This article comprehensively explores second near-infrared (NIR-II, 1000–1700 nm) fluorescence imaging, an advanced optical modality offering superior tissue penetration and high-fidelity visualization for preclinical research and emerging clinical applications.
This article comprehensively explores second near-infrared (NIR-II, 1000â1700 nm) fluorescence imaging, an advanced optical modality offering superior tissue penetration and high-fidelity visualization for preclinical research and emerging clinical applications. We detail the fundamental principles that enable NIR-II light to minimize photon scattering, reduce tissue autofluorescence, and achieve a high signal-to-background ratio, surpassing the limitations of visible and NIR-I imaging. The scope encompasses the latest methodological advances, including a diverse palette of organic and inorganic fluorophores, state-of-the-art imaging systems, and their application in vascular mapping, cancer surgery navigation, and real-time therapeutic monitoring. We further address critical challenges in fluorophore optimization and biological barriers, provide a comparative analysis with established imaging techniques, and discuss the promising pathway toward clinical adoption, providing researchers and drug development professionals with a vital resource in this cutting-edge field.
The interaction of light with biological tissue is governed by three fundamental physical phenomena: scattering, absorption, and autofluorescence. These properties collectively determine the penetration depth, spatial resolution, and signal-to-background ratio (SBR) achievable in biomedical optical imaging [1] [2]. When photons enter tissue, they may be absorbed by chromophores such as hemoglobin, lipids, melanin, and water, with their energy converted to heat or re-emitted as fluorescence. Alternatively, photons may be scattered by microscopic variations in refractive index within cellular and extracellular structures, changing their direction of propagation without energy loss. Autofluorescence originates from endogenous fluorophores including collagen, elastin, and flavins, generating background signal that can obscure specific contrast agent signals [3] [4].
The second near-infrared window (NIR-II, 900-1700 nm) has emerged as a superior regime for deep-tissue fluorescence imaging compared to the traditional first near-infrared window (NIR-I, 700-900 nm) and visible spectrum (400-700 nm) [1] [3] [5]. This advantage stems from significantly reduced scattering coefficients, diminished autofluorescence, and lower absorption by dominant tissue chromophores in this spectral region. Consequently, NIR-II fluorescence imaging enables deeper tissue penetration (5-20 mm) and higher spatio-temporal resolution, making it particularly valuable for pre-clinical research and clinical translation [1] [4].
Table 1: Comparison of Optical Windows for Biomedical Imaging
| Optical Property | Visible (400-700 nm) | NIR-I (700-900 nm) | NIR-II (900-1700 nm) |
|---|---|---|---|
| Tissue Scattering | High | Moderate | Low |
| Tissue Autofluorescence | High | Moderate | Low |
| Absorption by Hemoglobin | High | Moderate | Low |
| Absorption by Water | Low | Low | Moderate (increases with wavelength) |
| Penetration Depth | Shallow (1-2 mm) | Moderate (2-5 mm) | Deep (5-20 mm) |
| Spatial Resolution | Low | Moderate | High |
Absorption in biological tissues follows the Beer-Lambert law, where the probability of photon absorption depends on both the concentration of chromophores and their wavelength-dependent extinction coefficients. Principal absorbers in the NIR region include hemoglobin (oxy- and deoxy-forms), lipids, melanin, and water [6] [5]. While absorption is traditionally viewed as detrimental to signal detection, moderate absorption can paradoxically enhance image quality by preferentially attenuating multiply-scattered photons that contribute to background noise [5]. This phenomenon creates an absorption-based gating effect that improves spatial resolution and SBR by suppressing longer-path-length photons [5].
The absorption spectrum of water, the dominant tissue component, reveals several peaks in the NIR region at approximately 980 nm, 1200 nm, 1450 nm, and 1930 nm [5]. Regions surrounding these peaks (900-1000 nm, 1100-1250 nm, 1300-1500 nm, and 1700-1880 nm) benefit from this absorption-mediated background suppression, enabling high-fidelity imaging despite increased signal attenuation [5].
