (b) A Gaussian image stack with structures color-coded by
depth obtained with a 1.05-NA Gaussian focus. (From Ref. [15])
Fig. 7 Bessel focus scanning technology improves imaging throughput while maintaining synaptic lateral
resolution in calcium imaging of GCaMP6s+ neurites in an awake mouse brain in vivo. (a) Average intensity
projection of an image stack acquired by 60 2D scans of a Gaussian focus (1.05 NA), with structures color-
coded by depth. (b) A single 2D scan of a Bessel focus (0.4 NA, 53 μm FWHM) probes the same volume. Insets:
zoomed-in views of dendritic spines. (c) Calcium activity traces from axonal varicosities (putative boutons) and
dendritic spines (arrowheads in b). (From Ref. [15])
High-Speed Neural Imaging with Synaptic Resolution 309
Fig. 8 An axicon-based Bessel module enables 50 Hz volumetric calcium imaging of spinal projection neurons
in zebrafish larvae. (a) Image acquired by Gaussian focus scanning at 127 μm from the dorsal surface of the
head (relative depth z ¼ 17 μm). (b) Averaged calcium transients of neurons evoked by the acoustomecha-
nical tapping stimuli. (c, e, g) Volumetric images obtained by scanning a short (14 μm axial FWHM), medium
(24 μm axial FWHM), and long (39 μm axial FWHM) Bessel foci, respectively. (d, f, h) Averaged calcium
transients of responsive neurons. (i, j) Average intensity projections of a 66-μm-thick image stack acquired by
Gaussian focus scanning. Color in (i) encodes relative depth. Eleven trials were averaged in (b), (d), (f), and (h).
Shadow represents standard deviations. (From Ref. [16])
Fig. 9 Bessel focus scanning is resistant to axial motion artifacts. (a, b) Images obtained with 2D scans of a
Gaussian focus and a Bessel focus, respectively, from an awake Thy1-YFP line H mouse. (c, d) Brain motion
(upper panel, quantified as the lateral image displacement with time) causes (c) large changes of fluorescence
signal from two YFP+ dendrites (ROI1 and ROI2) in Gaussian focus scanning mode, (d) but not Bessel focus
scanning mode. (From Ref. [15])
2.4 Notes 1. The beam expander after the EOM (Fig. 2a and Subheading
2.3.1) is optional, but included in all our custom-built 2PFM
systems. The output from typical laser usually has a small (e.g.,
~1 mm) diameter and short Rayleigh length, thus diverging
quickly during propagation. Expanding the beam increases
Rayleigh length and reduces divergence during beam
propagation.
High-Speed Neural Imaging with Synaptic Resolution 311
3.2 Material and Figure 10 shows an example schematic diagram of a SIM setup.
Equipment After passing through an acousto-optic tunable filter (AOTF; AA
Opto-Electronic, AOTFnC-400.650-TN) (Fig. 11), an excitation
3.2.1 Optical
laser beam is expanded to match the active area of a spatial light
Components
modulator. To produce sinusoidal illumination patterns at the focal
plane, the laser light reflects off the spatial light modulator (SLM;
Forth Dimension Displays Ltd., QXGA-3DM) displaying binary
gratings. Two achromatic half-wave plates (HWP1 and HWP2;
Bolder Vision Optik, BVO AHWP3) and a polarizing beam splitter
(PBS; Thorlabs, PBS251) direct the laser to the SLM and maximize
its diffraction efficiency off the SLM, by ensuring the right polari-
zation direction. The diffracted light then transmits through the
PBS and has its polarization further controlled by another HWP
(HWP3; Bolder Vision Optik, BVO AHWP3) and a quarter-wave
plate (QWP; Bolder Vision Optik, BVO AQWP3), both of which
are mounted in a fast rotator (FR; Finger Lakes Instrumentation,
A24201). A dichroic mirror (D1; Semrock, Di-R405/488/561/
635-t3–25 36) reflects the illumination laser and transmits the
emitted fluorescence, and an identical compensating dichroic mir-
ror (D2, shown in Fig. 12) is used to minimize polarization scram-
bling (more details below). An objective lens (Nikon, CFI Apo
LWD 25X, 1.1 NA and 2 mm WD) is used for both illumination
and fluorescence collection. The emitted fluorescence is focused
and imaged on a camera (Hamamatsu, Orca Flash 4.0). Focal
lengths of all lenses (L1-L8) in the microscope: 150 mm,
125 mm, 400 mm, 400 mm, 175 mm, 300 mm, 85 mm, and
75 mm.
3.2.2 Fixed Mouse Brain To demonstrate structural imaging with OS-SIM, we prepared
Slices Preparation brain slices from a Thy1-GFP line M transgenic mouse (The Jack-
son Laboratory, stock 007788). After being completely sedated
with isoflurane (Piramal), the mouse was transcardially perfused
first with phosphate buffered saline (PBS, Invitrogen) followed by
4% paraformaldehyde (PFA, Electron Microscopy Sciences). The
mouse brain was dissected and immersed in 2% PFA and 15%
sucrose in PBS solution overnight at 4 C. Then the immersion
solution was replaced with 30% sucrose in PBS. After 24 h, the
Fig. 10 Detailed optical layout of a structured illumination microscope. (a) An amplitude mask in a rotational
mount selectively transmitting first-order diffraction beams. (b) Structured illumination generated by
two-beam interference at the sample plane. See Subheading 3.1 for detailed information of the optical
components
Fig. 11 Laser beam multiplexing with single-edge laser dichroic beam splitters
High-Speed Neural Imaging with Synaptic Resolution 315
3.2.3 Drosophila Larvae To demonstrate functional imaging with OS-SIM, we used trans-
Preparation genic Drosophila third instar larvae. A GCaMP6f-based postsynap-
tically targeted genetically encoded calcium indicator was expressed
in Drosophila larval muscle throughout development (genotype:
w1118; OK6-Gal4/UAS-CpxRNAi (BDSC Line #42017);
MHC-CD8-GCaMP6f-Sh/+). Larvae were dissected using a tradi-
tional semi-intact fillet preparation in HL3 solution (concentration
in mM: 70 NaCl, 5 KCl, 0.45 CaCl2·2H2O, 20 MgCl2·6H2O,
10 NaHCO3, 5 trehalose, 115 sucrose, 5 HEPES, and with pH
adjusted to 7.2) before imaging. During imaging, to maintain
viability, the larval fillet was immersed in HL3 containing 1.5 mM
CaCl2·2H2O and 25 mM MgCl2·6H2O.
3.3.1 SIM Setup The illumination and emission light paths for a structured illumi-
nation microscope follow that of a standard widefield microscope,
with added components/modules to generate and optimize the
structured (e.g., sinusoidal fringe) illumination at the sample
plane (Fig. 10). In the following subsections we discuss critical
steps in generating SIM illumination.
Maximizing Diffraction A polarizing beam splitter (PBS) is used to reflect the excitation
Efficiency and Pattern beam toward the SLM and transmit the diffracted beam traveling
Contrast at the Sample away from the SLM (Fig. 10). As the PBS reflects the s-polarized
Plane component and transmits the p-polarized component of the inci-
dent light, an achromatic half-wave plate (HWP) is placed before
the PBS to control excitation beam polarization. The HWP is
rotated to minimize the transmitted power, thus maximizing the
beam power delivered to the SLM. Our SLM has maximal diffrac-
tion efficiency for p-polarized light. Therefore, another HWP is
placed between the SLM and the PBS. With a binary pattern
displayed on the SLM, we rotate this HWP to minimize the
power in the 0th-order diffraction beam.
High modulation contrast is essential for both OS-SIM and
SR-SIM imaging. For OS-SIM, high contrast ensures strong mod-
ulation of in-focus signals, providing maximal optical sectioning.
For SR-SIM, high contrast ensures large magnitude of high fre-
quency components, supporting the extended spatial resolution.
To maximize the illumination contrast, the two interfering beams
should have s-polarization at the sample plane. In principle, one can
use a HWP set at an optimized angle for the grating orientation for
OS-SIM, or mount a HWP in a fast rotator and maintain
s-polarization for all illumination orientations used for SR-SIM.
However, in practice the s-polarization state may not be maintained
when the illumination beams reach the sample plane. This is
because optical components in the illumination light path may
alter beam polarization. For example, a dichroic mirror used to
separate excitation and emission light (Fig. 10, D1) may reflect
and transmit the p- and s-polarization components differently,
and therefore scramble the polarization of the illumination and
make an originally linearly polarized light to become elliptically
polarized.
High-Speed Neural Imaging with Synaptic Resolution 319
Fine Alignment SIM is very sensitive to misalignment, and here we provide two
useful methods to ensure perfect alignment. With both methods,
fine tip/tilt adjustment is made with two successive mirrors before
the SLM.
The first method utilizes the back-reflection pattern from the
objective lens. We first move the mask out of the light path (or flip
down if a flip mount is used). We then let the SLM display a flat
pattern so that it acts as a mirror and allows standard widefield
illumination. To achieve perfect alignment with the illumination
beam entering the objective center along its optical axis, we hold a
piece of white paper with a hole between L3 (Fig. 10) and the
objective, letting the illumination beam transmit through the hole
and reach the objective lens. We then can observe the light reflected
from the lens components inside the objective onto the white
paper. In the case of perfect alignment, the pattern appears as
concentric rings (Fig. 13a); when it is misaligned, the pattern
appears off-centered and scrambled (Fig. 13b).
With the second method, we directly observe the two interfer-
ing beams below the objective lens. For this, we let the SLM display
a grating pattern and put the mask back in path. Here we display on
the SLM a fine grating pattern so that the two diffraction orders
enter the objective lens close to its edge (typically at 80% of the NA;
see Subheading 3.3.1.2 for detailed information). In the case of
perfect alignment, the two beams below the objective appear sym-
metric in shape and with similar brightness (Fig. 14a). Choosing a
fine pattern makes any clipping of the two beams easy to detect. If
the light path is misaligned, the two beams appear asymmetric and
320 Guanghan Meng et al.
Fig. 13 Back-reflected illumination light observed before the objective. (a) Perfect alignment pattern. (b)
Misalignment pattern
3.3.2 SIM Detection Path In our system, the emitted fluorescence is collected by the same
objective and focused onto a camera for imaging (Fig. 10) with
achromatic lenses (L4–L8). Magnification from the sample plane to
the camera should ensure Nyquist sampling for the desired resolu-
tion. For SR-SIM that can double the diffraction-limited resolu-
tion, the pixel size should be smaller than a quarter of the
diffraction-limited widefield resolution.
Refined OS-SIM The basic idea of OS-SIM is that only in-focus information can be
Reconstruction Method effectively modulated. Figure 15 shows how structured illumina-
tion only preserves its contrast for structures in the focal plane (blue
arrows). When a structure is out-of-focus (orange arrows), its
fluorescence signal is not or only weakly modulated. Taking advan-
tage of this difference in signal modulation, we could computation-
ally reject out-of-focus background and retrieve in-focus
information.
One popular OS-SIM implementation was proposed by Neil
et al. [23], which requires three SI images with equally spaced
High-Speed Neural Imaging with Synaptic Resolution 321
Fig. 15 Structured illumination only modulates in-focus signals effectively. Images at two sample planes with
different structures in focus. Blue arrows: in-focus structures, orange arrows: out-of-focus structures.
Scale bar: 3 μm
Fig. 16 Widefield, basic SIM, refined OS-SIM, and two-photon fluorescence images of Thy1-GFP line M brain
slices. (a–d) Maximum intensity projections (MIPs) of 8-μm-thick widefield (WF), basic SIM, refined OS-SIM,
and 2-photon image stacks (0.1 μm Z step, 21–29 μm depth, 440 396 pixels at 86 nm pixel size),
respectively. Insets are single optical sections. Scale bar: 5 μm; insets: 1 μm
Fig. 17 In vivo functional imaging of quantal releases at the neuro-muscular junctions (NMJs) of a Drosophila
larva with OS-SIM. (a–b) Averages of widefield (WF) and OS-SIM image sequences (frames without calcium
activity) of NMJs (at a depth of 20 μm, 492 492 pixels at 86 nm pixel size). Scale bar: 5 μm; insets: 2 μm. (c)
Lateral line profiles across the structure in the insets (b, along red dashed line). (d) Spontaneous calcium
transients from 8 regions of interests (8 s of recording, orange circles in a). Widefield transients were
increased by 8 times for better visualization. (e) Averaged calcium transients over 5 events (black asterisks in
d) measured with widefield and OS-SIM
Other Optical Sectioning We described and demonstrated our refined OS-SIM method for
Reconstruction Methods in vivo structural and functional imaging in previous sections.
There are other optical sectioning reconstruction methods as well
as structured illumination strategies. For example, differential illu-
mination focal filtering (DIFF) microscopy [25], a variant of HiLo,
reconstructs one optical section from two images with structured
and complementary illumination patterns. HiLo microscopy
[24, 52], another SIM method, reconstructs one optical section
from one SI image and one uniform illumination image. HiLo has
faster imaging speed and is less sensitive to illumination distortion
and sample motion. For systems that cannot provide precise SI
translation (e.g., by an SLM), HiLo is a better option. The
OS-SIM system described here can also implement these alternative
SIM methods. A method should be chosen after considering the
application, implementation difficulty, and budget.
Other Considerations for In In addition to the often low signal-to-noise ratio, two other issues
Vivo Imaging of applying OS-SIM to in vivo imaging are sample-induced wave-
front distortion and motion-induced reconstruction artifacts. Pre-
viously [9], we demonstrated that aberration correction by adaptive
optics is essential for OS-SIM in both structural and functional
imaging of in vivo structures. For example, the imaging of the
Drosophila larva suffered from spherical aberrations coming from
the muscle layers above the focal plane and the high sucrose con-
centration immersion saline, resulting in severe reconstruction arti-
facts and abnormal calcium dynamics. All our presented results in
this section were aberration corrected. For motion artifacts, we
used a phase-corrected algorithm to remove motion-induced
reconstruction artifacts [9], which is especially important for
in vivo imaging.
Since the demonstrated 25 Hz rate is more than sufficient for
calcium imaging, we did not push for the highest imaging speed
that our camera is capable of. The imaging speed of OS-SIM is
theoretically limited by the frame rate of the camera, with the
maximum full chip (2048 2048) frame rate at 100 Hz. The
frame rate can be increased by simply reducing the line number of
readout. The camera in our system (Hamamatsu, Orca Flash 4.0)
can operate at 400 Hz at 512 lines and 800 Hz at 256 lines. By
implementing an interleaving OS-SIM reconstruction [53], the
OS-SIM frame rate equals that of the raw image frame rate. Thus,
for an imaging area of 256 2048 pixels, the 800 Hz frame rate
makes voltage imaging possible with OS-SIM.
