RESEARCH
Peer reviewed scientific video journal
Video encyclopedia of advanced research methods
Visualizing science through experiment videos
EDUCATION
Video textbooks for undergraduate courses
Visual demonstrations of key scientific experiments
BUSINESS
Video textbooks for business education
OTHERS
Interactive video based quizzes for formative assessments
Products
RESEARCH
JoVE Journal
Peer reviewed scientific video journal
JoVE Encyclopedia of Experiments
Video encyclopedia of advanced research methods
EDUCATION
JoVE Core
Video textbooks for undergraduates
JoVE Science Education
Visual demonstrations of key scientific experiments
JoVE Lab Manual
Videos of experiments for undergraduate lab courses
BUSINESS
JoVE Business
Video textbooks for business education
Solutions
Language
English
Menu
Menu
Menu
Menu
A subscription to JoVE is required to view this content. Sign in or start your free trial.
Research Article
Haoyu Xu*1, Wei-Wen Chen*2, Jessica Z. Dixon2, Davendra S. Maharaj3, Karthikeya M. Sharma3, Xavier Audier2, Marcus T. Cicerone2
1Department of Biomedical Engineering,Georgia Institute of Technology, 2School of Chemistry and Biochemistry,Georgia Institute of Technology, 3School of Electrical and Computer Engineering,Georgia Institute of Technology
Erratum Notice
Important: There has been an erratum issued for this article. View Erratum Notice
Retraction Notice
The article Assisted Selection of Biomarkers by Linear Discriminant Analysis Effect Size (LEfSe) in Microbiome Data (10.3791/61715) has been retracted by the journal upon the authors' request due to a conflict regarding the data and methodology. View Retraction Notice
This article demonstrates amultimodal imaging platform combining broadband coherent anti-Stokes Raman scattering (BCARS), two-photon excitation fluorescence (TPEF), and two-photon fluorescence lifetime imaging microscopy (2p-FLIM), enabling ultra-high information-content chemical bioimaging.
Raman fingerprint spectroscopy and fluorescence lifetime imaging are emerging tools for studying metabolic profiles of biological specimens. While Raman fingerprint spectroscopy detects intrinsic molecular vibrations that reflect the molecular composition and chemical environment of a sample, fluorescence lifetime imaging measures changes in the excited-state lifetime of fluorophores that are sensitive to their microenvironments. Here, we present a multimodal imaging platform combining broadband coherent anti-Stokes Raman scattering (BCARS) and two-photon fluorescence lifetime imaging (2p-FLIM) microscopy that can acquire biologically relevant Raman fingerprint spectra and fluorescence lifetime signals in vivo and simultaneously. The tremendous chemical information obtained from spatially co-registered BCARS and 2p-FLIM images allows us to characterize the subtle differences between sub-cellular compartments and verify the potential false-positive results generated by fluorescence imaging alone. This is demonstrated by directly comparing the BCARS, 2p-FLIM, and two-photon excitation fluorescence (TPEF) signals simultaneously obtained from the same dye-stained organelle in the live, intact C. elegans expressing a green fluorescent protein (GFP) marker. In this work, we introduce the BCARS/2p-FLIM/TPEF setup scheme, the image acquisition steps, data processing, and representative results showing that the cross-modality imaging method enables rigorous characterization and in vivo detection at sub-cellular resolution. This protocol provides a framework for simultaneous chemical and fluorescence lifetime imaging to improve the accuracy of biological interpretation in complex living systems.
One challenge in biological studies is to accurately detect biological building materials in their native states and within the intact architecture of cells and tissues. Raman spectroscopy shows unprecedented potential for detecting intrinsic metabolites and chemical species within biological specimens and has been recently recognized as an emerging tool for metabolic profiling of biological specimens1. By analyzing the peaks of various metabolic groups shown in Raman fingerprint spectra, increasing evidence shows that the fingerprint features can be used to detect the early stages of diseases such as cancer2,3,4, cardiac disorders5,6, and neurodegenerative diseases7,8 and is particularly useful for addressing fundamental biological problems9,10 including imaging intracellular structures without any labeling perturbation11, differentiating various cell states12,13, and cell type classification14,15,16. Most of the chemical information ( > 90%) encoded in the Raman spectra is in the fingerprint region (500 to 1800 cm-1), making this region by far the most useful for discriminating chemical compositions in biological samples17. However, the weak Raman signal in this frequency range makes the applications of Raman fingerprint spectroscopy to biological studies challenging, particularly for live, intact biological specimens1.