Scattering events in tissues arise from refractive index mismatches at cellular and subcellular interfaces, including membranes, organelles, and collagen fibrils. The scattering coefficient (μs) and anisotropy factor (g) collectively determine the reduced scattering coefficient (μs' = μs(1-g)), which decreases with increasing wavelength according to an approximate power law relationship [5]. This wavelength dependence means that NIR-II photons experience significantly less scattering than visible or NIR-I photons, enabling better preservation of ballistic and snake photons that carry spatial information [1] [4].
Reduced scattering in the NIR-II window allows light to be focused more tightly and deeply within tissue, dramatically improving spatial resolution for microscopic applications. The point spread function becomes narrower, and the degree of light diffusion decreases, permitting high-resolution visualization of subcellular structures and microvascular networks at unprecedented depths [4] [5].
Autofluorescence originates from endogenous fluorophores including collagen, elastin, flavins, lipofuscin, and NAD(P)H [3]. These compounds typically exhibit excitation and emission in the visible spectrum, with minimal emission beyond 900 nm. This property creates an exceptionally low autofluorescence background in the NIR-II window, resulting in significantly higher SBR compared to visible and NIR-I imaging [1] [4]. The combination of reduced scattering, favorable absorption profiles, and diminished autofluorescence in the NIR-II window enables high-contrast imaging of deep-tissue structures with micron-scale resolution [3] [5].
Table 2: Key Optical Properties Across the Near-Infrared Spectrum
| Wavelength Range | Reduced Scattering Coefficient (μs') | Absorption Coefficient (μa) of Water | Autofluorescence Intensity | Recommended Applications |
|---|---|---|---|---|
| NIR-I (700-900 nm) | ~1.0 mmâ»Â¹ | Low | Moderate | Superficial vascular imaging, lymph node mapping |
| NIR-IIa (900-1300 nm) | ~0.5 mmâ»Â¹ | Low to Moderate | Low | Deep-tissue vascular imaging, tumor detection |
| NIR-IIb (1500-1700 nm) | ~0.3 mmâ»Â¹ | Moderate | Very Low | High-resolution microscopic imaging |
| NIR-IIx (1400-1500 nm) | ~0.3 mmâ»Â¹ | High (near water peak) | Very Low | Maximum contrast for bright probes |
| NIR-IIc (1700-1880 nm) | ~0.2 mmâ»Â¹ | Moderate to High | Negligible | Emerging applications with new detectors |
Monte Carlo simulations of photon transport in biological media have refined our understanding of optimal imaging windows within the NIR-II spectrum [5]. These simulations demonstrate that while scattering decreases monotonically with increasing wavelength, strategic utilization of water absorption peaks can further enhance image quality by preferentially suppressing multiply-scattered background photons [5]. This counterintuitive principle has led to the identification of several sub-windows within the broader NIR-II region that offer exceptional imaging performance.
The NIR-IIx window (1400-1500 nm) and NIR-IIc window (1700-1880 nm) are located near water absorption peaks and provide superior image contrast compared to adjacent regions despite higher absolute absorption [5]. Beyond 1880 nm, the NIR-III window (2080-2340 nm) has been proposed as potentially offering the ultimate imaging quality due to minimal scattering, though current detector technology limits its practical implementation [5].
Diagram 1: Hierarchical classification of near-infrared imaging windows, showing the relationship between major windows and sub-windows with their respective wavelength ranges.
The performance advantages of NIR-II imaging can be quantified through several key metrics. Spatial resolution in scattering media improves with increasing wavelength due to reduced scattering, with NIR-II imaging demonstrating approximately 2-3 times higher resolution than NIR-I imaging at equivalent depths [1] [5]. The signal-to-background ratio (SBR) benefits from both reduced autofluorescence and absorption-mediated suppression of multiply-scattered photons, with NIR-II imaging achieving 5-10 times higher SBR compared to NIR-I imaging [1] [4]. Penetration depth increases to 5-20 mm in the NIR-II window, enabling non-invasive visualization of deep-tissue structures in small animal models and potentially in clinical applications [1] [3].
For fluorescence microscopy applications, NIR-II wide-field microscopy in the brain has demonstrated exceptional optical sectioning capability with imaging depths reaching ~1.3 mm through intact skull, representing the deepest in vivo NIR-II fluorescence microscopy achieved to date [5]. This performance underscores the transformative potential of optimized window selection for advanced imaging applications.