326 Guanghan Meng et al.
4 Discussion
References
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International
License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing, adaptation, distribution
and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative Commons license,
unless indicated otherwise in a credit line to the material. If material is not included in the chapter’s Creative
Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use,
you will need to obtain permission directly from the copyright holder.
Chapter 11
Abstract
The development of all-optical techniques and analytical tools to visualize and manipulate the activity of
identified neuronal ensembles enables the characterization of causal relations between neuronal activity and
behavioral states. In this chapter, we review the implementation of simultaneous two-photon imaging and
holographic optogenetics in conjunction with population analytical tools to identify and reactivate neuronal
ensembles to control a visual-guided behavior.
1 Introduction
Eirini Papagiakoumou (ed.), All-Optical Methods to Study Neuronal Function, Neuromethods, vol. 191,
https://doi.org/10.1007/978-1-0716-2764-8_11, © The Author(s) 2023
331
332 Luis Carrillo-Reid et al.
different brain areas [4, 14–18], paving the pathway to design more
sophisticated experiments to understand complex mental states in
health and disease [19].
The optimal design of all-optical experiments to control
learned behaviors with single-cell resolution requires the simulta-
neous reading and writing of neuronal activity. This could be
achieved in many ways [20], but in this chapter we will focus mainly
on scanning and parallel optical techniques guided by analytical
tools. We describe the implementation of scanning two-photon
imaging and parallel two-photon optogenetics using a spatial light
modulator (SLM). We also describe the main concepts necessary to
identify and target neurons with pattern completion capability
[6, 21] that can recall neuronal ensembles related to a visually
guided behavior [4].
2.1 Background Optical methods are powerful to record and manipulate neuronal
activity at single-cell resolution across a large population—a key
2.1.1 Light-Sensitive
requirement to investigate neuronal ensembles. Comparing with
Sensors and Actuators of
electrophysiology, optical methods can simultaneously sample and
Neuronal Activity
target a large group of neurons with high spatial specificity in a
noninvasive manner and can be conveniently implemented for
in vivo studies. The cornerstone of optical methods is the optical
microscope, which uses light to record and manipulate the neurons
that are illuminated. As the neurons in the brain do not typically
respond to light, they can be loaded or transfected with light-
sensitive sensors and actuators that can respectively report and
manipulate their neuronal activity upon light illumination. Using
these neuronal activity sensors and actuators, an optical microscope
can simultaneously read and write neuronal activity across a large
field of view with high spatial specificity.
Calcium [22–26] and voltage indicators [27, 28] are two com-
monly used sensors for neuronal activity. These indicators are
embedded with fluorophores, which can absorb the illumination
or excitation light, and emit fluorescence. The efficiency of this
light-absorption, fluorescence-emission process is modulated by
the intracellular calcium concentration and membrane potential
respectively in calcium indicators and voltage indicators (See also
Chaps. 1 and 2). Thus, the neuronal activity can be deduced from
the time-lapse recording of the fluorescence in individual cells. The
commonly used calcium indicators include GCaMP6 [26],
jGCaMP7 [29], jRGECO [30], etc. Compared with calcium indi-
cators, voltage indicators are less mature with smaller signal-to-
noise ratio and prone to photo-bleach though much progress has
been made in the past years [31] (Chap. 2). The counterparts of
All-Optical Probing of Neuronal Ensembles 333
2.1.2 One-Photon and Depending on how light is interacting with the light-sensitive
Multiphoton Excitation sensors and actuators, optical methods can be classified into two
categories: one-photon and multiphoton (i.e., two-photon
[36, 37] or three-photon [38, 39]). In one-photon case, the inter-
action is linear. The fluorescence emission rate of the sensors and
the actuation strength of the actuators are proportional to the
intensity of the excitation light before saturation. While
one-photon excitation is straightforward, it lacks spatial specificity
in 3D. As the excitation would form a double cone-shape pattern
along the axial direction (Fig. 1), all neurons with the sensors or
actuators within the double cone may absorb the light. It is thus
very challenging to just excite a single neuron in a 3D volume.
Furthermore, in terms of imaging, the out-of-focus fluorescence
emission typically becomes background on the image captured at
the focal plane and thus reduces the image signal-to-noise ratio. As
the light-absorption and fluorescence-emission cycles create pho-
totoxicity, one-photon excitation pays a high price to image the
focal plane by creating phototoxicity within the entire double cone
volume. Multiphoton excitation can greatly overcome these limita-
tions, as the interaction between the light and the sensors or
actuators is nonlinear. In the two-photon case, the excitation effect
is proportional to the intensity square of the excitation light before
saturation, leading to a strong gradient of the excitation effect
along the optical axis in the double cone. Thus, it is feasible to
control the incident light so only the light intensity at the focal
point (i.e., the tip of the double cone) is strong enough to excite
the neuronal sensors or actuators (Fig. 1). This greatly improves the
spatial specificity or resolution and reduces the phototoxicity in
out-of-focus region. Another advantage of two-photon excitation
is that the excitation light has a longer wavelength, which reduces
the light scattering effect in scattering brain tissues. Thus
two-photon excitation can penetrate much deeper in the brain
compared with one-photon excitation. The drawback of
two-photon excitation is that much more laser power is required
due to a lower efficiency of multiphoton absorption. As the excita-
tion light eventually turns into heat, more heat will be generated
inside the brain. This is typically not a concern as in a typical
imaging experiment, the laser power used is less than the brain
damage threshold [40], even at very deep layers. In the past two
334 Luis Carrillo-Reid et al.
Fig. 1 Comparison between one-photon and two-photon excitation. In one-photon excitation (a), a double
cone is formed by focusing the visible excitation light (indicated by blue color); fluorophores (green) in the
entire double cone could be excited. In two-photon excitation (b), while a double cone is still formed by the
focus of infrared excitation light (indicated by red color), the fluorescence generation (green) is localized at the
vicinity of the focal region. Reprinted with permission from [101], Springer Nature. Experimental illustration of
the one-photon and two-photon excitation (0.16 NA for both cases) is shown in the bottom panel. (Reprinted
with permission from Ref. [102], Springer Nature)
2.1.3 Basic Setup of Since the multiphoton absorption rate is generally low, and it has a
Multiphoton Microscope nonlinear relationship with the light intensity, a femtosecond laser
is required for multiphoton excitation. The femtosecond laser deli-
vers a periodic pulse train, typically with a repetition rate of
1~80 MHz. The temporal width of a pulse is below 300 fs, yielding
a very high instantaneous peak power and thus a high multiphoton
absorption rate. Unlike one-photon excitation where a large area or
volume of sample could be simultaneously illuminated and their
fluorescence could be detected through a camera in the case of
imaging, the illumination of multiphoton light is typically through
a rapid scanning of the laser beam on the sample. In multiphoton
imaging, a single pixel detector (versus pixel array in a camera) is
used to record the emitted fluorescence (see also Chap. 2). By
correlating the temporal signal from the detector and the scanning
trajectory, the image can be built. The microscope setup (Fig. 2a)
typically includes a raster lateral (xy) scanning system composed of a
galvanometer scanner and a resonant scanner (30~60 Hz frame
rate), or two galvanometer scanners (4~10 Hz frame rate). For
volumetric imaging, an axial (z) scanning system is implemented
by adding a piezo-electric controller on the objective lens or insert-
ing axial focusing devices, such as electrically tunable lens [42],
spatial light modulator [43], or remote focusing unit [44–46] (~
ms focus switching time). Different axial planes are sequentially
scanned (Fig. 2b). While this configuration of lateral and axial
scanning is straightforward, it may not be the most efficient as it
blindly samples the brain tissue. There could be a large volume of
“empty” extracellular space without neurons, and among all the
neurons, typically only a subset of them are labeled with the calcium
indicators. Random access scanning techniques [47–51] can over-
come this issue. The laser focus spot can jump rapidly between
different regions in the sample in 3D (<20 μs transition time
between different regions) (Fig. 2c). This is enabled by acousto-
optic deflectors (AODs). The challenges of this technique are to
overcome the spatial and temporal distortion of the ultrashort laser
pulses, caused by the angular dispersion from the AODs’ phase
grating and the group delay dispersion of the AOD crystals, respec-
tively. These limit the imaging field of view. While these distortions
can be compensated [51], the optical setup is complex. AODs also
have a limitation of high insertion loss. Recently, many advanced
scanning mechanisms including beam multiplexing [43, 52–55] are
proposed and demonstrated to increase the imaging throughput.
For extensive reviews, see [56–59].
336 Luis Carrillo-Reid et al.
Fig. 3 Simultaneous two-photon calcium imaging and two-photon holographic optogenetics. (a) Two-photon
microscope setup with two independent beam paths for calcium imaging and photostimulation. An electrically
tunable lens is equipped in the imaging path so the focal plane can be rapidly switched for high-speed
volumetric imaging. The dichroic mirror 3 is used to spectrally separate the fluorescence into green and red
channel. In the photostimulation path, a half-wave plate is placed before the spatial light modulator to align
the polarization of the laser to the active axis of the spatial light modulator. The spatial light modulator then
creates a hologram in the sample to photostimulate the regions of interest (e.g., a group of neurons). A pair of
relay lens (4f system) is used to transfer the light field to a set of galvanometer mirrors which can spirally scan
the holographic pattern. At the intermediate plane of this pair of relay lens, a zeroth-order beam block is used
to block the residue light that is not modulated by the spatial light modulator. The imaging laser and
photostimulation laser have different wavelengths, and they are combined through the dichroic mirror 1 before
being delivered to the sample through the scan lens, tube lens, and objective lens. HWP half-wave plate, PMT
photomultiplier tube. (b) Schematics for simultaneous volumetric calcium imaging and 3D holographic
patterned photostimulation in mouse cortex. (c) An exemplary field of view showing neurons co-expressing
GCaMP6s (green) and C1V1-mCherry (magenta). (Reprinted and adapted from Ref. [69])
Fig. 4 Computer-generated hologram. (a) Principle of computer-generated hologram. The collimated laser
beam is incident onto the spatial light modulator (SLM), which spatially modulates the wavefront of the light.
The light field then propagates through a lens and forms the desired 3D image in the imaging domain through
interference. f, focal length of the lens. (b) By modulating the spatial phase profile at the back focal plane of
the object lens, different focal spot patterns can be formed in the imaging domain. (Reprinted with permission
from Ref. [103], Elsevier). (c) Example of a two-photon SLM hologram. The four panels illustrate the binary
target image, the spatial light modulator phase hologram generated by Gerchberg-Saxton algorithm, the
squared image (to mimic two-photon excitation) of the projected pattern back calculated from the phase
hologram, and the experimentally measured two-photon fluorescence image generated by the SLM. A stylized
picture of Cajal is used as a target image. (Reprinted from Ref. [11], Frontiers)
Fig. 5 Gerchberg-Saxton algorithm. This algorithm is used to calculate the phase hologram φs for the spatial
light modulator (SLM) based on the amplitude of the target pattern At in the imaging space and the amplitude
of the incident light field As onto the SLM
Table 1
Analytical expression of the phase hologram and the Zernike polynomials and Zernike coefficients in
superposition algorithm
M
P
A i e 2πj fx i uþy i vþ½Z 2 ðu,vÞC 2 ðzi ÞþZ 4 ðu,vÞC 4 ðzi ÞþZ 6 ðu,vÞC 6 ðzi Þg
0 0 0 0 0 0
Phase hologram on the SLM: ϕðu, v Þ ¼ phase
i¼1
[xi, yi, zi] (i¼1,2. . .M ), the coordinate of the cell body centroid (M targeted cells in total); Ai, the
electrical field weighting coefficient for the ith target (which controls the laser power it receives);
Z 0m ðu, v Þ and C 0m ðz i Þ, the Zernike polynomials and Zernike coefficients, respectively, which sets the
defocusing and compensates the first-order and second-order spherical aberration due to defocusing.
pffiffiffi
Zernike polynomials (defocus) Z 02 ðu, vÞ ¼ 3 2 u2 þ v 2 1
2
Zernike coefficients (defocus) pffiffi α 1 þ 1 sin 2 α þ 9 sin 4 α þ 1 sin 6 α þ . . .
C 02 ðz Þ ¼ nkz8πsin3 4 80 16
2.2.3 Spiral Scan Versus As mentioned in Subheading 2.1, photostimulation can be per-
Scanless Approach in formed by raster or spirally scanning the laser focal spot across the
Holographic neuron, or by projecting a disk pattern which matches the mor-
Photostimulation phology of the neuron so the entire neuron can be stimulated at
once. The same applies in holographic photostimulation. In the
first approach, a group of focal spots are created in the hologram,
and each focal spot lies on the centroid of the individual targeted
neuron [2, 12, 69]. A set of galvanometer scanners then spirally
scan the entire group of focal points (with certain repetitions), so
each focal spot spirally scans across the corresponding targeted
neuron. In the second approach, a group of disk patterns are
created in the hologram and projected to the brain tissue [72, 77,
78]. Each disk spatially overlaps with a targeted neuron. Since the
hologram already generates multiple disks for all the target
All-Optical Probing of Neuronal Ensembles 343
neurons, this method can work without scanners, and thus this is a
scanless approach. In two-photon excitation, these two approaches
have their own advantages and limitations. The spiral scan approach
spatially concentrates the laser intensity into focal spots; as
two-photon excitation effect is proportional to light intensity
square, this approach has a high excitation efficiency. As the entire
neuron is not photostimulated at the same time, to reach the
threshold of evoking action potentials, it relies on the accumulation
of the excitation effect along the spiral trajectory. The decay con-
stant of the opsin kinetics (tau-off) should thus be longer than the
duration of each spiral (i.e., the opsin channels can stay open during
the one single spiral scan, which could be less than 1 ms); otherwise
the excitation effect cannot be accumulated. The scanless approach,
on the other hand, disperses the light intensity across the entire
neuron, and thus it requires a higher average power to reach the
same excitation strength as the spiral scan approach. However, as
the entire neuron is being stimulated at once, it does not pose a
limitation on the kinetics of the opsin. It is also expected that the
jitter of the delay between the onset of photostimulation and onset
of action potential is smaller. As the opsin decay constant (tau-off)
is typically larger than the duration of a single spiral scan (which can
be <1 ms), the scanning approach is in favor as it takes lower laser
power to photoactivate the neurons. In our experiments, we used
C1V1 as the opsin. Using spiral scan, it takes about 1.8 less power
to photoactivate the neurons than the scanless approach, with the
same photostimulation duration (Fig. 6) [69]. For opsins with
faster kinetics, a faster spiral scan or the scanless approach may be
preferred.