Coherent Raman imaging (CRI) approaches significantly shorten the acquisition time of Raman signals, but most of these acquire relatively narrow spectral segments and are optimized for the strong CH stretch signals between 2800 - 3100 cm-1. Narrowband CARS and stimulated Raman scattering (SRS) (Figure 1A,B) use two picosecond pulses, usually to acquire a narrow portion of the strong CH signal 18,19,20,21,22,23. Multiplex CRI approaches, such as multiplex CARS (Figure 1C), use a combination of a picosecond and ~100 fs pulses to obtain a moderate spectral range (usually ≤ 400 cm-1) usually in the CH region24,25,26,27,28,29,30. Acquiring a whole biological-relevant Raman spectrum (from 600 - 3200 cm-1) per pixel using these imaging approaches is usually prohibitive because most fingerprint signals are 10-100 times weaker than CH-stretch signals31. Unlike most CRI approaches, broadband coherent anti-Stokes Raman scattering (BCARS) is optimized for single-shot broadband acquisition that covers the fingerprint and CH-stretch regions31,32,33. BCARS uses intra-pulse interactions of a ~10 fs pulse to efficiently excite low-frequency (fingerprint) vibrations (Figure 1D), and uses the same mechanism as multiplex CARS, the combination of the fs pulse with a ps pulse to generate vibrational coherence in the CH stretch region (Figure 1C)32,33. Furthermore, the BCARS signals are intrinsically calibrated due to their interaction with the non-resonant background (NRB), resulting in a quantitative and instrument-invariant full Raman spectrum per laser pulse at each pixel31, with the capability of rapidly acquiring high-information-content data from cells, intact tissues, and live biological organisms32,34,35,36,37,38,39.
The acquired raw BCARS spectra contain a strong so-called "non-resonant background" (NRB). The NRB is the contribution from the electronic response to the laser pulses that is not chemically specific. The NRB coherently interferes with the vibrationally resonant signal, acting as a heterodyne amplifier32. The NRB also distorts the underlying Raman spectrum (Figure 2A), but the Raman spectrum can be retrieved as long as significant contiguous spectral sections are acquired. Currently, there is no post-processing method available to separate the resonant CARS signal from the non-resonant component using a narrowband CARS approach. Raman signal retrieval from the broadband CARS signal can be achieved by utilizing the time-domain Kramers-Kronig relation (or Hilbert transform)33,40 (Figure 2B), maximum entropy41,42, or machine-learning43,44,45,46,47,48 approaches. Note that the first two methods analytically retrieve the phase (the resonant Raman signal) from the raw BCARS spectra, the machine-learning approaches can be highly dependent on the instrumentation and the training models.
While BCARS microscopy allows for label-free characterization of metabolic profiling of biological specimens, the same supercontinuum laser (950 - 1200 nm for the BCARS setup presented in this work) used for BCARS signal generation can also excite a wide range of dyes, such as Rhodamine-based, Bodipy-based, Alexa Fluor-based, and cyanine-based dyes, and various fluorescent proteins, including GFP and mCherry49,50. Compared to single-photon excitation using visible continuous wave (CW) lasers, multiphoton fluorescence microscopy, such as two-photon fluorescence lifetime imaging microscopy (2p-FLIM) and two-photon excitation fluorescence microscopy (TPEF), using near infrared (NIR) light offers a deeper sample penetration depth, intrinsic 3D sectioning capability, high-contrast images, and good resolution comparable to state-of-the-art confocal microscopy, making it an attractive tool for biological research51,52. Fluorescence lifetime imaging microscopy (FLIM) using a time-correlated single photon counting (TCSPC) approach detects the fluorescence decay signals repeatedly excited by a time-modulated laser source53,54. The fluorescence spectrum of a given fluorophore may change with environmental factors such as solvent polarity, but these changes are usually subtle. On the other hand, fluorescence lifetimes often vary by factors of two or more with environmental changes such as pH values55, viscosity56, polarity57, and concentration of ionic species58,59, binding interactions, the presence of quenchers, or structural changes53,54. This unique property of fluorescence lifetime provides an additional signal contrast in fluorescence imaging and has successfully addressed the issues of autofluorescence interference in cell and tissue imaging60, unmixing multiple fluorophores using a spectrally resolved FLIM approach61, cancer diagnosis and treatment monitoring62, label-free detection of cellular heterogeneity63, and metabolic changes64,65.