Table 3: Key Research Reagents for NIR-II Fluorescence Imaging
| Reagent Category | Specific Examples | Key Properties | Primary Applications |
|---|---|---|---|
| Inorganic Nanomaterials | PbS/CdS QDs [5], Rare-earth doped nanoparticles (RENPs) [1] [3], Carbon nanotubes (CNTs) [1] | High quantum yield, narrow emission, tunable wavelengths, excellent photostability | Multiplexed imaging, deep-tissue angiography, cellular tracking |
| Organic Small Molecules | CH1055 [3], FEB dyes [3], IR-BEMC6P [3], D-A-D structured dyes [1] [3] | Well-defined structures, renal clearance, modifiable targeting, tunable optical properties | Molecular imaging, tumor detection, pharmacokinetic studies |
| Activated Probes | APNO-1080 (NO sensing) [6], Enzyme-activated probes [7] | Low background, target-activated fluorescence, high specificity | Pathological biomarker detection, enzymatic activity monitoring, tumor microenvironment sensing |
| Protein-Hybrid Systems | HSA@CO-1080 [8], Biomimetic fluorescent proteins [8] | Enhanced brightness and photostability, improved biocompatibility, covalent binding | Lymphography, angiography, long-term circulation studies |
This protocol describes wide-field fluorescence microscopy in the NIR-IIx window (1400-1500 nm) for high-resolution imaging of cerebral vasculature through intact skull, achieving exceptional penetration depth and minimal background [5].
Materials Required:
Procedure:
Technical Notes: This method achieves ~1.3 mm penetration depth in mouse brain with excellent spatial resolution. The 1400-1500 nm window provides superior performance compared to NIR-IIb imaging due to optimal balance of scattering reduction and absorption-mediated background suppression [5].
This advanced technique combines ultrasound focusing with NIR fluorescence to achieve spatial resolution beyond the acoustic diffraction limit in deep tissue [9].
Materials Required:
Procedure:
Technical Notes: USF imaging achieves spatial resolution beyond the acoustic diffraction limit (potentially <100 μm) at depths of ~8 mm in tissue. The ION/IOFF ratio of optimal contrast agents reaches 9.1, enabling excellent background suppression [9].
Diagram 2: Experimental workflow for ultrasound-switchable fluorescence (USF) imaging, showing the sequence from sample preparation to super-resolution image reconstruction.
The unique advantages of NIR-II imaging have enabled diverse applications in biomedical research. Multiplexed imaging leverages multiple spectrally distinct NIR-II probes to simultaneously visualize different molecular targets or biological processes [1]. This approach enables researchers to study complex cellular interactions and disease mechanisms with unprecedented detail. Activatable molecular probes that remain dark until specifically triggered by disease biomarkers (e.g., enzymes, reactive oxygen species, or abnormal pH) provide exceptionally high target-to-background ratios for precise disease detection [7].
Emerging technologies continue to push the boundaries of NIR-II imaging. Bioluminescence resonance energy transfer (BRET)-based NIR-II systems eliminate the need for external excitation, further reducing background and enabling ultra-sensitive imaging [1]. Hybrid approaches that combine NIR-II fluorescence with other modalities like photoacoustic imaging [6] or ultrasound [9] provide complementary information for comprehensive tissue characterization.
The future clinical translation of NIR-II imaging will depend on developing fluorophores with optimal combinations of brightness, biocompatibility, and clearance properties [4] [8]. Refined imaging window definitions and the development of specialized optical systems will further enhance performance, potentially revolutionizing diagnostic imaging and therapeutic monitoring in clinical practice.
Fluorescence imaging in the second near-infrared window (NIR-II) has emerged as a transformative modality for in vivo biomedical research, enabling unprecedented capabilities for deep-tissue visualization at high spatial and temporal resolution. This imaging window fundamentally leverages favorable light-tissue interactions within the 1000-1700 nm spectrum, where reduced photon scattering, minimal tissue autofluorescence, and lower photon absorption by biological components occur compared to traditional visible (400-700 nm) and first near-infrared (NIR-I, 700-900 nm) windows [10]. The superior imaging performance within the NIR-II window provides researchers with powerful tools for investigating physiological and pathological processes in live subjects, facilitating applications ranging from vascular imaging and cancer detection to intraoperative guidance and therapeutic monitoring.