2.2.4 Detailed In this section, we explain the details of how the microscope can be
Implementation of the constructed with two beam paths for two-photon imaging and
Microscope for two-photon holographic photostimulation (Fig. 3a). We aim to
Simultaneous Two-Photon help the readers understand the inside of a commercial microscope
Imaging and Two-Photon system and meanwhile provide basic guidelines for those who want
Holographic to home-build the microscope. Here, we use GCaMP6 as calcium
Photostimulation indicators, and C1V1-mCherry as the red-shifted opsin. The
mCherry can be used to indicate if the opsin is expressed in each
neuron. We combine the holographic illumination with spiral scan
approach to minimize the laser power onto the brain.
A Stimulation 20 ms 10 ms 5 ms 1 ms
Duration
4 mW 5.6 mW 8 mW 18 mW
Spiral Scan
1X Relative Power
0.8
Normalized ∆ F/F
0.6
0.4
0.2
Spiral Scanless Scanless Spiral Scanless Scanless Spiral Scanless Scanless Spiral Scanless Scanless
Modality
Scan Disk Disk Scan Disk Disk Scan Disk Disk Scan Disk Disk
Relative
Power 1x 1x 1.8x 1x 1x 1.8x 1x 1x 1.8x 1x 1x 1.8x
Duration 20 ms 10 ms 5 ms 1 ms
C 1
0.8
Normalized ∆ F/F
0.6
0.4
0.2
Modality Spiral Scanless Scanless Spiral Scanless Scanless Spiral Scanless Scanless Spiral Scanless Scanless
Scan Disk Disk Scan Disk Disk Scan Disk Disk Scan Disk Disk
Relative
Power 1x 1x 1.8x 1x 1x 1.8x 1x 1x 1.8x 1x 1x 1.8x
Duration 20 ms 10 ms 5 ms 1 ms
Fig. 6 Comparison between spiral scan and scanless holographic approaches for photostimulation. In the
scanning approach, the laser spot is spirally scanned over the cell body; in the scanless approach, a disk
pattern is generated by the SLM, covering the entire cell body at once. (a) Photostimulation triggered calcium
response of a targeted neuron in vivo at mouse layer 2/3 of V1, for different stimulation modalities. For each
All-Optical Probing of Neuronal Ensembles 345
Fig. 6 (continued) modality, the multiplication of stimulation duration and the square of the laser power was
kept constant over four different stimulation durations. The average response traces are plotted over those
from the individual trials. (b) ΔF/F response of individual neurons on different photostimulation conditions
(layer 2/3 of V1, over a depth of 100 ~ 270 μm from pial surface; one-way ANOVA test). For each neuron and
each stimulation duration, the laser power used in the scanless disk modality is 1 and 1.8 times relative to that
in the spiral scan. For each neuron and each modality, the multiplication of the stimulation duration and the
square of the laser power was kept constant over four different stimulation durations. (c) Boxplot summarizing
the statistics in (b). The central mark indicates the median, and the bottom and top edges of the box indicate
the 25th and 75th percentiles, respectively. The whiskers extend to the most extreme data points (99.3%
coverage if the data are normal distributed) not considered outliers, and the outliers are plotted individually
using the “+” symbol. In this experiment, the mice were transfected with GCaMP6f and C1V1-mCherry.
Repetition rate of the photostimulation laser is 1 MHz. The spiral scan consists of 50 rotations to cover the
neuronal cell body, and the scanning speed is adjusted to make different stimulation durations. (Reprinted
with permission from Ref. [69])
346 Luis Carrillo-Reid et al.
Coordinate Calibration
axial focus offset between the two beam paths. An affine trans-
formation can be extracted to map the coordinates between the
hologram generation algorithm and the actual imaging system.
This can be repeated for a few defocusing depths set in the
spatial light modulator, and a linear interpolation of the
mapping can be applied for the depths in between.
(4.2) Due to the chromatic dispersion and finite pixel size of SLM,
the SLM’s beam steering efficiency, also called diffraction effi-
ciency, drops with larger angle, leading to a lower beam power
for targets further away from the center field of view (in xy),
and nominal focus (in z). This efficiency drop can be analyti-
cally calculated [55, 75] or experimentally measured by raster
scanning the photostimulation beam on the autofluorescence
slide and detecting the fluorescence strength. A linear compen-
sation can be applied in the weighting coefficient among differ-
ent points in the target pattern to counteract this
non-uniformity (see Note 3).
Optogenetics Experiment
3.1.1 Multidimensional It has been demonstrated that the activity of neuronal ensembles
Reduction Techniques could be defined as an array of multidimensional population vec-
Applied to Population tors, where the dimensionality of the array is given by all recorded
Recordings cells in the field of view [5, 8, 82]. Similar population vectors could
be visualized by diverse computational techniques that define clus-
ters in a reduced dimensional space, where each cluster depicts a
neuronal ensemble [5, 83]. A neuronal ensemble represents a
group of neurons with coordinated activity that repeats at different
points in time [10]. The characterization of brain states using
multidimensional population vectors is independent of the record-
ing length and could be implemented in chronic experiments where
the activity of identified neuronal ensembles could be compared at
different days.
3.1.2 Targeting Two experimental designs are possible to control behavior target-
Visualized Neuronal ing visualized neuronal populations (see Note 5). One solution is to
Populations with Two- target all available neurons that could respond to a given behavioral
Photon Optogenetics cue with single-cell precision [14, 16, 18]. The other solution is to
target neurons with pattern completion capabilities that could
recall physiologically relevant neuronal ensembles related to behav-
ior [4, 6, 10, 19, 21].
In cortical circuits, the synchronous activation of several neu-
rons could simultaneously result in two unwanted scenarios: (i) the
generation of epileptiform activity or (ii) the forced engagement of
GABAergic circuits. In both scenarios, the physiological recalling of
a specialized neuronal ensemble is compromised (see Note 5).
On the other hand, the identification and targeting of neurons
with pattern completion capabilities could keep the balance
between excitation and inhibition inherent to the microcircuit
under study, allowing the recalling of physiological neuronal
ensembles that work as attractors [84] evoking a given behavior
[84, 85].
3.2 Implementation In this part of the chapter, we describe the procedures and concepts
of Analytical Methods to analyze population activity extracted from calcium imaging
to Recall Neuronal recordings. We will focus on the steps after individual neurons
Ensembles Relevant to have been identified from imaging recordings and their activity
a Learned Behavior has been inferred from calcium transients.
Fig. 7 Spike inference from holographic calcium imaging recordings. (a) Representative experiment where two
focal planes were recorded simultaneously using holographic two-photon microscopy. (b) Regions of interest
(ROIs) detected from the recordings in a. Each ROI corresponds to an identified neuron. (c) Representative
changes in fluorescence obtained from ROIs shown in b. (d) Binary arrays representing the activity of one
neuron obtained from inferred spikes. (Modified from Ref. [5, 10])
and 0’s indicate silent periods [82, 86]. Recently, several indepen-
dent research groups have released open-source code to preprocess
calcium imaging recordings to extract activity information from a
series of images [87–90].
There are four main steps that need to be considered before
performing population analysis on binary arrays (Fig. 7):
(i) Motion correction: Lateral displacements are commonly cor-
rected using TurboReg [91], fast Fourier transform [92] or
field approaches [93]. Translations in depth should be avoided
or the data should be discarded.
(ii) Identification of neurons: Automatic identification of regions of
interest (ROI’s) can be performed using different approaches
[87, 88].
(iii) Determination of changes in fluorescence: After the identifica-
tion of ROI’s, fluorescence traces for each neuron should be
extracted.
(iv) Activity inference from calcium transients: The first time deriv-
ative, deconvolution, or fitting could be used to infer spiking
activity from calcium transients [82, 87, 88]. Note that infer-
ring co-activity between pairs of neurons from raw calcium
transients introduces an artifact from the decaying phase of
the changes in fluorescence. Such common mistake generates
spurious correlations between neurons as demonstrated
before [7].
3.2.2 Binary Arrays from To construct binary arrays, inferred activity should be threshold
Inferred Activity usually 3 S.D. above noise level (Fig. 7d). This procedure has been
shown to match bursting activity of neurons above 2 action poten-
tials [82]. The onsets from inferred activity can be used to represent
the overall population activity. A binary matrix [N T] is con-
structed, where N denotes the total number of neurons in the field
of view and T represents the total number of time points recorded.
350 Luis Carrillo-Reid et al.
a b population vectors
1 0 1 0 0 1 0 0 0 1
a
similarity index
t1
1
0
1
0
1
0
1
0
1 1
0
1 1
0
1 1
0
b
1 1 1
vectors dimensionality
0 1 0 1 0 0 0 1 0 0
1 0 1 0 0 1 0 0 0 1
0 0 0 0 0 0 0 0 0 0
frames in time
0 0 0 0 1 0 1 0 1 0
0 0 0 0 0 0 0 0 0 0
neuron
t2 1 1 1 1 1 1 1 1 1 1
0 1 0 1 0 0 0 1 0 0
1 0 1 0 0 1 0 0 0 1
0 0 0 0 0 0 0 0 0 0
population vector
θ
0 0 0 0 1 0 1 0 1 0
...
0 0 0 0 0 0 0 0 0 0
0 1 0 1 0 0 0 1 0 0
1 0 1 0 0 1 0 0 0 1
0 0 0 0 1 0 1 0 1 0
tn 0 1 0 1 0 0 0 1 0 0 neuronal ensemble
t1 t2 population vector tn cluster of vectors
n dimensions
Fig. 8 Neuronal ensembles represented as multidimensional population vectors. (a) Schematic representation
of different frames where active neurons are shown in black-filled circles (left). Binary matrix illustrating the
overall population activity, where each row is a neuron and each column denotes a population vector. The total
number of recorded neurons gives the dimensionality of the array. (b) Neuronal ensembles representing
recurrent groups of neurons that are active at different times can be understood as clusters of similar
population vectors in a multidimensional space. The cosine similarity could be used as a metric to calculate
the angle between population vectors. If the angle is close to 0 , the population vectors are similar, therefore
almost the same neurons fired at different times. If the angle is close to 90 , the population vectors are
different. (Reprinted from Ref. [10])
Each row in the binary matrix depicts the activity of one neuron and
each column in the binary matrix represents the activity profile of a
neuronal population (Fig. 8a). To visualize the overall network
activity, the binary matrix can be plotted as a raster plot where 1’s
are dots. The population synchronicity can be extracted from the
time histogram of the raster plot [5, 8].
3.2.4 Similarity It has been shown that the use of significant population vectors can
Measurements on be used to discriminate similar patterns of activity repeated at
Population Vectors different times [4–6, 8]. The representation of network activity as
population vectors allows an exhaustive comparison of all popula-
tion vectors to visualize different experimental conditions. Similar-
ity maps of population vectors represent a valuable tool to visualize
recurrent activity of neuronal ensembles. To construct similarity
maps, all possible combinations of vector pairs need to be com-
puted. Since population vectors representing the activity of a given
group of neurons point to a similar place in a multidimensional
space, the angle between population vectors could be used to create
a similarity map [8, 82]. The cosine of the angle between a pair of
population vectors is defined by their normalized dot product: cos
(θ)¼V1 l V2 / llV1ll llV2ll. Thus, if the angle is close to zero, the
vectors are similar whereas if the vectors are different, they tend to
be orthogonal (Fig. 8b).
3.2.5 Identification of Similarity maps constructed from all possible vector pairs could be
Neuronal Ensembles from understood as a low-dimensional representation of the original
Population Vectors network activity, where the angles between all population vectors
as a function of time are highlighted. Thus, high similarity values in
the same row denote recurrent groups of neurons firing together. A
cluster of population vectors pointing in a similar direction gives
the definition of a neuronal ensemble using the population vector
approach. Similarity maps can be transformed to a binary matrix S
of size [T T] (Fig. 9a, b). The factorization of S using singular
value decomposition (SVD) factorizes S as the multiplication of
three factors: V, ∑, and VT, where V and VT are orthonormal basis
and ∑ contains the singular values [5, 8]. To detect neuronal
ensembles from recurrent patterns of activity observed in the
matrix S, the rate of singular values decay is determined (Fig. 9c).
Taking the singular values above chance levels can reproduce the
original matrix S with high accuracy (Fig. 9d). Then each factor of
the SVD decomposition represents a linearly orthogonal compo-
nent and each factor defines a neuronal ensemble (Fig. 9e). In the
case of primary visual cortex, each neuronal ensemble represents a
different orientation of drifting-gratings, a different natural scene
or the Go signal from a visually guided Go/No-Go task [4, 5,
7]. Another approach to identify neuronal ensembles is the projec-
tion of the multidimensional population vectors into a low dimen-
sional space, in such reduced dimensional space clusters of
population vectors depicted with cluster analysis define neuronal
ensembles or network states (Fig. 8b) [82]. The advantage of using
SVD factorization is that the total number of neuronal ensembles
could be systematically detected from the magnitude of the singular
values. Thus, recurrent population vectors can be assigned to a
given neuronal ensemble and population vectors with low repeti-
tion rate are excluded.
352 Luis Carrillo-Reid et al.
a b c d SVD σ1 1 1 +...+ σ6 6 6
120
1 data
shuffled
similarity index
2X shuffled
magnitude
vectors (t)
vectors (t)
vectors (t)
60
1 cutoff (6th)
1
0 ∗
1 200 400 1 200 400 1 5 10 30 1 200 400
vectors (t) vectors (t) log(singular value) vectors (t)
e σ1 σ2 σ3 σ4 σ5 σ6
1 1 2 2 3 3 4 4 5 5 6 6
Fig. 9 Neuronal ensembles defined by similarity maps of population vectors using singular value decomposi-
tion (SVD). (a) Similarity map illustrating the angles between all possible pair combinations of population
vectors. Note that the recurrent activation of a given group of neurons is visualized as increased similarity
index in the same row. An angle between two vectors close to 0 represents a similarity value close to
1, whereas an angle between two vectors close to 90 represents a similarity value close to 0. (b) Binary
matrix computed from the similarity map in a, representing significant patterns of activity factorized by SVD.