In this work, we introduce a multimodal imaging platform combining BCARS, 2p-FLIM, and TPEF. To the best of our knowledge, this is the first time presenting an imaging platform allowing simultaneous BCARS and 2p-FLIM/TPEF imaging. We demonstrate the in vivo imaging capabilities of this instrument using a widely used genetically tractable model organism, Caenorhabditis elegans (C. elegans). We show that the BCARS spectra, the 2p-FLIM decays of Nile Red (a vital lipid dye), and the two-photon fluorescence GFP signals of a reported lipid marker (DHS-3::GFP) can be simultaneously acquired in vivo from Nile-Red-stained worms. A practical guide for the imaging approach presented here includes (i) the sample should be optically transparent and with a thickness typically less than 100 µm; (ii) the staining condition can be different depending on the dyes and samples. In the Nile-Red staining of C. elegans results shown here, the Nile Red concentration is at ~µM level, and staining time can range from a few hours to overnight66; (iii) a dark sample environment is required to avoid visible light leakage into the 2p-FLIM/TPEF detectors. These imaging conditions are identical to the imaging conditions of most nonlinear optical and fluorescence microscopy. In addition, our system demonstrated here can be further applied to detect other dyes and fluorescent reporters in addition to Nile Red and GFP by introducing an additional Ti:Sapphire fs laser that covers 700 -950 nm range for the two-photon fluorescence excitation. Thus, while the representative results described in this protocol are the in vivo imaging of dye-stained C. elegans expressing GFP, the ultra-high information-content chemical bioimaging platform has applications that are broad and far-reaching.
1. C . elegans sample preparation and Nile Red Vital staining
NOTE: C. elegans has been heavily used for multiple imaging techniques due to its whole-body transparency. The internal organs and tissues can be easily imaged in the intact animal without dissection. In addition, ethics approval is not required for research involving C. elegans, as they are not considered vertebrates and are not subject to the same ethical regulations as higher animals like rodents. This section describes the synchronization of C. elegans worms and the preparation of Nile Red-stained samples for fluorescence imaging. Synchronized one-day adult worms are obtained via a two-generation egg-laying protocol and stained using freshly prepared Nile Red-OP50 plates.
2. Probe beam generation
NOTE: This protocol describes the generation of the picosecond (ps) probe beam using a home-built optical parametric oscillator (OPO) followed by second harmonic generation (SHG) with a home-built signal feedback loop system to stabilize the output power of the probe beam.
3. Supercontinuum (SC) generation for BCARS
NOTE: The simplified scheme of the BCARS-based multimodal system is shown in Figure 4. This protocol describes the generation of the SC beam.
4. Setting up the fluorescence detection for 2p-FLIM and TPEF and introducing an additional Ti:Sapphire fs laser
5. Setting up the BCARS detection path
NOTE: This protocol describes setting up a descanned detection path for line-scanning hyperspectral imaging. The 1D galvo mirror scans the excitation beam along the y-axis with 2-4 ms dwell time per pixel for sufficient signal integration. After each line scan, the sample stage advances one step along the x-axis. Descanning using a synchronized secondary galvo mirror is essential to maintain consistent spatial registration of the spectral signal on the camera during fast beam scanning, eliminating signal crosstalk between adjacent pixels and preserving spatial-spectral fidelity39.
6. Signal optimization and spectral resolution calibration
7. Signal acquisition
Figure 5 shows the measured spectra of the narrowband probe and the broadband supercontinuum (SC) pulses used for BCARS excitation. The probe pulse, generated via second harmonic generation of a ~1530 nm signal from a home-built optical parametric oscillator (OPO, Figure 3), is centered at 764.2 nm with a full width at half maximum (FWHM) of 0.2 nm, providing the spectral resolution at ~6.8 cm-1 in the resulting Raman spectra. The SC pulse, produced by coupling a pre-chirped femtosecond 1030 nm laser pulse into a 10 cm PCF (see Table of Materials), spans from approximately 950 nm to 1200 nm (≥ 1800 cm-1), covering the whole biologically relevant Raman spectral region (400 to ≥ 3200 cm-1) from fingerprint (through intra-pulse 3-color excitation), silent region (the mix of the 3- and 2-color excitations), and the CH/OH stretch region (through inter-pulse 2-color excitation). Together, these beams enable effective coherent Raman excitation with high spectral specificity and broad vibrational coverage.
To achieve precise synchronization between scanning, descanning, and image acquisition, we implemented a custom control system using an FPGA. The FPGA generates analog waveforms that drive both the scanning galvo and the descanning galvo (Figure 7A). The FPGA receives a line clock to mark the beginning of each scan line and produces duration-adjustable pixel triggers synchronized with the camera exposure (2-4 ms per trigger, equal to pixel dwell time). These signals are distributed across all detection channels, including FLIM, TPEF, and BCARS, to ensure accurate temporal and spatial co-registration. To compensate for slight optical differences introduced by varying objective lenses, the descanning amplitude can be finely tuned via a 7-bit digital potentiometer. A timing diagram illustrating the relationship between galvo signals and trigger outputs is shown in Figure 7B.