The development of NIR-II imaging represents a continuous pursuit of optimal wavelength ranges for probing biological systems. While conventional NIR-I imaging using FDA-approved dyes like indocyanine green (ICG) has provided clinical value for decades, it faces inherent limitations in penetration depth and spatial resolution due to greater tissue scattering and autofluorescence [11] [10]. The extension into NIR-II wavelengths has demonstrated remarkable improvements, achieving approximately 10-fold higher spatial resolution and significantly deeper tissue penetration up to several centimeters, making it particularly valuable for preclinical research and emerging clinical applications [12] [10].
The NIR-II window is formally defined as the spectral range from 1000 nm to 1700 nm [10] [13]. Within this primary window, researchers have established further subdivisions to categorize regions with distinct optical properties and imaging characteristics:
Table 1: Standard NIR-II Window Sub-Regions
| Spectral Region | Wavelength Range (nm) | Key Characteristics |
|---|---|---|
| NIR-II | 1000-1700 | Primary second near-infrared window |
| NIR-IIa | 1300-1400 | Reduced scattering versus NIR-II [14] |
| NIR-IIb | 1500-1700 | Minimal tissue autofluorescence and scattering [13] |
The NIR-IIb sub-region (1500-1700 nm) has received particular attention due to significantly reduced scattering and virtually negligible tissue autofluorescence, which enables superior image contrast compared to shorter wavelengths [15] [13]. However, this advantage must be balanced against increased water absorption in this region, which can attenuate fluorescence signals.
Recent research has further refined the NIR-II window definition and proposed extensions based on detailed analysis of light-tissue interactions:
Table 2: Extended and Proposed NIR-II Classifications
| Spectral Region | Wavelength Range (nm) | Basis/Rationale |
|---|---|---|
| NIR-II (Perfected) | 900-1880 | Incorporates regions around water absorption peaks [13] |
| NIR-IIx | 1400-1500 | Region around water absorption peak at ~1450 nm [13] |
| NIR-IIc | 1700-1880 | Comparable properties to NIR-IIb [13] |
| NIR-III | 2080-2340 | Newly proposed window with potential for superior imaging [13] |
The concept of a "perfected" NIR-II window from 900-1880 nm acknowledges the constructive role of moderate water absorption in enhancing image quality by preferentially attenuating multiply scattered photons that contribute to background noise [13]. The proposal of the NIR-III window (2080-2340 nm) suggests future directions for optical imaging, though current detector technology primarily limits practical implementation to wavelengths below 1700 nm.
This protocol describes methodology for conducting NIR-II fluorescence imaging using multi-wavelength LED excitation sources, based on recent research demonstrating the viability of LEDs as alternatives to laser-based systems [16].
Fluorophore Preparation:
System Calibration:
Animal Preparation:
Image Acquisition:
Image Processing and Analysis:
This protocol describes methodology for implementing NIR-IIb imaging using the long emission tails of commercially available NIR-I dyes, providing an accessible entry point to NIR-II imaging [11] [13].