Black patterns in the same row indicate recurrent coactive neurons at different times. (c) Magnitude of
singular values used to define the number of neuronal ensembles repeated above chance levels. Red line
indicates the double of singular values from shuffled data. The cutoff indicates the number of neuronal
ensembles that account for >90% of the data. In general microcircuits of ~100 neurons stimulated with four
drifting-gratings can be defined by ~6 neuronal ensembles. (d) Binary matrix of the first 6 factors obtained
from SVD of b. (e) Factors from SVD that reproduce the overall response of the imaged focal plane to visual
stimuli. Bars on top indicate orientations of drifting-gratings. Empty squares represent spontaneously active
neuronal ensembles that appeared in the absence of visual stimuli. (Modified from Ref. [5])
3.2.6 Pattern Completion In the neuronal ensemble framework, pattern completion refers to
Properties of Neuronal the ability to recall a whole neuronal ensemble by the activation of
Ensembles few neurons with strong functional connectivity [6]. It has been
recently shown that the stimulation of the same neuronal popula-
tion for several times (~100) was able to imprint an artificial neuro-
nal ensemble in layer 2/3 of primary visual cortex of awake head
fixed mice [6]. Imprinted ensembles were composed by neurons
that had low probability to fire together, but after the imprinting
protocol, the stimulated neurons fired spontaneously in the absence
of stimuli (Fig. 10a). The mechanism governing the creation of
artificial neuronal ensembles could be explained by Hebbian synap-
tic plasticity [94], where the connectivity between neurons firing
together was strengthened (Fig. 10b). Indeed, neurons responding
to a given drifting-grating or belonging to a neuronal ensemble
have increased probability to be connected [8, 95]. Imprinted
All-Optical Probing of Neuronal Ensembles 353
Fig. 10 Imprinting and recalling of artificial neuronal ensembles in layer 2/3 of primary visual cortex. (a) Spatial
map of neurons activated with two-photon optogenetics several times (~100). Red neurons represent
repeatedly responding neurons (left). Scale bar: 50 μm. Imprinted ensemble (red neurons) shows spontaneous
activity in the absence of any stimuli (middle). The activation of one neuron with pattern completion capability
(blue arrow) was able to recall the imprinted ensemble (red neurons). (b) Cartoon illustrating how imprinted
and recalled ensembles could be explained by the strengthening of connections of pre-existing ensembles and
photostimulated neurons. (Modified from Ref. [10])
3.2.7 Recalling Neuronal To prove the causal relation between neuronal activity and a learned
Ensembles Related to behavior, it is necessary to identify representative members of neu-
Behavior ronal ensembles that could trigger the behavior [81]. During the
behavioral training phase photostimulation should be avoided (see
Note 6). It has been recently shown, in a visually guided Go/No-
Go task, that neuronal ensembles responding to drifting-gratings in
layer 2/3 of primary visual cortex increase their reliability for the
Go signal whereas reduce their reliability for the No-Go signal
[4]. Increased reliability is reflected as strengthened functional
connectivity of the Go ensemble allowing neurons with pattern
completion capabilities [21] to recall the whole Go ensemble and
trigger the perception of the Go signal evoking the learned behav-
ior (Fig. 11).
a
Go signal Go ensemble
visual stimuli layer 2/3 V1 licking
b
p.c neurons recall ensemble
opto stimuli layer 2/3 V1 licking
5 Notes
6 Outlook
Acknowledgments
References
1. Alivisatos AP, Chun M, Church GM, Green- term synaptic plasticity in cortical networks.
span RJ, Roukes ML, Yuste R (2012) The Int J Neural Syst 25:1550026
brain activity map project and the challenge 9. Yuste R (2015) From the neuron doctrine to
of functional connectomics. Neuron 74: neural networks. Nat Rev Neurosci 16:
970–974 487–497
2. Packer AM, Russell LE, Dalgleish HW, Haus- 10. Carrillo-Reid L, Yang W, Kang Miller JE,
ser M (2015) Simultaneous all-optical manip- Peterka DS, Yuste R (2017) Imaging and
ulation and recording of neural circuit activity optically manipulating neuronal ensembles.
with cellular resolution in vivo. Nat Methods Annu Rev Biophys 46:271–293
12:140–146 11. Nikolenko V, Watson BO, Araya R,
3. Rickgauer JP, Deisseroth K, Tank DW (2014) Woodruff A, Peterka DS, Yuste R (2008)
Simultaneous cellular-resolution optical per- SLM microscopy: scanless two-photon imag-
turbation and imaging of place cell firing ing and photostimulation with spatial light
fields. Nat Neurosci 17:1816–1824 modulators. Front Neural Circuit 2:1–14
4. Carrillo-Reid L, Han S, Yang W, Akrouh A, 12. Packer AM, Peterka DS, Hirtz JJ, Prakash R,
Yuste R (2019) Controlling visually guided Deisseroth K, Yuste R (2012) Two-photon
behavior by holographic recalling of cortical optogenetics of dendritic spines and neural
ensembles. Cell 178(447–457):e445 circuits. Nat Methods 9:1202–1205
5. Carrillo-Reid L, Miller JE, Hamm JP, 13. Pandarinath C, O’Shea DJ, Collins J,
Jackson J, Yuste R (2015) Endogenous Jozefowicz R, Stavisky SD, Kao JC, Traut-
sequential cortical activity evoked by visual mann EM, Kaufman MT, Ryu SI, Hochberg
stimuli. J Neurosci 35:8813–8828 LR et al (2018) Inferring single-trial neural
6. Carrillo-Reid L, Yang W, Bando Y, Peterka population dynamics using sequential auto-
DS, Yuste R (2016) Imprinting and recalling encoders. Nat Methods 15:805–815
cortical ensembles. Science 353:691–694 14. Marshel JH, Kim YS, Machado TA, Quirin S,
7. Miller JE, Ayzenshtat I, Carrillo-Reid L, Yuste Benson B, Kadmon J, Raja C,
R (2014) Visual stimuli recruit intrinsically Chibukhchyan A, Ramakrishnan C, Inoue M
generated cortical ensembles. Proc Natl Acad et al (2019) Cortical layer-specific critical
Sci U S A 111:E4053–E4061 dynamics triggering perception. Science 365:
8. Carrillo-Reid L, Lopez-Huerta VG, Garcia- eaaw5202
Munoz M, Theiss S, Arbuthnott GW (2015) 15. Daie K, Svoboda K, Druckmann S (2021)
Cell assembly signatures defined by short- Targeted photostimulation uncovers circuit
358 Luis Carrillo-Reid et al.
motifs supporting short-term memory. Nat 28. Peterka DS, Takahashi H, Yuste R (2011)
Neurosci 24:259–265 Imaging voltage in neurons. Neuron 69:9–21
16. Dalgleish HW, Russell LE, Packer AM, 29. Dana H, Sun Y, Mohar B, Hulse BK, Kerlin
Roth A, Gauld OM, Greenstreet F, Thomp- AM, Hasseman JP, Tsegaye G, Tsang A,
son EJ, Hausser M (2020) How many neu- Wong A, Patel R et al (2019) High-
rons are sufficient for perception of cortical performance calcium sensors for imaging
activity? elife 9:e58889 activity in neuronal populations and micro-
17. Gill JV, Lerman GM, Zhao H, Stetler BJ, compartments. Nat Methods 16:649–657
Rinberg D, Shoham S (2020) Precise holo- 30. Dana H, Mohar B, Sun Y, Narayan S,
graphic manipulation of olfactory circuits Gordus A, Hasseman JP, Tsegaye G, Holt
reveals coding features determining percep- GT, Hu A, Walpita D et al (2016) Sensitive
tual detection. Neuron 108(382–393):e385 red protein calcium indicators for imaging
18. Robinson NTM, Descamps LAL, Russell LE, neural activity. elife 5:e12727
Buchholz MO, Bicknell BA, Antonov GK, 31. Bando Y, Grimm C, Cornejo VH, Yuste R
Lau JYN, Nutbrown R, Schmidt-Hieber C, (2019) Genetic voltage indicators. BMC Biol
Hausser M (2020) Targeted activation of hip- 17:71
pocampal place cells drives memory-guided 32. Adams SR, Tsien RY (1993) Controlling cell
spatial behavior. Cell 183:2041–2042 chemistry with caged compounds. Annu Rev
19. Carrillo-Reid L (2021) Neuronal ensembles Physiol 55:755–784
in memory processes. Semin Cell Dev Biol 33. Ellis-Davies GC (2007) Caged compounds:
125:136–143 photorelease technology for control of cellu-
20. Emiliani V, Cohen AE, Deisseroth K, Hausser lar chemistry and physiology. Nat Methods 4:
M (2015) All-optical interrogation of neural 619–628
circuits. J Neurosci 35:13917–13926 34. Nagel G, Szellas T, Huhn W, Kateriya S,
21. Carrillo-Reid L, Han S, O’Neil D, Taralova E, Adeishvili N, Berthold P, Ollig D,
Jebara T, Yuste R (2021) Identification of Hegemann P, Bamberg E (2003)
pattern completion neurons in neuronal Channelrhodopsin-2, a directly light-gated
ensembles using probabilistic graphical mod- cation-selective membrane channel. Proc
els. J Neurosci 41:8577–8588 Natl Acad Sci U S A 100:13940–13945
22. Tsien RY (1980) New calcium indicators and 35. Boyden ES, Zhang F, Bamberg E, Nagel G,
buffers with high selectivity against magne- Deisseroth K (2005) Millisecond-timescale,
sium and protons: design, synthesis, and genetically targeted optical control of neural
properties of prototype structures. Biochem- activity. Nat Neurosci 8:1263–1268
istry 19:2396–2404 36. Denk W, Strickler JH, Webb WW (1990)
23. Yuste R, Katz LC (1991) Control of postsyn- Two-photon laser scanning fluorescence
aptic Ca2+ influx in developing neocortex by microscopy. Science 248:73–76
excitatory and inhibitory neurotransmitters. 37. Yuste R, Denk W (1995) Dendritic spines as
Neuron 6:333–344 basic functional units of neuronal integration.
24. Miyawaki A, Llopis J, Heim R, McCaffery JM, Nature 375:682–684
Adams JA, Ikura M, Tsien RY (1997) Fluo- 38. Horton NG, Wang K, Kobat D, Clark CG,
rescent indicators for Ca2+ based on green Wise FW, Schaffer CB, Xu C (2013) three-
fluorescent proteins and calmodulin. Nature photon microscopy of subcortical structures
388:882–887 within an intact mouse brain. Nat Photonics
25. Stosiek C, Garaschuk O, Holthoff K, Kon- 7:205–209
nerth A (2003) In vivo two-photon calcium 39. Ouzounov DG, Wang T, Wang M, Feng DD,
imaging of neuronal networks. Proc Natl Horton NG, Cruz-Hernandez JC, Cheng YT,
Acad Sci U S A 100:7319–7324 Reimer J, Tolias AS, Nishimura N et al (2017)
26. Chen TW, Wardill TJ, Sun Y, Pulver SR, In vivo three-photon imaging of activity of
Renninger SL, Baohan A, Schreiter ER, Kerr GCaMP6-labeled neurons deep in intact
RA, Orger MB, Jayaraman V et al (2013) mouse brain. Nat Methods 14:388–390
Ultrasensitive fluorescent proteins for imag- 40. Podgorski K, Ranganathan GN (2016) Brain
ing neuronal activity. Nature 499:295–300 heating induced by near infrared lasers during
27. Gross D, Loew LM (1989) Fluorescent indi- multi-photon microscopy. J Neurophysiol
cators of membrane potential: microspectro- 116(3):1012–1023:jn 00275 02016
fluorometry and imaging. Methods Cell Biol 41. Rowlands CJ, Park D, Bruns OT, Piatkevich
30:193–218 KD, Fukumura D, Jain RK, Bawendi MG,
All-Optical Probing of Neuronal Ensembles 359
Boyden ES, So PTC (2017) Wide-field three- two-photon microscopy. Proc Natl Acad Sci
photon excitation in biological samples. Light U S A 110:13138–13143
Sci Appl 6:e16255 54. Wu J, Liang Y, Hsu C-L, Chavarha M, Evans
42. Grewe BF, Voigt FF, van‘t Hoff M, Helmchen SW, Shi D, Lin MZ, Tsia KK, Ji N (2019)
F (2011) Fast two-layer two-photon imaging Kilohertz in vivo imaging of neural activity.
of neuronal cell populations using an electri- bioArxiv:543058
cally tunable lens. Biomed Opt Express 2: 55. Yang W, Miller JE, Carrillo-Reid L,
2035–2046 Pnevmatikakis E, Paninski L, Yuste R, Peterka
43. Han S, Yang W, Yuste R (2019) Two-color DS (2016) Simultaneous multi-plane imaging
volumetric imaging of neuronal activity of of neural circuits. Neuron 89:269–284
cortical columns. Cell Rep 27(2229–2240): 56. Yang W, Yuste R (2017) In vivo imaging of
e2224 neural activity. Nat Methods 14:349–359
44. Anselmi F, Ventalon C, Begue A, Ogden D, 57. Ji N, Freeman J, Smith SL (2016) Technolo-
Emiliani V (2011) Three-dimensional imag- gies for imaging neural activity in large
ing and photostimulation by remote-focusing volumes. Nat Neurosci 19:1154–1164
and holographic light patterning. Proc Natl 58. Weisenburger S, Vaziri A (2018) A guide to
Acad Sci U S A 108:19504–19509 emerging technologies for large-scale and
45. Botcherby EJ, Juskaitis R, Booth MJ, Wilson whole-brain optical imaging of neuronal
T (2008) An optical technique for remote activity. Annu Rev Neurosci 41:431–452
focusing in microscopy. Opt Commun 281: 59. Mertz J (2019) Strategies for volumetric
880–887 imaging with a fluorescence microscope.