Figure 8 shows example images of BCARS, 2p-FLIM, and TPEF images acquired simultaneously from a live, Nile-Red-stained adult worm expressing DHS-3::GFP, which has been accepted as a lipid droplet marker in C. elegans lipid research69. As a reference, we first acquired the BCARS and 2p-FLIM signals from the autofluorescent gut granules of adult wild-type worms without any staining. The autofluorescence gut granules show a low Raman lipid signal (2850 cm-1) and a short fluorescence lifetime ~1 ns (the black spectrum and FLIM decay curve shown in Figure 8E,F, respectively), where our 2p-FLIM results are consistent with recent FLIM findings70. To investigate whether those autofluorescent gut granules contained starch components that were possibly originated from digested bacteria consumed by the worms, we measured the decay curve of a second harmonic generation (SHG) signal. Starch is known to generate SHG signals with zero fluorescence lifetime and therefore can be an SHG reference to estimate the instrument response function (IRF)71. Using our 2p-FLIM system and a thin ZnSe crystal as an SHG reference, we recorded the SHG response. The SHG decay curve showed a very fast lifetime of ~0.12 ns (gray curve in Figure 8F), which is much shorter than the ~0.86 ns lifetime we observed in the autofluorescent granules. This comparison indicates that starch SHG is unlikely to be the main contributor to the measured lifetime of autofluorescent granules. Next, we compared the retrieved Raman spectra from several intestinal regions with DHS-3::GFP signal (ROI 1-4 shown in Figure 8A,C,E) in the Nile-Red stained adult worms expressing DHS-3::GFP. The spectra from ROI 1 and 3 show weak protein signatures but an obvious 2850 cm-1 lipid Raman peak and other lipid-related signatures including the vibration signals at 1740 cm-1 (C=O), 1640 cm-1 (C=C), 1450 cm-1(CH def.), 1080 cm-1 (C-C), and 870 cm-1(C-C). In contrast, although ROI 2 exhibits strong DHS-3::GFP and Nile-Red signals (Figure 8A,B), the absence of lipid signatures and the presence of strong protein peaks including 1003 cm-1 (Phe.), 1665 cm-1 (Amide I), and 2930 cm-1 (CH3) in the ROI2 indicates that this particle was falsely labeled with lipid marker DHS-3::GFP and Nile Red dye. We also found that the ROI 4 spatially contacts a larger lipid particle ROI 3 (Figure 8C) and contains phospholipid-related Raman signatures such as 760 cm−1 (ethanolamine) and 880 cm-1 (νasN+(CH3)3). The spectrum of ROI 4 shows strong lipid and protein peaks indicated in Figure 8E, suggesting that this particle is a yolk particle interacting with a lipid droplet (ROI 3). This can be further confirmed by the Nile Red 2p-FLIM signals that were obtained simultaneously with BCARS. Both ROI 1 and ROI 3 are triglyceride-rich lipid droplets that exhibit a relatively longer lifetime (~3.38 ns and ~3.21 ns, respectively), but ROI 4 shows a shorter lifetime at ~2.73 ns (Figure 8F), which is a typical lifetime value when Nile Red stains polar lipids, consistent with our previous 2p-FLIM finding66. While ROI 2 was stained by Nile Red, its lifetime is significantly shorter than that of lipid-binding Nile Red dye, suggesting that Nile Red falsely tagged ROI2, but this non-specific staining can be separated based on the distinct lifetime. Overall, these suggest that although ROI 1-4 were all tagged by a well-accepted lipid droplet marker DHS-3::GFP, two of them (ROI 2 and 4), which are not triglyceride-rich lipid droplets, were labeled inaccurately. We conclude that multimodal imaging combining BCARS, 2p-FLIM, and TPEF generates high information-content data and can provide rigorous results.

Figure 1: Jablonski diagrams of coherent Raman scattering. (A) Stimulated Raman Scattering (SRS), where Raman-active vibrational modes are coherently driven by two incident fields (pump and Stokes), and the resulting modulation is detected on the pump (SRL) or probe (SRG) beam. (B) Narrowband Coherent Anti-Stokes Raman Scattering (CARS), using narrowband pump and Stokes beams to excite a single Raman mode, with the anti-Stokes signal read out using a degenerate probe. (C) Two-color or multiplex CARS, where a narrowband probe is used in combination with a supercontinuum (SC) Stokes beam to access multiple Raman shifts. (D) Three-color CARS, where both the pump and Stokes pulses are derived from a broadband SC. Abbreviations: SRS - Stimulated Raman Scattering; CARS - Coherent Anti-Stokes Raman Scattering. Please click here to view a larger version of this figure.