Dye Preparation:
System Configuration:
In Vivo Imaging:
Data Analysis:
Table 3: Key Research Reagent Solutions for NIR-II Imaging
| Reagent/Material | Function/Application | Examples/Notes |
|---|---|---|
| Organic NIR-II Fluorophores | Small-molecule fluorescent probes | TPA-TQT [16], CH1055-PEG [11], IR-FGP [11], HQL2 AIE dots [15] |
| NIR-I Dyes with NIR-II Tails | Clinically relevant probes for NIR-II imaging | Indocyanine Green (ICG) [11], IRDye800CW [11] |
| Nanoparticle Fluorophores | Inorganic probes with tunable emission | PbS/CdS Quantum Dots [13], Single-Walled Carbon Nanotubes (SWCNTs) [10], Rare-Earth Doped Nanoparticles [10] |
| Protein Carriers | Enhance fluorescence brightness and biocompatibility | Human Serum Albumin (HSA) [11], Bovine Serum Albumin (BSA) [16] |
| Encapsulation Matrices | Improve aqueous solubility and stability | PEG-phospholipids [17], F127 copolymer [17] |
| Targeting Moieties | Enable molecular-specific imaging | Antibodies [12], Peptides, Folic Acid [18] |
| BAY-1316957 | ||
| BMS-764459 | BMS-764459 CRF1 Receptor Antagonist|RUO | BMS-764459 is a high-affinity, selective CRF1 receptor antagonist for research. This product is for Research Use Only. Not for diagnostic or personal use. |
The following diagram illustrates the logical workflow for selecting appropriate NIR-II spectral regions and fluorophores based on specific research applications:
NIR-II Experimental Design Workflow
The precise definition of NIR-II spectral regions and their appropriate application is fundamental to advancing deep-tissue fluorescence imaging research. The standardized subdivisions of the NIR-II window (NIR-IIa, NIR-IIb) provide a framework for selecting optimal imaging parameters based on specific research needs, while newly proposed regions (NIR-IIx, NIR-IIc, NIR-III) suggest future directions for technological development. The experimental protocols outlined herein enable researchers to implement both cutting-edge approaches using specialized NIR-II fluorophores and accessible methods utilizing clinically available dyes with NIR-II emission tails. As the field continues to evolve, the refined understanding of light-tissue interactions across the extended near-infrared spectrum will further enhance the capabilities of in vivo imaging for basic research and drug development applications.
Fluorescence imaging in the second near-infrared window (NIR-II, 1000â1700 nm) represents a significant advancement over traditional NIR-I (700â900 nm) imaging for deep-tissue biological investigations. The core advantage of NIR-II bioimaging stems from fundamentally superior light-tissue interactions in this spectral region, which enable researchers to overcome the pervasive challenges of light scattering and absorption in biological materials [1] [19]. When propagating through living tissue, NIR-II photons experience diminished absorption by biomolecules such as hemoglobin and water, reduced tissue autofluorescence, and suppressed photon scattering compared to both visible light and NIR-I radiation [1] [20]. These physical phenomena collectively enable NIR-II fluorescence imaging to achieve deeper tissue penetration (typically reported in the range of 5-20 mm) and a significantly higher signal-to-background ratio (SBR), providing unprecedented clarity and resolution for non-invasive in vivo studies [1] [19]. This application note details the quantitative metrics validating NIR-II superiority and provides standardized protocols for its application in deep-tissue imaging research.
The performance advantages of NIR-II imaging can be quantified across several key metrics. The following tables consolidate experimental data from recent studies, providing researchers with benchmark values for experimental planning and validation.
Table 1: Comparative Optical Properties of NIR-I and NIR-II Windows
| Performance Metric | NIR-I Window (700-900 nm) | NIR-II Window (1000-1700 nm) | Experimental Basis |
|---|---|---|---|
| Tissue Penetration Depth | < 1 mm [19] | 5â20 mm [1] | Measured in tissue phantoms and in vivo murine models |
| Photon Scattering | High | Significantly reduced (scales as λ-α, α=0.2-4) [11] | Turbid medium measurements |
| Tissue Autofluorescence | Significant | Minimal/negligible [20] | Ex vivo tissue spectroscopy |
| Spatial Resolution | ~10-30 μm at superficial depths | Sub-millimeter at >3 mm depth [1] | Resolution target imaging through tissue |
| Signal-to-Background Ratio (SBR) | Moderate (e.g., ~2-4 for tumors) [11] | High (e.g., 2-4 fold improvement over NIR-I) [11] | In vivo tumor imaging with targeted probes |
Table 2: Performance of Representative NIR-II Fluorophores
| Fluorophore Type | Emission Range (nm) | Quantum Yield (%) | Clearance Pathway | Key Applications |
|---|---|---|---|---|
| Organic D-A-D Dyes (e.g., CH1055) | 900-1600 [11] | 0.03-0.53 [11] | >90% Renal [11] | Vascular imaging, tumor targeting |
| Lanthanide Nanoparticles | 1000-1700 [1] | Varies by composition | MPS/Liver [20] | Multiplexed imaging |
| Quantum Dots (PbS) | >1500 [1] | High (varies) | MPS/Liver [20] | Deep-tissue ovary imaging |
| Repurposed NIR-I Dyes (ICG tail) | 1000-1500 [11] | N/A (tail emission) | Hepatobiliary [20] | Clinical translation studies |
This protocol describes methodology for high-resolution visualization of murine cerebral vasculature using the organic NIR-II fluorophore CH1055-PEG [11].