46. Botcherby EJ, Smith CW, Kohl MM, Optica 6:1261–1268
Debarre D, Booth MJ, Juskaitis R, 60. Oron D, Tal E, Silberberg Y (2005) Scanning-
Paulsen O, Wilson T (2012) Aberration-free less depth-resolved microscopy. Opt Express
three-dimensional multiphoton imaging of 13:1468–1476
neuronal activity at kHz rates. Proc Natl
Acad Sci U S A 109:2919–2924 61. Zhu G, van Howe J, Durst M, Zipfel W, Xu C
(2005) Simultaneous spatial and temporal
47. Kaplan A, Friedman N, Davidson N (2001) focusing of femtosecond pulses. Opt Express
Acousto-optic lens with very fast focus scan- 13:2153–2159
ning. Opt Lett 26:1078–1080
62. Papagiakoumou E, de Sars V, Oron D, Emi-
48. Duemani Reddy G, Kelleher K, Fink R, Sag- liani V (2008) Patterned two-photon illumi-
gau P (2008) Three-dimensional random nation by spatiotemporal shaping of
access multiphoton microscopy for functional ultrashort pulses. Opt Express 16:
imaging of neuronal activity. Nat Neurosci 11: 22039–22047
713–720
63. Papagiakoumou E, de Sars V, Emiliani V,
49. Grewe BF, Langer D, Kasper H, Kampa BM, Oron D (2009) Temporal focusing with spa-
Helmchen F (2010) High-speed in vivo cal- tially modulated excitation. Opt Express 17:
cium imaging reveals neuronal network activ- 5391–5401
ity with near-millisecond precision. Nat
Methods 7:399–405 64. Papagiakoumou E, Ronzitti E, Emiliani V
(2020) Scanless two-photon excitation with
50. Kirkby PA, Nadella KMNS, Silver RA (2010) temporal focusing. Nat Methods 17:571–581
A compact acousto-optic lens for 2D and 3D
femtosecond based 2-photon microscopy. 65. Papagiakoumou E, Anselmi F, Begue A, de
Opt Express 18:13720–13744 Sars V, Gluckstad J, Isacoff EY, Emiliani V
(2010) Scanless two-photon excitation of
51. Katona G, Szalay G, Maak P, Kaszas A, channelrhodopsin-2. Nat Methods 7:
Veress M, Hillier D, Chiovini B, Vizi ES, 848–854
Roska B, Rozsa B (2012) Fast two-photon
in vivo imaging with three-dimensional ran- 66. Mardinly AR, Oldenburg IA, Pegard NC,
dom-access scanning in large tissue volumes. Sridharan S, Lyall EH, Chesnov K, Brohawn
Nat Methods 9:201–208 SG, Waller L, Adesnik H (2018) Precise mul-
timodal optical control of neural ensemble
52. Cheng A, Goncalves JT, Golshani P, activity. Nat Neurosci 21:881–893
Arisaka K, Portera-Cailliau C (2011) Simulta-
neous two-photon calcium imaging at differ- 67. Dal Maschio M, Donovan JC, Helmbrecht
ent depths with spatiotemporal multiplexing. TO, Baier H (2017) Linking neurons to net-
Nat Methods 8:139–142 work function and behavior by two-photon
holographic optogenetics and volumetric
53. Ducros M, Goulam Houssen Y, Bradley J, de imaging. Neuron 94(774–789):e775
Sars V, Charpak S (2013) Encoded multisite
360 Luis Carrillo-Reid et al.
68. Forli A, Vecchia D, Binini N, Succol F, sub-millisecond precision. In: Optics in the
Bovetti S, Moretti C, Nespoli F, Mahn M, Life Sciences. The Optical Society of America,
Baker CA, Bolton MM et al (2018) BrM3B.4
Two-photon bidirectional control and imag- 79. Pologruto TA, Sabatini BL, Svoboda K
ing of neuronal excitability with high spatial (2003) ScanImage: flexible software for
resolution in vivo. Cell Rep 22:3087–3098 operating laser scanning microscopes. Biomed
69. Yang W, Carrillo-Reid L, Bando Y, Peterka Eng Online 2:13
DS, Yuste R (2018) Simultaneous 80. Yang W (2018). https://github.com/
two-photon imaging and two-photon opto- wjyangGithub/Holographic-
genetics of cortical circuits in three dimen- Photostimulation-System
sions. elife 7:e32671 81. Carrillo-Reid L, Yuste R (2020) Playing the
70. Yizhar O, Fenno LE, Prigge M, Schneider F, piano with the cortex: role of neuronal ensem-
Davidson TJ, O’Shea DJ, Sohal VS, Goshen I, bles and pattern completion in perception and
Finkelstein J, Paz JT et al (2011) Neocortical behavior. Curr Opin Neurobiol 64:89–95
excitation/inhibition balance in information 82. Carrillo-Reid L, Tecuapetla F, Tapia D,
processing and social dysfunction. Nature Hernandez-Cruz A, Galarraga E, Drucker-
477:171–178 Colin R, Bargas J (2008) Encoding network
71. Akerboom J, Carreras Calderon N, Tian L, states by striatal cell assemblies. J Neurophy-
Wabnig S, Prigge M, Tolo J, Gordus A, siol 99:1435–1450
Orger MB, Severi KE, Macklin JJ et al 83. Cunningham JP, Yu BM (2014) Dimension-
(2013) Genetically encoded calcium indica- ality reduction for large-scale neural record-
tors for multi-color neural activity imaging ings. Nat Neurosci 17:1500–1509
and combination with optogenetics. Front
Mol Neurosci 6:2 84. Cossart R, Aronov D, Yuste R (2003)
Attractor dynamics of network UP states in
72. Pegard NC, Mardinly AR, Oldenburg IA, neocortex. Nature 423:283–289
Sridharan S, Waller L, Adesnik H (2017)
Three-dimensional scanless holographic 85. Hopfield JJ (1982) Neural networks and
optogenetics with temporal focusing physical systems with emergent collective
(3D-SHOT). Nat Commun 8:1228 computational abilities. Proc Natl Acad Sci
U S A 79:2554–2558
73. Chen IW, Ronzitti E, Lee BR, Daigle TL,
Dalkara D, Zeng H, Emiliani V, Papagiakou- 86. Sasaki T, Matsuki N, Ikegaya Y (2007) Meta-
mou E (2019) In vivo submillisecond stability of active CA3 networks. J Neurosci
two-photon optogenetics with temporally 27:517–528
focused patterned light. J Neurosci 39: 87. Pnevmatikakis EA, Soudry D, Gao Y,
3484–3497 Machado TA, Merel J, Pfau D, Reardon T,
74. Gerchberg RW, Saxton WO (1972) A practi- Mu Y, Lacefield C, Yang W et al (2016)
cal algorithm for the determination of the Simultaneous denoising, deconvolution, and
phase from image and diffraction pictures. demixing of calcium imaging data. Neuron
Optik 35:237–246 89:285–299
75. Golan L, Reutsky I, Farah N, Shoham S 88. Mukamel EA, Nimmerjahn A, Schnitzer MJ
(2009) Design and characteristics of holo- (2009) Automated analysis of cellular signals
graphic neural photo-stimulation systems. J from large-scale calcium imaging data. Neu-
Neural Eng 6:066004 ron 63:747–760
76. Yang SJ, Allen WE, Kauvar I, Andalman AS, 89. Giovannucci A, Friedrich J, Gunn P, Kalfon J,
Young NP, Kim CK, Marshel JH, Brown BL, Koay SA, Taxidis J, Najafi F, Gau-
Wetzstein G, Deisseroth K (2015) Extended thier JL, Zhou P et al (2019) CaImAn an
field-of-view and increased-signal 3D holo- open source tool for scalable calcium imaging
graphic illumination with time-division multi- data analysis. elife 8:e38173
plexing. Opt Express 23:32573–32581 90. Pnevmatikakis EA (2019) Analysis pipelines
77. Hernandez O, Papagiakoumou E, Tanese D, for calcium imaging data. Curr Opin Neuro-
Fidelin K, Wyart C, Emiliani V (2016) Three- biol 55:15–21
dimensional spatiotemporal focusing of holo- 91. Thevenaz P, Ruttimann UE, Unser M (1998)
graphic patterns. Nat Commun 7:11928 A pyramid approach to subpixel registration
78. Mardinly A, Pegard N, Oldenburg I, based on intensity. IEEE Trans Image Process
Sridharan S, Hakim R, Waller L, Adesnik H 7:27–41
(2017) 3D all-optical control of functionally 92. Guizar-Sicairos M, Thurman ST, Fienup JR
defined neurons with cellular resolultion and (2008) Efficient subpixel image registration
algorithms. Opt Lett 33:156–158
All-Optical Probing of Neuronal Ensembles 361
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International
License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing, adaptation, distribution
and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative Commons license,
unless indicated otherwise in a credit line to the material. If material is not included in the chapter’s Creative
Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use,
you will need to obtain permission directly from the copyright holder.
Chapter 12
Abstract
Connecting neuronal activity to perception requires tools that can probe neural codes at cellular and circuit
levels, paired with sensitive behavioral measures. In this chapter, we present an overview of current methods
for connecting neural codes to perception using precision optogenetics and psychophysical measurements
of synthetically induced percepts. We also highlight new methodologies for validating precise control of
optical and behavioral manipulations. Finally, we provide a perspective on upcoming developments that are
poised to advance the field.
1 Introduction
Eirini Papagiakoumou (ed.), All-Optical Methods to Study Neuronal Function, Neuromethods, vol. 191,
https://doi.org/10.1007/978-1-0716-2764-8_12, © The Author(s) 2023
363
364 Jonathan V. Gill et al.
cellular composition and timing are exactly known. This allows the
experimenter to take a step beyond validating the participation of
individual circuits for behavior and to determine precisely what
features of the activity of neural circuits guide perception. This is
accomplished by holding individual features constant (e.g., what
neurons are stimulated), while parametrically varying other features
(e.g., timing), and performing sensitive psychophysical measure-
ments of the perceptual impact.
2.1 Using Detection The first experimental strategy is to assess the influence of specific
to Test the Relevance features within neural activity patterns on the detectability of an
of Neural Codes artificial stimulus. Animal survival relies on the detection of brief,
faint cues to signal the presence of food, mates, and predators.
A multitude of studies across species have revealed the exquisite
sensitivity of sensory systems to their preferred stimuli
[37, 38]. Exploring how critical information is conveyed by sensory
circuits at the perceptual limit provides a lens to examine the
essential coding features connecting neural activity to behavior.
However, even in this minimal regime, sensory-evoked responses
can be complex, simultaneously encoding stimuli in the identity,
rate, and timing of specific neurons. By replacing external stimuli
with targeted activation of neurons in sensory areas, we can directly
test which features of the observed activity affect the detectability of
the artificial stimulus. We can then infer the features of activity
essential for perception.
Early studies using either electrical or 1P optogenetic stimula-
tion revealed that rodents can detect changes in the spike rate of
single neurons [39, 40], or across populations of hundreds of
neurons [41] in the somatosensory and visual cortices, using sti-
muli that lasted for hundreds of milliseconds. Additional studies
have explored the role of relative spike timing [41, 42], and sensi-
tivity to latency [43, 44] in populations of hundreds to thousands
of optogenetically activated cortical neurons. These studies were
successful in demonstrating that rodents can perceive minimal
perturbations in sensory cortical neurons, and perceptibility may,
or may not, vary along certain feature dimensions, like timing.
However, the techniques used in these experiments lacked the
ability to target specific sets of neurons according to their func-
tional properties or tuning. By optogenetically labeling olfactory
sensory neurons expressing a specific receptor (M72-ChR2), a key
related study in the olfactory system revealed that mice could detect
brief activation (10 ms) of a single glomerulus [45]. This well-
defined input channel to the olfactory bulb is a site of convergence
where ~5,000 olfactory sensory neurons provide input to ~25–30
mitral and tufted cells (projection neurons), demonstrating that the
elementary features of olfactory perception operate at, or below,
this spatiotemporal scale.
Illuminating Neural Computation Using Precision Optogenetics 367
A B C
2P Imaging 2P Stimulation 5 Go
920 nm + 1028 nm Tone 1 4
6
Lick Hit
2 3 7
sound cues Sniff
8
12
9
10
No-Lick Miss
11 13
16x Photostim. 14
30 cells
No-Go
16 23
sniff signal 15
Response 18
22
19 20
Lick
False
17
Interval 25
21
Alarm
27
24 26 28
Lick 29 No-Lick
Correct
water rewards 30
WT
Reject
1s no stim.
Fig. 1 Testing detection of precision photostimulation. (a) Schematic of photostimulation detection experi-
ment. A head-fixed mouse with a chronically implanted window above the olfactory bulb is positioned in front
of a lickspout and pressure sensor to monitor respiration (sniff). (b) Trial structure for detection experiment. A
tone signaled the start of trials and photostimulation was timed to a fixed delay relative to sniff (c) Left,
neurons in the mitral cell layer (MCs and GCs) co-expressing ChrimsonR-tdTomato (red) and GCaMP6s
(green). Thirty neurons were targeted for simultaneous photostimulation (white circles). Scale bar – 40 μm.
Right, outcomes for responses to the “go” and “no-go” trials. Red circles indicate stimulation of a particular
cell, while empty circles indicate no stimulation. (Adapted from Ref. [25])
A B C inhalation
inhalation inhalation
N=9 ∆t = 10 ms
Neuron #
Neuron #
Neuron #
Decreasing Synchrony
Decreasing Number
Increasing Latency 25
N=5 ∆t = 30 ms
45
N=2 ∆t = 50 ms
Fig. 2 Independent control of multiple activity features. (a) Schematic of stimuli which vary in the number of
neurons targeted, but not the timing of activation. (b) Schematic of stimuli which vary in the synchrony across
neurons, but not the number of neurons. Stimuli were presented with a mean latency of 45 ms across
conditions. (c) Schematic of stimuli which vary in the latency of photostimulation relative to the onset of
inhalation, but not synchrony, or number of neurons. (Adapted from Ref. [25])
2.2 Technical Behavioral Task and Training The previously described experi-
Implementation of ments all made use of a similar strategy for assessing the detectabil-
Detection Experiments ity of 2P photostimulation. They used a go/no-go behavioral
paradigm, in which an animal makes a binary decision about the
presence or absence of a stimulus. In this paradigm, an animal is
trained to respond in some manner (e.g., licking, lever press, nose
poke) for a reward, or withhold a response (e.g., do not lick, press,
or poke), to signal whether a “go” stimulus cue is present or absent
(“no-go”). Typically, the “go” stimulus is randomly, or pseudo-
randomly presented on a fraction of trials (typically 0.5), where on
the remaining fraction of trials an animal experiences either a “dis-
tractor” stimulus or nothing. Animal behavior is progressively
shaped to associate the availability of reward with the presence or
absence of the stimulus. Once the task is acquired, the experimenter
can then change the parameters of the stimulus in either a trial or
blockwise manner to determine the sensitivity of animals to each
parameter based on the frequency of correct choices or error type.
Fig. 3 Spatial and temporal perturbations of synthetic odors. (a) Schematic of the experimental setup. Dorsal
olfactory bulb was exposed by a chronically implanted 3 mm window. Spatiotemporal stimulation patterns,
created by a digital micromirror device, were projected onto the olfactory bulb of a head-fixed OMP-ChR2
mouse in front of a pressure sensor for sniff monitoring, and lick spouts delivering water. (b) Schematics for
pattern discrimination task. Animals were trained to recognize Target versus Non-target patterns defined on a
stimulation grid. Target patterns comprised six spots, initialized randomly but fixed across subsequent
sessions, activated in an ordered sequence defined in time where 0 marks inhalation onset. Non-target
patterns were six off-Target spots, randomly chosen from trial to trial, with randomized timing within 300 ms
from inhalation (~single sniff). (c) Illustration of spatial perturbations: One or multiple spots in Target patterns
were randomly replaced with Non-target spots. (d) Illustration of temporal perturbations in which one or
multiple spots in Target patterns were temporally shifted. (Adapted from Ref. [30])
374 Jonathan V. Gill et al.
A B C
Photostim.