Figure 2: Simulated BCARS, NRB, and retrieved Raman spectra. (A) Raw BCARS spectrum and the exact estimation of the NRB reference. (B) Retrieved Raman signal when the Kramers-Kronig (KK) transformation is applied correctly without errors, representing an ideal spectrum retrieval. Abbreviations: NRB: Non-resonant Background; KK: Kramers-Kronig. Please click here to view a larger version of this figure.

Figure 3: Dual output laser and the home-built OPO/SHG system for BCARS probe beam generation. (A) The femtosecond branch passes through a grating compressor, generating a broadband supercontinuum (SC) beam through photonic crystal fiber (PCF). The pump branch drives an optical parametric oscillator (OPO) with a focusing lens (L), two curved mirrors (CM), an end mirror (EM), an output coupler (OC), and a periodically poled lithium niobate (PPLN) crystal. The narrowband signal at ~1530 nm is frequency-doubled in a second PPLN crystal to produce a ~765 nm probe via second harmonic generation (SHG) and is combined with the SC beam using a dichroic mirror (DM). An additional Ti:Sapphire beam is collinearly combined with the probe beam through a polarizing beam splitter (PBS) before the DM. A photodiode (PD) monitors SHG output for cavity stability feedback by adjusting the OPO cavity length (the position of EM) in real-time. (B) Jablonski diagram of the OPO process showing parametric down-conversion from 1030 nm to produce a narrowband signal and idler photon pair. (C) Jablonski diagram of the SHG process for frequency doubling the signal beam to generate the probe pulse. (D) Simplified schematic of the full beam path, highlighting major components and nonlinear optical conversions used to generate the narrowband probe beam. Abbreviations: PPLN: Periodically Poled Lithium Niobate; CM: Curved Mirror; EM: End Mirror; OC: Output Coupler; PD: Photodiode; SPF: Shortpass Filter; PCF: Photonic Crystal Fiber; DM: Dichroic Mirror; L: Lens; TC: Temperature Controller; WW: Wedge Window; PBS: Polarizing Beam Splitter; OPO: Optical Parametric Oscillator; SHG: Second Harmonic Generation. Please click here to view a larger version of this figure.

Figure 4: The multi-modal imaging setup combining BCARS, 2p-FLIM, and TPEF. The supercontinuum (SC) is generated from a 10 cm photonic crystal fiber (PCF) by pumping a 1 W, 130 fs 1030 nm laser. The prechirp (grating compressor) and prism compressor are set to compensate for the dispersion caused by the PCF, aspherical lens (Asph. lens), glasses, and objective lens before the sample stage. The probe pulse is generated from a home-built OPO/SHG system (see the details in Figure 3). A mechanical delay line is set to adjust the length of the probe beam path to match that of SC. After collinearly combining SC and probe beams with a dichroic mirror (DM), the beams are guided into a 1D-galvo/reflective beam expansion system, resulting in a 1D laser scanning geometry with a beam size similar to the back aperture of the focusing objective lens. The forward BCARS signal is collected by another objective lens, through a 4f-descan system, short-pass filters that reject the SC and probe, and finally focused onto the slit of the spectrometer. The 2p-FLIM and TPEF signals are collected with a non-descan geometry in the epi-direction. The additional Ti:Sapphire laser (not shown in this scheme) is collinearly combined with the probe before the DM (see the details in Figure 3). Abbreviations: DM: Dichroic Mirror; Asph. Lens: Aspherical Lens; PCF: Photonic Crystal Fiber; OAP: Off-axis Parabolic mirror; SP filter: Shortpass filter. Please click here to view a larger version of this figure.

Figure 5: The spectra of SC and probe. Both probe and SC spectra are shown on the same plot for comparison. The probe spectrum is centered at 764.2 nm with a full width at half maximum (FWHM) of 0.2 nm. The supercontinuum (SC) spectrum spans from 950 nm to 1200 nm, covering ≥ 1800 cm-1. Please click here to view a larger version of this figure.

Figure 6: BCARS signal processing pipeline. (A) The pipeline of raw BCARS signal processing. (B) The user interface and the steps using customized CRIKit2 software68 to retrieve Raman signal by applying time-domain Kramers-Kronig (TDKK), phase error correction, scale error correction, and wavenumber correction. The detailed mathematical formula can be found in33. Please click here to view a larger version of this figure.