20 ms
Latency
10 ms
5mV
1 Trial
{
5 ms 5 ms Jitter
A B C
* Opsin+
Opsin+ Response
Targeted
100 1 Opsin-
250
g
200
150 50 0.75
100
∆ F/F (norm)
50 0 0.05 0.50
Targ.
*
∆ F/F
-50 0.25
0 *
-100 0
-0.05
-100 -50 0 50 100 0 50 100 150 200 250
Distance from target neuron (μm) Distance from target neuron (μm)
Targeted
1 Opsin- 1 Opsin-
g
g
0.75 0.75
∆ F/F (norm)
∆ F/F (norm)
0.25 0.25
0 0
Fig. 5 Assessing specificity of photostimulation-evoked activity. (a) Example FOV centered on the mitral cell
layer of a Tbet-Cre mouse. One mitral cell is targeted for 2P photostimulation (small white circle, 15 μm
diameter). Red labeling corresponds to FLEX-ChrimsonR-tdTomato expression limited to a subset of mitral
cells, and green labeling corresponds to pan-neuronal GCaMP6s expression. Normalized fluorescence (ΔF/F)
is averaged for every neuron in the 100 ms period following photostimulation and averaged across neurons
occupying the same radial distance from the target (example bin size ¼ 50 μm). (b) A spatial heatmap of
average response vs. ROI position centered on 49 mitral cell targets (n ¼ 2 Tbet-cre mice, 138 total neurons,
3631 photostimulations, 3 pulses per photostimulation: 10 ms on – 10 ms off, at 30 mW, or 0.19 mW/μm2).
Only cells labeled with ChrimsonR-tdTomato were included. The “targeted” bin is outlined by a black circle. (c)
The average radial decay of responses across ChrimsonR-tdTomato positive neurons (Opsin+, red) and
ChrimsonR-tdTomato negative neurons (Opsin, green). Cell responses radially binned and averaged for
each targeted neuron and bin means were averaged across targets (mean s.e.m., 49 targeted neurons,
n ¼ 2 Tbet-cre mice). Asterisks indicate a significant difference between the average binned response of
ChrimsonR+ and ChrimsonR neurons (*p < 0.05, two sample t-test, Holm-Bonferroni corrected for multiple
comparisons). (d) A cartoon of the hypothesis that photostimulation activates neighboring Opsin+ neurons due
to off-target photostimulation: Opsin+ responses will exceed the Opsin responses, reflecting inadvertent
stimulation of nearby Opsin+ neurons. (e) A cartoon of the hypothesis that single-cell photostimulation leads
to responses of nearby neurons due to network effects: Opsin+ responses do not exceed the responses of
Opsin neurons. (Adapted from Ref. [25])
380 Jonathan V. Gill et al.
**
0.1
Response (∆F/F)
vs.
0.05
Omit One Target
Target 1 Target 2 Target n
, , ... , -0.05
3.3 Assessing the While initial measurements of the responses evoked by photosti-
Reliability of mulation as well as calibrations to ensure proper targeting are
Behavioral Readout essential for experiments combining all-optical manipulations with
behavior, they are not the only controls necessary to be confident in
a behavioral result. As an additional precaution, it is extremely
useful to ensure that the observed behavioral effects depend only
384 Jonathan V. Gill et al.
Fig. 7 Visualizing light patterns in vivo with HOCUS. (a) The 2P imaging path is combined with the holographic
2P photostimulation path for in vivo experiments in head-fixed, behaving mice. The reflected stimulation light
passes through the PBS, is descanned by the mirrors, and is then reflected by the dichroic mirror through the
pinhole onto the detector. PBS polarizing beam splitter, PMT1, PMT2 photomultiplier tubes, SLM spatial light
modulator, DET detector. (b) A merged image of GCaMP6s expression (green) and the HOCUS-imaged
reflected stimulation light (red) showing a pattern of four light spots projected onto the brain, 120 μm deep
in the olfactory bulb, and positioned on four respective neurons. Scale bar: 25 μm. (Adapted from Ref. [31])
1/f
A B
Control
Stim
Power
th
2P stim
Time W Heat
W
L Indirect
effects
C D
20 trials
Detection Accuracy
300 10 ms
2P Stim
Firing rate (spikes/sec)
Control
200 0.75
100
0.5
0
-100 0 100
Time (ms)
Fig. 8 Sham-photostimulation control. (a) A schematic demonstrating the effect of tuning the laser pulse
duration. Time dependence of laser power for pulse trains with the same pulse frequency ( f ) and average
power, but different pulse durations: short pulse, τS (red), and long pulse, τL (gray). To photostimulate a cell,
laser power must exceed a certain threshold, Pth. (b) Left, a table summarizing the differences in effects
evoked by the short and long pulse duration stimuli. Right, a schematic of the behavioral setup for the sham
photostimulation control experiment. (c) Representative example raster plots (top) and peristimulus time
histograms (PSTHs) (bottom) for short pulse photostimulation (~200 fs, red) and long pulse sham photo-
stimulation (control, 15 ps, gray) (20 trials per condition, 30 mW, 10 ms illumination, n ¼ 1 cell in 1 WT
mouse). (d) Detection accuracy as a function of photostimulation condition. During the sham-photostimulation
control blocks, detection accuracy dropped to chance level (0.5 0.003, mean s.e.m., p ¼ 0.37,
one-sample t-test, 0.06–0.125 mW/μm2, n ¼ 5 mice, 2 WT (filled circles) and 3 Tbet-cre (empty circles)
and was significantly different from both pre- and post-control measurements ( p < 0.001, Fisher’s exact test,
0.06–0.125 mW/μm2, n ¼ 5 mice, 2 WT (filled circles) and 3 Tbet-cre (empty circles). (Adapted from Ref. [25])
4 Notes
5 Outlook
neurons [28]. Using the inferred structure, one could predict how
the effects of stimulation will propagate through the circuit, and the
stimulation could be designed to activate specific modes of activity,
testing their effect on perception. Ultimately, the emergence of
newly informative behavioral paradigms along with novel concep-
tual frameworks will likely be some of the most exciting outcomes
to come from the application of precision optogenetics and syn-
thetic perception.
References
1. Kandel ER, Schwartz JH, Jessell TM (2000) activity by sculpted light. Proc Natl Acad Sci
Principles of neural science, vol 4. McGraw- U S A 107(26):11981–11986
Hill, New York 14. Papagiakoumou E et al (2010) Scanless
2. Dayan P, Abbott LF (2001) Theoretical neuro- two-photon excitation of channelrhodopsin-2.
science: computational and mathematical mod- Nat Methods 7(10):848–854
eling of neural systems, Computational 15. Rickgauer JP, Deisseroth K, Tank DW (2014)
Neuroscience Series. MIT Press Simultaneous cellular-resolution optical per-
3. Jun JJ et al (2017) Fully integrated silicon turbation and imaging of place cell firing fields.
probes for high-density recording of neural Nat Neurosci 17(12):1816–1824
activity. Nature 551:232 16. Prakash R et al (2012) Two-photon optoge-
4. Sofroniew NJ et al (2016) A large field of view netic toolbox for fast inhibition, excitation and
two-photon mesoscope with subcellular reso- bistable modulation. Nat Methods 9(12):
lution for in vivo imaging. elife 5:e14472 1171–1179
5. Shusterman R et al (2011) Precise olfactory 17. Shemesh OA et al (2017) Temporally precise
responses tile the sniff cycle. Nat Neurosci single-cell-resolution optogenetics. Nat Neu-
14(8):1039–1044 rosci 20(12):1796–1806
6. Fu Y et al (2014) A cortical circuit for gain 18. Mardinly AR et al (2018) Precise multimodal
control by behavioral state. Cell 156(6): optical control of neural ensemble activity. Nat
1139–1152 Neurosci 21(6):881–893
7. Schneider DM, Nelson A, Mooney R (2014) A 19. Chen IW et al (2019) In vivo submillisecond
synaptic and circuit basis for corollary discharge two-photon optogenetics with temporally
in the auditory cortex. Nature 513(7517): focused patterned light. J Neurosci 39(18):
189–194 3484–3497
8. McGinley MJ, David SV, McCormick DA 20. Packer AM et al (2015) Simultaneous
(2015) Cortical membrane potential signature all-optical manipulation and recording of neu-
of optimal states for sensory signal detection. ral circuit activity with cellular resolution
Neuron 87(1):179–192 in vivo. Nat Methods 12(2):140–146
9. Hires SA et al (2015) Low-noise encoding of 21. Yang W et al (2018) Simultaneous two-photon
active touch by layer 4 in the somatosensory imaging and two-photon optogenetics of cor-
cortex. elife 4:e06619 tical circuits in three dimensions. elife 7:
10. Cury KM, Uchida N (2010) Robust odor cod- e32671
ing via inhalation-coupled transient activity in 22. Chettih SN, Harvey CD (2019) Single-neuron
the mammalian olfactory bulb. Neuron 68(3): perturbations reveal feature-specific competi-
570–585 tion in V1. Nature 567(7748):334–340
11. Salzman CD, Britten KH, Newsome WT 23. Carrillo-Reid L et al (2019) Controlling visu-
(1990) Cortical microstimulation influences ally guided behavior by holographic recalling of
perceptual judgements of motion direction. cortical ensembles. Cell 178(2):447–457.e5
Nature 346(6280):174–177 24. Marshel JH et al (2019) Cortical layer–specific
12. Rickgauer JP, Tank DW (2009) Two-photon critical dynamics triggering perception. Science
excitation of channelrhodopsin-2 at saturation. 365:eaaw5202
Proc Natl Acad Sci U S A 106(35): 25. Gill JV et al (2020) Precise holographic manip-
15025–15030 ulation of olfactory circuits reveals coding fea-
13. Andrasfalvy BK et al (2010) Two-photon sin- tures determining perceptual detection.
gle-cell optogenetic control of neuronal Neuron 108(2):382–393.e5
Illuminating Neural Computation Using Precision Optogenetics 391
26. Dalgleish HW et al (2020) How many neurons 43. Yang Y et al (2008) Millisecond-scale differ-
are sufficient for perception of cortical activity? ences in neural activity in auditory cortex can
elife 9:e58889 drive decisions. Nat Neurosci 11(11):
27. Robinson NTM et al (2020) Targeted activa- 1262–1263
tion of hippocampal place cells drives memory- 44. O’Connor DH et al (2013) Neural coding
guided spatial behavior. Cell 183(6): during active somatosensation revealed using
1586–1599.e10 illusory touch. Nat Neurosci 16(7):958–965
28. Daie K, Svoboda K, Druckmann S (2021) Tar- 45. Smear M et al (2013) Multiple perceptible sig-
geted photostimulation uncovers circuit motifs nals from a single olfactory glomerulus. Nat
supporting short-term memory. Nat Neurosci Neurosci 16(11):1687–1691
24(2):259–265 46. Stanford TR et al (2010) Perceptual decision
29. Dal Maschio M et al (2017) Linking neurons making in less than 30 milliseconds. Nat Neu-
to network function and behavior by rosci 13(3):379–385
two-photon holographic optogenetics and vol- 47. Thorpe SJ, Fabre-Thorpe M (2001) Seeking
umetric imaging. Neuron 94(4):774–789.e5 categories in the brain. Science 291:260–263
30. Chong E et al (2020) Manipulating synthetic 48. Luna R et al (2005) Neural codes for percep-
optogenetic odors reveals the coding logic of tual discrimination in primary somatosensory
olfactory perception. Science 368(6497):1329 cortex. Nat Neurosci 8(9):1210–1219
31. Lerman GM et al (2019) Real-time in situ 49. Resulaj A, Rinberg D (2015) Novel behavioral
holographic optogenetics confocally unraveled paradigm reveals lower temporal limits on
sculpting microscopy. Laser Photonics Rev mouse olfactory decisions. J Neurosci 35(33):
13(9):1900144 11667–11673
32. Klapoetke NC et al (2014) Independent opti- 50. Wesson DW et al (2008) Rapid encoding and
cal excitation of distinct neural populations. perception of novel odors in the rat. PLoS Biol
Nat Methods 11(3):338–346 6(4):e82
33. Resulaj A et al (2018) First spikes in visual 51. Lerman GM et al (2018) Precise optical prob-
cortex enable perceptual discrimination. elife ing of perceptual detection. bioRxiv 456764
7:e34044 52. Smear M et al (2011) Perception of sniff phase
34. Wilson CD et al (2017) A primacy code for in mouse olfaction. Nature 479(7373):
odor identity. Nat Commun 8(1):1477 397–400
35. Packer AM, Roska B, Hausser M (2013) Tar- 53. Busse L et al (2011) The detection of visual
geting neurons and photons for optogenetics. contrast in the behaving mouse. J Neurosci
Nat Neurosci 16(7):805–815 31(31):11351
36. Haddad R et al (2013) Olfactory cortical neu- 54. Knutsen PM, Pietr M, Ahissar E (2006) Haptic
rons read out a relative time code in the olfac- object localization in the vibrissal system:
tory bulb. Nat Neurosci 16:949 behavior and performance. J Neurosci 26(33):
37. Dewan A et al (2018) Single olfactory recep- 8451
tors set odor detection thresholds. Nat Com- 55. Pologruto TA, Sabatini BL, Svoboda K (2003)
mun 9(1):2887 ScanImage: flexible software for operating laser
38. Hecht S, Shlaer S, Pirenne MH (1942) Energy, scanning microscopes. Biomed Eng Online 2:
quanta, and vision. J Gen Physiol 25(6): 13
819–840 56. Pnevmatikakis EA, Giovannucci A (2017)
39. Houweling AR, Brecht M (2008) Behavioural NoRMCorre: an online algorithm for piece-
report of single neuron stimulation in somato- wise rigid motion correction of calcium imag-
sensory cortex. Nature 451(7174):65–68 ing data. J Neurosci Methods 291:83–94
40. Doron G et al (2014) Spiking irregularity and 57. Smeds L et al (2019) Paradoxical rules of spike
frequency modulate the behavioral report of train decoding revealed at the sensitivity limit
single-neuron stimulation. Neuron 81(3): of vision. Neuron 104(3):576–587.e11
653–663 58. Picot A et al (2018) Temperature rise under
41. Huber D et al (2008) Sparse optical microsti- two-photon optogenetic brain stimulation.
mulation in barrel cortex drives learned beha- Cell Rep 24(5):1243–1253.e5
viour in freely moving mice. Nature 59. Panzeri S et al (2017) Cracking the neural code
451(7174):61–64 for sensory perception by combining statistics,
42. Histed MH, Maunsell JH (2014) Cortical neu- intervention, and behavior. Neuron 93(3):
ral populations can guide behavior by integrat- 491–507
ing inputs linearly, independent of synchrony.