Figure 7: Descanning and synchronization architecture. (A) Schematic diagram of the optical and electronic pathways used for signal descanning and synchronization. Fluorescence emission is collected via dichroic mirrors (DM) and directed to photomultiplier tubes (PMTs) for time-correlated single-photon counting (TCSPC) and analog detection. The anti-Stokes signal for BCARS is descanned using a galvo mirror placed in a 4f relay system. The scanning and descanning galvo mirrors are driven by analog signals from a field-programmable gate array (FPGA), which also generates synchronized trigger signals for pixel clocking for camera, ADC (Analog-to-Digital Converter) and TCSPC. The data are collected by a computer. (B) Timing diagram showing the scanning and descanning galvo drive signals, along with the corresponding line and pixel clock outputs. The FPGA precisely controls both galvos to maintain the anti-Stokes beam within a fixed region on the camera during scanning. Pixel and line clocks are used to synchronize data acquisition across all detection channels. Abbreviations: FPGA: field programmable gate array; PMT: photomultiplier tube; TCSPC: time-correlated single-photon counting; PA: pre-amplifier; ADC: analog-to-digital converter. Please click here to view a larger version of this figure.

Figure 8: The representative results of simultaneous in vivo imaging of BCARS, 2p-FLIM, and TPEF obtained from a Nile-Red stained adult worm expressing DHS-3::GFP. The overlapping images (A) of BCARS (yellow) and GFP signals (cyan) and (B) 2p-FLIM Nile Red (magenta) and GFP (cyan) signals. The scale bar = 10 µm. (C)(D) The enlarged image showing the BCARS, 2p-FLIM, TPEF signal of two connecting particles. (E) Raman spectra retrieved from selected regions of interest (ROIs 1 - 4 shown in (A-D)) along with an autofluorescent reference region from adult wild-type worms without any Nile-Red staining. (F) The fluorescence decay curves of ROI 1 – 4, the autofluorescence reference (from autofluorescent gut granules in wild-type adult worms), and SHG decay curve (gray). The fluorescence lifetimes were exponentially fitted after removing the interference of the instrument response function (IRF). Please click here to view a larger version of this figure.
This work demonstrates the architecture of BCARS and 2p-FLIM/TPEF imaging platform in detail. While a dual-output 1030 nm laser is used in our current setup, most high-power fs 1030 nm fiber lasers can be used as a master laser source for building a BCARS microscope. By carefully choosing the length and period of OPO/SHG PPLN crystals, generating a few-ps probe pulse with a sub-ps OPO 1030 nm pump is achievable. Although a relatively high laser repetition rate (100 MHz) or a relatively lower pulse peak power is demonstrated here, the BCARS signal can still be effectively generated using our BCARS signal excitation scheme that combines intra- and inter-stimulation (3-color and 2-color signal generation). The in vivoC. elegans images shown in the representative results further confirmed the feasibility of performing bioimaging using home-built SC/probe BCARS laser sources. While we previously showed BCARS spectroscopic bioimaging can be achieved by the sample-stage-scanning method32,34, the same quality of the spectral and spatial resolution can also be achieved by using the 1D-galvo scanning and descan geometry demonstrated here. Finally, we demonstrate the expansion of BCARS imaging platform by integrating 2p-FLIM and TPEF modality. Although the BCARS system is relatively complex, the multimodal BCARS/2p-FLIM/TPEF imaging platform can provide pixel-coregistered high-quality images with a combination of unbiased chemical and spatial information; changes in the fluorescence lifetime responding to the microenvironment, and good spatial distribution of both intrinsic metabolites and fluorescence labels of interest. More importantly, this can be done within live, intact specimens, providing invaluable insights into biological research and can be broadly applied to various biomedical research topics.
The BCARS data processing pipeline using customized CRIKit2 software33,68 is described in detail (Figure 6). While these procedures can be done in CRIKit2 with a customized Python script, the denoise steps (the steps in the yellow and orange regions shown in Figure 6A) can be done in real-time through a carefully orchestrated combination of multithreaded CPU stages and sequential GPU processing. The CPU pipeline (yellow region shown in Figure 6A) is responsible for high-throughput data preparation and cleaning, operating on a stream of spatial pixels captured from the camera. It is implemented as a series of thread-isolated stages connected through queues, enabling parallel and efficient data flow. The multithreaded CPU stages are configurable based on the signal being processed. For instance, NRB signal processing bypasses stages like Anscombe transformation and quantization. This modular design allows the system to adapt to different acquisition modes with minimal changes to the pipeline structure. Once accumulated, the signal tensor is transferred to the GPU for more compute-intensive sequential processing. The GPU pipeline (orange region shown in Figure 6A) begins with a singular value decomposition (SVD) to identify dominant spectral components. First, all the SVD components will be saved for the case of user-defined postprocessing and reconstruction. After that, a randomized SVD is performed multiple times, allowing the system to determine the number of significant singular values to be retained statistically. The signal is then reconstructed using only the selected components, enabling denoising and dimensionality reduction, followed by inverse Anscombe transformation. This on-the-fly processing approach has been implemented in our acquisition software, greatly reducing the overall BCARS data processing time. We expect that all data processing will be automated in the future by applying parallel calculation for the remaining steps, including TDKK, phase error correction, and scale error correction.
The multimodal system described here can be further expanded by adding additional FLIM and TPEF detectors that enable many-channel detection through implementing a wavelength dispersion element to achieve spectral-resolved FLIM and TPEF imaging. While the 2p-FLIM signal shown in this work used the BCARS SC ranging from 950 nm to 1200 nm as the excitation laser source, an additional Ti:Sapphire fs laser can be guided into the BCARS system. As shown in Figure 3, a tunable Ti:Sapphire fs laser, which has been implemented in our multimodal imaging platform (the dashed cyan line), was collinearly combined with the probe pulse, significantly broadening our two-photon fluorescence imaging capability for a wide range of fluorophores covering from blue to deep red fluorescence emission that have been routinely used in biological studies. However, the additional Ti:Sapphire laser has to be carefully adjusted and calibrated to ensure the overlap between BCARS and its point spread functions. This can be done by using dyed polystyrene beads for fine-tuning the axial and lateral overlapping. Other potential issues with introducing a third laser could be laser damage and interference between the fluorescence excited by SC and Ti:Sapphire pulses if the fluorophore exhibits a wide range of two-photon absorption wavelengths. The former issue can be addressed by experimental determination of the maximum input laser power. The latter can be mitigated by choosing a different laser repetition rate and polarization direction. In our case, we introduced an fs laser with a repetition rate of 76 MHz, which is different from our BCARS laser's 100 MHz repetition. The polarizing beam splitter (The PBS shown in Figure 3A) was used to combine the Ti:Sapphire fs laser and BCARS probe beam, allowing an orthogonal polarization between BCARS SC/probe pulses and the 2p-FLIM excitation pulse to minimize the possible interference between them.
Both BCARS and 2p-FLIM results are system-invariant. The BCARS signal is intrinsically calibrated with an intrinsic reference NRB signal, which has a fixed phase and amplitude relationship to the BCARS signal33. Consequently, BCARS yields Raman spectra with absolute amplitudes that do not vary day-to-day, or even between instruments. The lifetime data obtained from 2p-FLIM are also system-invariant. The results of FLIM phasor analysis are solely dependent on the intrinsic fluorescence decay property and the microenvironment of the fluorophore because its fluorescence intensity is normalized during the phasor component conversion72,73. These unique properties enable BCARS and 2p-FLIM to be attractive tools for quantitative comparison among samples and across instruments.
In the representative results, the BCARS spectra and fluorescence decay curves of several particles labeled with DHS-3::GFP are shown. Consistent with our previous finding34, we confirmed that in the adult worms, the DHS-3::GFP can report false-positive results, for example, although ROI 2 shown in Figure 8 was tagged by the DHS-3::GFP reporter, it did not show effective lipid signals in BCARS and 2p-FLIM detections. In our previous 2p-FLIM work66, it required a staining time of 4-8 h to examine lipid particles in Nile-Red stained living worms. By contrast, we found that only 1-2 h of staining time is sufficient for the 2p-FLIM imaging using the multimodal system described here. This is because the two-photon fluorescence excitation efficiency using our BCARS SC beam is much higher. Since the SC wavelength is longer than the laser wavelength we used previously for Nile-Red 2p-FLIM imaging66, less interference of autofluorescence from the gut granules was observed. The SHG-based signals can be possible separated due to their very short lifetimes. Although several fluorescence decay curves from selected particles are demonstrated in this protocol (Figure 8F), the 2p-FLIM signal of each pixel can be converted to phasor components72,73 and analyzed with various machine-learning (ML) approaches, as increasing ML methods have been successfully applied to 2p-FLIM data, including ensemble machine-learning for identifying sub-populations of lipid particles66, U-Net convolutional neural network (CNN) and Uniform Manifold Approximation and Projection (UMAP) for cell classification74,75, and combinatorial analysis of FLIM and spatial transcriptomics76. While the images of Nile-Red-stained living worms are shown as an example here, the introduction of an additional Ti:Sapphire fs laser described in this work can be broadly applied to almost all the fluorescence labeling routinely used in biological studies, facilitating the understanding of fluorescently labeled objects and their chemical compositions, surrounding microenvironments, and the possible metabolic activities and changes.