Proc Natl Acad Sci U S A 111(1):E178–E187
392 Jonathan V. Gill et al.
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International
License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing, adaptation, distribution
and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative Commons license,
unless indicated otherwise in a credit line to the material. If material is not included in the chapter’s Creative
Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use,
you will need to obtain permission directly from the copyright holder.
Chapter 13
Abstract
Stimulated Raman Scattering (SRS) microscopy is a light-based non-linear imaging method for visualizing a
molecule based on its chemical properties, i.e., the vibrational energy states reflecting the molecule’s
structure and its environment. This technique, relying on the specificity of the molecule’s spectral finger-
print, enables label-free, high-sensitivity, and high-resolution 3D reconstruction of the distribution and the
properties of a molecule within a tissue. Despite its tremendous potentials, the application of SRS is still not
frequent in the field of life science, where it could be applied over an extremely broad investigation range,
from the study of the molecular interactions at subcellular level to the characterization of tissue alterations
in clinical studies. Trying to fill this gap, here, after describing the general principles of SRS, we present the
materials and the methods to integrate spectrally focused Stimulated Raman Spectroscopy (sf-SRS) on
commercial multiphoton microscopes and highlight the critical aspects to consider.
1 Introduction
Eirini Papagiakoumou (ed.), All-Optical Methods to Study Neuronal Function, Neuromethods, vol. 191,
https://doi.org/10.1007/978-1-0716-2764-8_13, © The Author(s) 2023
393
394 Tamás Váczi et al.
A B
Energy diagram Input Field Intensity Output Field Intensity
Virtual state
Stokes Pump Stokes Pump
Spontaneous
Stokes beam
Pump beam
ΔIs SRG
Raman
ΔIp SRL
ω
Δω Δω
Vibraonal state
Δω
Ground state ωs ωp
ω
ωs ωp
ω
C D E F
Pump ωp
Frequency
δω Stokes ωs
ω0 δω
Time
Δω
Fig. 1 (a) The SRS Jablonsky diagrams showing the energy states and the transitions associated with
Spontaneous Raman and Stimulated Raman signals with the y-axis indicating the energetic level as
represented in terms of the radiation frequency. (b) Changes of the laser intensities during Stimulated
Raman Spectroscopy in the case of Stimulated Raman Gain (SRG) and Stimulated Raman Loss (SRL). (c)
Frequency-time properties of ultrashort Stokes and pump pulses with chirping parameter β ¼ 0 and the
corresponding spectral resolution below. (d) Dispersion of the spectral component in chirped pulse with β 6¼ 0
with matched chirping condition and the corresponding spectral resolution below. (e) Dispersion of the
spectral component in chirped pulse with β > 0 with un-matched chirping condition and the corresponding
spectral resolution below. (f) Dispersion of the spectral component in chirped pulse with β > 0 with matched
chirping condition with different temporal shift and the corresponding spectral window. Time axis corresponds
to the time of arrival of the spectral components at the same point in the space
1.3 Spectrally Implementing a system for SRS with spectral focusing capabilities
Focused SRS: The within a microscope for 3D beam scanning typically requires the
General Hardware integration of a set of add-on systems: (1) the optical pathway for
Layout the routing, the temporal control, the high-frequency modulation,
and the conditioning of the pump and the Stokes beams; (2) the
implementation of a dedicated signal detection apparatus; (3) the
arrangement of a signal processing chain based on a lock-in ampli-
fier to return the SRS signal intensity corresponding to the illumi-
nated spot (“pixel”). While these aspects will be described in detail
in the Methods session, here we present the basic principles guiding
the sf-SRS implementation (Fig. 2). SRS relies on a pump-probe
scheme based on the spatial and temporal superposition of the
pulses of two beams characterized by different frequencies, i.e.,
ωp and ωS. As for the spatial overlap, this requires the use of ultrafast
routing mirrors, to ensure the XY (directions transversal with
respect to the light propagation direction) collinear overlap of the
two beams at the sample, and the adoption of beam expanders
(BE), to properly match the diameters and the degree of diver-
gence/convergence of the beams in order to have a good overlap of
the focal spots along the Z direction (longitudinal direction with
respect to the light propagation direction). Along with spatial
overlap of the beams, SRS requires in general a precise control of
the temporal overlap between the pulses of the two beams. It is
then fundamental that the pulses for pump and for Stokes beams
are emitted with a fixed phase delay one from the other. This relies
398 Tamás Váczi et al.
time
Pump beam
time Stokes beam (probe)
BE BDC Stokes beam (balanced detecon)
XY scanner
DM
Delay line
Scan lens
Tube lens
BE AOM BE
time
Main dichroic
PMT
Sample
I I
Microscope
DAQ
1.4 NA Condenser
GDD PD Filter
time
PBS
PD
time
Lock-in Differenal
Stokes beam Pump beam
amplifier amplifier
Fig. 2 The hardware layout for the integration of a sf-SRS in a laser scanning microscope. The main
components include: the time-locked pulsed laser sources (pump and Stokes beam), a Group Delay Dispersion
(GDD) system, a laser Intensity control (I) unit, a Pulse Stretching Apparatus (PSA), a set of Beam Expanders
(BE), an Acousto-Optic Modulator (AOM), a delay line, a Balanced Detection Component (BDC), a Dichroic
Mirror (DM), a Photomultiplier Tube (PMT), a Polarization-sensitive Beam Splitter (PBS), a pair of Photodiodes
(PD), a differential amplifier and a Lock-in amplifier. On the upper part, the temporal features of the Stokes
beam pulse and of the pump beams, presenting an orthogonal polarization and a temporal delay in the pulse
replica with respect to the Stokes pulse interacting with pump beam, are presented
on the use of two light sources or a single source with dual output
channels, whose pulse emissions are synchronized. Having syn-
chronized pulses does not imply necessarily that the pulses are
overlapping at the sample and, in general, it is not granted that
the amount of delay between the pulses remains the same when
varying one of the emission frequencies. Considering these aspects
requires a careful design of the optical paths (Fig. 2): first optimiz-
ing the difference in the distance travelled by the two beams in
order to minimize the range of possible delays; second, integrating
a delay line in one of the beam paths for finely tuning the residual
delay excursion range and getting the pulses overlapping at all the
selected frequencies. Along each of the two beam lines, it is practi-
cal to have a system for attenuating the laser intensity (I) and keep
the power for the two beams under control. As discussed in the
previous section, in the case of the sf-SRS the proper tuning of the
Stimulated Raman Spectroscopy to Unravel Molecular Mechanism in Living Organisms 399
2 Materials
3 Methods
3.1 Main Figure 3 shows the schematic of an SRS microscope with single-
Components for channel lock-in detection. It consists of the tunable wavelength
Integrating sf-SRS in a pump (TUN) and fixed wavelength Stokes (FIX) laser beams that,
Multiphoton in case of femtosecond laser sources, are guided along their light
Microscope paths with ultrafast-enhanced mirrors. These optical elements are
designed so that they introduce minimal distortions to the wave-
fronts of the ultrashort laser pulses during their reflection. Mirrors
are used to realize the criteria of SRS excitation described above,
namely to create the extra light path for one of the beams for the
coarse compensation of the relative delay of the pulses, to couple
the beam(s) in and out of the delay line, acousto-optical modulator
Fig. 3 Schematic layout of an SRS microscope system. FIX – fixed wavelength Stokes light source, TUN – tune-
able wavelength pump light source, M – mirrors, PD – photodiode, AUX in – signal input of the microscope for
imaging, AOM – acousto-optical modulator
Stimulated Raman Spectroscopy to Unravel Molecular Mechanism in Living Organisms 403
to operate far from the 1/frequency noise region of the source and
evaluate the SRS signal associated with the pixel over a sufficient
number of modulation cycles.
An SRS measurement starts with the adjustment of the wave-
length of the tunable light source, followed by setting the
corresponding delay line position and turning on the intensity
modulation. Then the intensities of the pump and Stokes beams
have to be adjusted. As a rule of thumb, the intensity of the
modulated beam has to be twice the other, and, in general, the
higher the laser intensity, the stronger the SRS effect. However, in
practice the laser-induced damage of the sample and the appearance
of different artefacts (Kerr effect, cross-phase modulation, etc.)
limit the useful laser intensity levels.
3.2 Methods for The net GDD (chirp) of the pulses comes from three main sources:
Spectral Tuning of the the chirp introduced within the laser, the optical setup including
System and Its the SRS system and the microscope, and the additional element
Optimization: Pulse (s) used to achieve pulse stretching. The contribution of these is
Chirp and Delay difficult to calculate or estimate precisely, therefore, the experimen-
Control tal measurement of the spectral and temporal widths of the laser
beams passing through the SRS system, and the microscope has to
be performed with a spectrometer and an autocorrelator, respec-
tively. The results can be used to determine the amount of chirp to
be introduced into the pump and Stokes beams to achieve the chirp
matching condition.
While the chirp of the emitted pulses cannot be adjusted
directly for some femtosecond lasers, other systems include the
option of compensating the group delay dispersion introduced by
the optical elements. This is a convenient tool to fine-tune the net
GDD and to establish precise chirp matching in spectrally focused
SRS measurements under a variety of laser wavelengths, i.e., Raman
sampling ranges.
The chirp related to the optical components of the SRS micro-
scope is an intrinsic property of the system that has to be taken into
account when adjusting the chirp matching conditions. The main
source is represented by transmissive optical elements—divergence
correctors, filters, beam splitters, objective lens, etc.
The third, and most substantial, component is the pulse
stretching device. In practice, the temporal redistribution of
the spectral components within the pulse can be achieved by
using combinations of gratings, prisms, or multilayered mirrors.
Another approach is to use a dispersive medium, in which the
components of the laser pulse with different wavelengths will travel
with slightly different velocity. The velocity difference per unit
length of the dispersive medium is given as the Group Velocity
Dispersion (GVD), and GDD ¼ GVD∙length of the medium. As
a consequence, a chirp will be introduced into the pulse, the extent
Stimulated Raman Spectroscopy to Unravel Molecular Mechanism in Living Organisms 405
Table 1
Group velocity dispersion of the Ohara S-TIH53 high density flint glass at different wavelengths [20]
Fig. 4 Chirp matching as a function of glass length (green line; Lp: glass length in pump beam, with additional
300 mm glass in Stokes beam). At the value of βP/βS ¼ 1, the difference between the instantaneous
frequencies of the temporally overlapping pulses is constant. The violet line shows the calculated best
spectral resolution without distortion effects
horizontally oriented mirror was used to reflect the beam above the
height of the first pass. At the end of the second pass another, now
vertically oriented hollow roof mirror is used to reflect it in the
height of the second pass, after which the first mirror reflects it onto
the knife edge mirror in the height of the first pass. So, the beam is
passing the 200 mm long glass four times. The tunable beam passes
the 200 mm glass rod only twice. Here a shorter rod of 75 mm
length is also used to achieve the required 550 mm
(2 200 mm + 2 75 mm) of glass length. The distances between
the knife edge and the hollow roof mirrors are adjusted to approxi-
mately maintain the temporal synchronization of the Stokes and
pump pulses.
The knife edge mirrors couple in and out the laser beams to the
spectral focusing unit. If needed, e.g., for femtosecond two-photon
measurements, these mirrors can be mounted on magnetic, detach-
able mounts so the system can be used in both femtosecond and
spectral focusing modes.
A spectral focusing SRS can be used to record vibrational
spectra of the samples. As it was detailed in the introduction, by
tuning the relative delay between the spectral focused pulses, they
will excite different vibrational transitions without tuning the wave-
length of the laser. By recording a calibration Raman spectrum on a
known, suitable sample, the delay line positions can be converted
into Raman shift values after appropriate curve fitting. Figure 6
demonstrates this capability and the high spectral resolution of a
sf-SRS microscope on succinic acid, a compound with intrinsically
narrow Raman lines. The resolution of the system was found to be
on par with the calculated resolution of 8 cm1 for this system
(cf. Figure 4, purple line minimum at 500–550 mm). Note that a
known chirp parameter at the chirp matching condition can also be
408 Tamás Váczi et al.
Fig. 6 Spontaneous Raman and sf-SRS spectra of succinic acid. The raw SRS spectrum (green) is shown with
respect to the delay line position on the upper X-axis and with respect to the corresponding Raman shift. With
a linear chirp, the stimulated Raman shift changes linearly with delay stage movement. The spontaneous
Raman spectrum (blue) is shown as reference
3.3 Methods for In the Differential Detection (DD), a reference signal is measured
Optimal Signal simultaneously along with the SRS one. This component is sub-
Detection: Differential tracted from the SRS signal in order to suppress the common noise
Detection of the laser light source. DD is an efficient solution to reduce the
noise and increase the sensitivity of the SRS measurements. The
main requirements of the high noise reduction are to minimize the
Stimulated Raman Spectroscopy to Unravel Molecular Mechanism in Living Organisms 409
Fig. 7 SRS spectrum recorded from the brain of a live zebrafish with the corresponding SRS images (large field
of view, upper line) and zoomed-in insets (lower line) at different vibrational levels. Scale bars are 50 (upper
line) and 10 μm (lower line)
delay between the signals coming from the reference and SRS arms
and to have similar signal levels at the two detectors.
For the SRS with DD, the reference signal can be generated in
several ways. In some configurations, the detected beam is split
before the main dichroic mirror (Fig. 2), and some part of it is
diverted into the reference detector, so the reference beam is differ-
ent from the SRS one, both spatially and temporally. As a conse-
quence of the latter, this solution requires the insertion of a delay—
either into the optical or the electronic path—that compensates for
the longer distance between the beam splitter and the SRS detector
under the stage of the microscope. The main problem with this
approach is that it cannot compensate for the fluctuations of the
signal level at the SRS detector caused by the varying optical density
of the sample during the scanning.