In summary, the ultra-high information-content chemical bioimaging combining BCARS, 2p-FLIM, and TPEF modalities can provide valuable, coupled chemical and spatial information within live, intact specimens at sub-cellular resolution. Most importantly, all the BCARS/2p-FLIM/TPEF and potentially second harmonic generation (SHG) / third harmonic generation (THG) images can be obtained simultaneously in a spatially co-registered manner, enabling more complex and comprehensive spatial-spectral and cross-modality data analysis. The platform setup and protocols described here offer interested readers the possibility to adapt this multimodal imaging platform to their research. The practical limitations of the presented system include (i) high equipment cost to build a new system, which can be close to1 M USD, largely depending on the costs of lasers (BCARS and the additional Ti:Sapphire lasers), microscope/objective lenses, spectrometer/camera, TPEF/FLIM detectors and TCSPC card, and galvo-scanning systems and other optics; (ii) high instrumental complexity as the whole system contains more than one hundred components and requires customized controlling, image acquisition, and data analysis programs and software; (iii) a nonlinear optical laser expert is required for Laser beam path and system maintenance to ensure the performance of the system; (iv) user expertise is required as the new users need to be properly trained by an experienced BCARS expert before they can acquire high-quality and reproducible data. Although the technical threshold of building such a system based on BCARS could be high compared to other CRI approaches, we believe that future improvements can be done by simplifying the SC and probe beam generation, integrating a 2D laser scanning and descanning geometry, or introducing a time-domain detection scheme.
The authors have nothing to disclose and have no competing financial interests.
W-W.C. acknowledges support from NIH (R21AG086974). M.T.C. acknowledges support from the U.S. Department of Energy (DOE BERDE-SC0022121).
| Camera | Hamamatsu | ORCA-Fusion | sCMOS, 2304×2304 pixels |
| Collection objective | Olympus | LUMPLFLN60XW | Water immersion, 1.0 NA |
| DAQ | Diligent | MCC USB-1208HS | data aquisition card |
| Delay line | Newport | DL225 | Mechanical delay for probe pulse |
| Dichroic | Semrock | FF925-Di01 | combining probe and SC |
| Dichroic | Semrock | FF705-Di01 | detecting epi-mode fluorescence |
| Dichroic | Thorlabs | DMLP550R | seperating the FLIM signal and the TPEF signal |
| Excitation objective | Olympus | UPLSAPO60XWIR | Water immersion, 1.2 NA |
| Filter | Semrock | FF01-1326/SP-25 | Reject the fundamental laser after SHG |
| Filter | Semrock | FF01-758/SP | Reject the fundamental laser before BCARS camera |
| Filter | Thorlabs | FESH0650 | Reject the fundamental laser before the fluorescence modality |
| FLIM PMT | PicoQuant | PMA Hybrid | FLIM single photon detector |
| FPGA | Redpitaya | STEMlab 125-14 | Scan synchronization and triggering |
| Galvo | Thorlabs | GVS101 | For scanning and descanning |
| Grating | Coherent | LightSmyth 1040nm | 1000 Groves/mm, pre-chirp for SC |
| Half-wave plate | Thorlabs | AHWP05M-980 | Polarization control |
| K-cube | Thorlabs | KPA101 | PSD driver for the auto-aligner |
| Laser | Prospective Instruments | FSX-Dual | Dual-output femtosecond laser |
| Laser | Coherent | Mira-900 | Femtosecond laser for fluorescence excitation |
| PBS | Thorlabs | PBS255 | Beam combination for SC and probe |
| PCF | Thorlabs | LMA-PM-5 | 10 cm photonic crystal fiber for SC generation |
| PD | Thorlabs | PDA100A2 | OPO feedback loop |
| Piezo mirror mount | Thorlabs | POLARIS-K1S2P | Piezoelectric Adjuster for the auto-aligner |
| Pre-amp | EdmundOptics | 59-179 | TPEF pre-amplifier |
| Prism | Lambda Research optics | EQP-30SF10 | SC compression |
| PSD | Thorlabs | PDQ80A | Position Sensitive Detector for the auto-aligner |
| Spectrometer | Teledyne Princeton Instruments | IsoPlane 160 | 300 g/mm grating, centered at 700 nm |
| Stage | ASI | MS-2000 | Motorized 3-axis stage |
| TCSPC | PicoQuant | MultiHarp 150 | Photon counting device for FLIM |
| TPEF PMT | Hamamatsu | H9305-03 | TPEF detector |
| Tube Lens | Thorlabs | TL200-2P2 | Used in descanning system |