The other approach is to have the reference beam passing the
same optical path as the SRS one and to place the reference detector
also under the sample. This allows for the two components to have
the same intensity change upon passing through the sample of
varying optical density. The reference beam can be realized by
creating a delayed replica of the pulses in the pump or Stokes
beam (the one detected) that will not be synchronized temporarily
with the pulses of the other beam [25, 26]. Then, the only differ-
ence between the pulse and its replica is that the intensity of the
former is affected by the SRS interaction, so their differential
detection will result in efficient common mode noise suppression.
The two arms of the balance detection can be conveniently multi-
plexed and de-multiplexed taking advantage orthogonal linear
polarization states.
410 Tamás Váczi et al.
Fig. 8 Schematic layout of the differential detection. HWP – half-wave plate, M – mirror, PBS – polarization-
dependent beamsplitter, A-B detector – differential detector
4 Notes
x104
4.8
4.6
4.4
Intensity [a.u.]
4.2
Signal
4.0
3.8
3.6 Noise
3.4
3.2
0 5 10 15 20 25 30 35 40
x [μm]
Fig. 10 High quality sf-SRS image recorded on polystyrene microbeads and an intensity profile measured on
the top center microparticle
4.2 Strategies for SRS requires tight focusing of the pump and Stokes beams. This
Ensuring Optimal criterion is very strict; the pulses should be in the very same focal
Spatio-Temporal volume at the same time. Several factors could affect the tight
Overlap Between focusing of the beams (even in an optimally adjusted SRS system)
Pump and Stokes that could have temporary or long-term effect, including the tem-
Beams perature and air density fluctuations, pointing stability of the two
beams, deformation of the mirror surfaces due to heating by the
intense laser beams, etc. On tunable SRS systems, the wavelength
dependence of the depth of focus of many objective lens could be
another issue.
The above problems can be addressed in several ways. The local
temperature and air density fluctuations in the surrounding of the
SRS system can be minimized by using air conditioning with precise
temperature control, proper shielding (closed box) around the
optical setup, and proper cooling of the overheating parts.
The pointing stability of the two beams can be improved by
integrating an automatic beam stabilizer into the optical paths of
the SRS system. The beam stabilizer for a single beam consists of
two steering mirrors, each with high-speed and high-resolution
angular adjustment capabilities in two orthogonal directions (rea-
lized through stepper motors or piezo drives) and two laser beam
position sensors that can detect the lateral misalignment of the laser
beam with high precision (e.g., a camera or a quadrant detector).
The two detectors are placed at a large distance from each other
into an auxiliary arm obtained by placing a beam picker or a beam
splitter into the main beam.
Stimulated Raman Spectroscopy to Unravel Molecular Mechanism in Living Organisms 413
1.3
Change in Laser divergence
1.2
1.1
1.0
0.9
0.8
0.7
-1 -0.5 0 0.5 1
Distance from ideal posion of the beam expander [mm]
Fig. 11 Dependence of the divergence of a laser beam of 800 nm on the distance of the two lenses of a
divergence corrector telescope
4.3 Optimizing the It has been described that detecting the SRS signal relies on lock-in
Beam Modulation amplifiers, which work by driving one of the excitation beams with
Frequency and Depth a high-frequency modulation. This is typically achieved using an
acousto-optical modulator (AOM), a crystal with a piezo actuator
attached to it. It works like a dynamic optical grating, where the
periodic modulation of the refractive index is created by the acous-
tic waves introduced into the crystal via the piezo actuator. As a
result, the crystal will behave as diffractive element and will divert
the collinear incident beam to the diffraction angle. By switching
the acoustic waves on and off, the beam can be switched between
straight and diverted states, so the intensity of both straight and
diverted beams will be modulated. As for optical gratings, the
414 Tamás Váczi et al.
1.0
0.8
Modulaon depth
0.6
0.4
0.2
0
100 101 102 103 104
Frequency [kHz]
Fig. 12 Frequency dependence of poor (blue) and optimized (red) modulation depths
Acknowledgments
References
1. Sanderson MJ, Smith I, Parker I, Bootman MD in noise. J Phys E 8:621–627. https://doi.org/
(2014) Fluorescence microscopy. Cold Spring 10.1088/0022-3735/8/8/001
Harb Protoc 2014:pdb.top071795. https:// 14. Shi L, Fung AA, Zhou A (2021) Advances in
doi.org/10.1101/pdb.top071795 stimulated Raman scattering imaging for tis-
2. Strack R (2022) Imaging without the labels. sues and animals. Quant Imaging Med Surg
Nat Methods 19:30–30. https://doi.org/10. 11:1078101–1071101. https://doi.org/10.
1038/s41592-021-01376-0 21037/qims-20-712
3. Orlando A, Franceschini F, Muscas C et al 15. Lee HJ, Cheng J-X (2019) 5 – Label-free sti-
(2021) A comprehensive review on Raman mulated Raman scattering imaging of neuronal
spectroscopy applications. Chemosensors 9: membrane potential. In: Alfano RR, Shi L
2 6 2 . h t t p s : // d o i . o r g / 1 0 . 3 3 9 0 / (eds) Neurophotonics and biomedical spec-
chemosensors9090262 troscopy. Elsevier, pp 107–122
4. Hendra PJ, Stratton PM (1969) Laser-Raman 16. Lee HJ, Zhang D, Jiang Y et al (2017) Label-
spectroscopy. Chem Rev 69:325–344. https:// free vibrational spectroscopic imaging of neu-
doi.org/10.1021/cr60259a003 ronal membrane potential. J Phys Chem Lett 8:
5. John N, George S (2017) Chapter 5 – Raman 1932–1936. https://doi.org/10.1021/acs.
spectroscopy. In: Thomas S, Thomas R, jpclett.7b00575
Zachariah AK, Mishra RK (eds) Spectroscopic 17. Lombardini A, Mytskaniuk V, Sivankutty S et al
methods for nanomaterials characterization. (2018) High-resolution multimodal flexible
Elsevier, pp 95–127 coherent Raman endoscope. Light Sci Appl 7:
6. Rigneault H, Berto P (2018) Tutorial: coher- 10. https://doi.org/10.1038/s41377-018-
ent Raman light matter interaction processes. 0003-3
APL Photonics 3:091101. https://doi.org/ 18. Saar BG, Johnston RS, Freudiger CW et al
10.1063/1.5030335 (2011) Coherent Raman scanning fiber endos-
7. Min W, Freudiger CW, Lu S, Xie XS (2011) copy. Opt Lett 36:2396–2398. https://doi.
Coherent nonlinear optical imaging: beyond org/10.1364/OL.36.002396
fluorescence microscopy. Annu Rev Phys 19. Zhao Z, Shen Y, Hu F, Min W (2017) Applica-
Chem 62:507–530. https://doi.org/10. tions of vibrational tags in biological imaging
1146/annurev.physchem.012809.103512 by Raman microscopy. Analyst 142:
8. Li Y, Shen B, Li S et al (2021) Review of 4018–4029. https://doi.org/10.1039/
stimulated Raman scattering microscopy tech- c7an01001j
niques and applications in the biosciences. Adv 20. RefractiveIndex.INFO – Refractive index data-
Biol 5:2000184. https://doi.org/10.1002/ base. https://refractiveindex.info/. Accessed
adbi.202000184 10 Apr 2022
9. Cheng J-X, Xie XS (2013) Coherent Raman 21. Kearney SP (2014) Bandwidth optimization of
scattering microscopy. CRC Press, Taylor & femtosecond pure-rotational coherent anti-
Francis Group, Boca Raton stokes Raman scattering by pump/stokes spec-
10. Saar BG, Freudiger CW, Reichman J et al tral focusing. Appl Optics 53:6579–6585.
(2010) Video-rate molecular imaging in vivo https://doi.org/10.1364/AO.53.006579
with stimulated Raman scattering. Science 330: 22. Cole RA, Slepkov AD (2018) Interplay of pulse
1368–1370. https://doi.org/10.1126/sci bandwidth and spectral resolution in spectral-
ence.1197236 focusing CARS microscopy. JOSA B 35:
11. Sirleto L, Ranjan R, Ferrara MA (2021) Analy- 842–850. https://doi.org/10.1364/JOSAB.
sis of pulses bandwidth and spectral resolution 35.000842
in femtosecond stimulated Raman scattering 23. Mohseni M, Polzer C, Hellerer T (2018) Res-
microscopy. Appl Sci 11:3903. https://doi. olution of spectral focusing in coherent Raman
org/10.3390/app11093903 imaging. Opt Express 26:10230–10241.
12. Dunlap B, Richter P, McCamant DW (2014) https://doi.org/10.1364/OE.26.010230
Stimulated Raman spectroscopy using chirped 24. Label-free DNA imaging in vivo with stimu-
pulses. J Raman Spectrosc 45:918–929. lated Raman scattering microscopy | PNAS.
https://doi.org/10.1002/jrs.4578 https://www.pnas.org/doi/full/10.1073/
13. Blair DP, Sydenham PH (1975) Phase sensitive pnas.1515121112. Accessed 31 Mar 2022
detection as a means to recover signals buried
416 Tamás Váczi et al.
25. Crisafi F, Kumar V, Scopigno T et al (2017) 26. Polli D, Cerullo G (2022) Chapter 5 – bal-
In-line balanced detection stimulated Raman anced detection SRS microscopy. In: Cheng
scattering microscopy. Sci Rep 7:10745. J-X, Min W, Ozeki Y, Polli D (eds) Stimulated
https://doi.org/10.1038/s41598-017- Raman scattering microscopy. Elsevier, pp
09839-1 81–90
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0 International
License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing, adaptation, distribution
and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the
source, provide a link to the Creative Commons license and indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative Commons license,
unless indicated otherwise in a credit line to the material. If material is not included in the chapter’s Creative
Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use,
you will need to obtain permission directly from the copyright holder.
INDEX
A H
Adaptive optics .....................................18, 193, 237, 252, Holographic microscope .............................................. 105
254, 263, 326, 356
All-optical neurophysiology..........................2–19, 21, 32, I
37, 56, 58 Image processing................................... 66, 208–209, 296
Imaging at depth.................................................. 263–282
B
Bessel beam ............................................ 61, 63, 236–238, L
240–242, 244–246, 252, 301, 311
Label-free imaging ........................................................ 400
Brain machine interfacing............................................. 142 Lens system design............................................... 188, 189
Light shaping .............................28, 29, 31, 89, 275, 276
C
Light-sheet microscopy .................................50, 231–258
Calcium imaging ....................................9, 11, 17, 18, 57,
104, 140–142, 144–146, 156, 170, 248, 249, M
284, 308, 309, 334, 336–339, 348, 378
Molecular tools ....................................... 3, 4, 11, 96, 105
Channelrhodopsin......... 2, 4–8, 10–15, 37, 59, 173, 181 Multiphoton microscopy .................................... 196, 239,
ChroME ............... 13, 14, 104, 123, 125–127, 129, 130 264, 265, 295–297
Compressive sensing ........................................... 250, 256,
257, 265, 272–275, 279, 286 N
Computer-generated holography (CGH) .................... 17,
21, 61, 90, 102, 112, 166, 232, 238, 242, 250, Neural imaging..................................................... 293–326
254, 339, 340 Neuronal ensembles................... 331, 332, 337, 347–353
Control of behavior ............................................. 331, 348 Neurophysiology ............................................ 2–19, 49–68
Cross-activation...................................168, 172, 179–181
O
E Olfactory system.......................................... 366, 369, 372
Event-related analysis of functional fluorescence One-photon excitation ........................................... 50, 59,
traces ................................................ 149, 150, 159 61, 141, 144, 148, 322, 333, 335
Opsin ....................................... 3, 5–8, 11–13, 20, 58, 59,
F 64, 78, 101, 104, 105, 112, 122–131, 143, 149,
156, 167, 172–176, 181, 343, 355, 376, 380, 387
Fiber optics .................................................................... 194 Optical microscopy .............................................. 293, 393
Fluorescence microscopy .............................193, 294–325 Optogenetics ................................. 7, 11, 27, 29, 38, 102,
Functional imaging ............................................. 8, 52, 66,
106, 114, 115, 123, 127, 138, 160, 232, 238,
138, 139, 142, 156, 283, 284, 313, 315, 323–325 247, 265, 288, 332–347, 356, 363–414
G P
Generalized phase contrast (GPC)..................... 3, 17, 19, Pattern completion .............................332, 348, 352–354
21–26, 91
Photobiophysics ..................................167, 168, 179–181
Genetically encoded calcium indicators (GECIs) ........ 75, Population vectors ............................................... 348, 350
76, 139 Psychophysics .............................................. 364, 366, 368
genetically encoded voltage indicators (GEVIs) .......... 10,
59, 60, 75
Eirini Papagiakoumou (ed.), All-Optical Methods to Study Neuronal Function, Neuromethods, vol. 191,
https://doi.org/10.1007/978-1-0716-2764-8, © The Editor(s) (if applicable) and The Author(s) 2023
417
ALL-OPTICAL METHODS TO STUDY NEURONAL FUNCTION
418 Index
S Two-photon imaging......................................14, 63, 102,
118, 122, 146, 172, 179, 197, 198, 208, 337,
Single-pixel imaging................................... 208, 236, 265, 343, 347
271, 272, 275–277, 282 Two-photon optogenetics .................................. 123, 129,
Soma-targeted .....................................7, 13, 14, 124, 177 138, 331, 337, 348
Spectral focusing ......................................... 397, 406, 407
Spectral independency ......................................... 141, 160 U
Stimulated Raman spectroscopy.......................... 394–396
Stimulation artefact avoidance .................................11, 58 Ultrafast pulse propagation ......................................53, 61
Structured illumination microscopy................... 294, 312,
V
314–316, 320
Synapse ........................................................ 294, 306, 313 Voltage imaging ..................................10–12, 16, 57, 325
System integration ..............................138, 140, 149, 158 Volumetric neuronal imaging................................. 16, 63,
212, 236, 295, 326, 336
T
W
Temporal focusing ...........................3, 16, 17, 19, 26–30,
36, 87, 91, 103, 105, 107, 264, 275, 285, 380 Widefield imaging ............................................... 217, 263,
3D holography .............................................................. 111 271, 275, 282, 284, 324
3D photostimulation ................................................88–92 Widefield microscopy...................................192, 293–326
3D-SHOT ........................................29, 31, 90, 102–108,
110–118, 122–124, 126–128, 131 Z
Two-photon excitation fluorescence
Zebrafish ...............................................51, 138, 232, 238,
microscopy ..................................3, 15, 20, 22, 29,
247–250, 254, 307, 408, 409
110, 141, 146, 156, 168, 181, 212, 239, 300,
333, 334, 343