Materials & Reagents
Procedure
Troubleshooting Notes
This protocol leverages the long fluorescence lifetime of certain probes (e.g., lanthanide-doped nanoparticles, ruthenium complexes) to suppress short-lived background via time-gated detection [22].
Materials & Reagents
Procedure
The following diagram illustrates the fundamental principles that give NIR-II imaging its performance advantage, focusing on photon-tissue interactions.
Successful NIR-II imaging requires a suite of specialized reagents and materials. The following table details key components for building a robust NIR-II imaging pipeline.
Table 3: Essential Reagents and Materials for NIR-II Bioimaging Research
| Item Category | Specific Examples | Function & Application Notes |
|---|---|---|
| NIR-II Fluorophores | CH1055-PEG, IR-FGP, FD-108 [11] | Organic small molecules for high-resolution vascular imaging and rapid renal clearance. Optimal for acute studies. |
| Clinical NIR-I/NIR-II Dyes | Indocyanine Green (ICG), IRDye800CW [11] | Leverage NIR-II "tail" emission (1000-1500 nm) for accelerated clinical translation studies. |
| Inorganic Nanoparticles | Rare-earth doped nanoparticles (RENPs), PbS Quantum Dots [1] | High brightness for multiplexed imaging and sensing; note potential long-term retention [20]. |
| Surface Coating Agents | DSPE-PEG, PS-PEG, Cyclodextrin [20] | Improve biocompatibility, aqueous solubility, and circulation half-life of nanoprobes. |
| Mounting/Imaging Media | FluoroBrite DMEM, OptiClear [21] | Specially formulated media with low autofluorescence for live-cell and ex vivo imaging. |
| Emission Filters | 1000 nm, 1300 nm, 1500 nm long-pass filters | Critical for blocking excitation light and isolating the NIR-II signal from NIR-I background. |
| Lhc-165 | LHC-165 | LHC-165 is a potent Toll-like receptor 7 (TLR7) agonist for research into intratumoral immunotherapy. For Research Use Only. Not for human or veterinary diagnostic or therapeutic use. |
| MK-3903 | MK-3903, CAS:1219737-12-8, MF:C27H19ClN2O3, MW:454.91 | Chemical Reagent |
The transition from NIR-I to NIR-II fluorescence imaging provides a paradigm shift in deep-tissue optical bioimaging, offering quantifiable and significant improvements in both penetration depth and signal-to-background ratio. The experimental protocols and performance metrics outlined in this application note provide a foundational framework for researchers to implement and validate this powerful technology in their investigative workflows. As the palette of biocompatible NIR-II contrast agents continues to expand, incorporating advanced features such as target-specificity and activability, the application of NIR-II imaging is poised to unlock new frontiers in non-invasive biological observation and preclinical drug development.
The emergence of fluorescence imaging in the second near-infrared window (NIR-II, 1000â1700 nm) represents a paradigm shift in biomedical optical imaging, addressing fundamental limitations of traditional visible and NIR-I (700â900 nm) techniques. The historical development of this field stems from the recognition that light propagation in biological tissues is governed by wavelength-dependent scattering and absorption phenomena. Compared to the NIR-I window, the NIR-II window offers significantly reduced photon scattering, diminished tissue autofluorescence, and lower light absorption by endogenous chromophores, enabling unprecedented capabilities for deep-tissue in vivo imaging with high spatial resolution and superior signal-to-background ratios (SBR) [4] [1] [23].
The pioneering work in this field dates to 2003 when Professor Frangioni and Professor Bawendi's team first developed near-infrared quantum dots and demonstrated optimized vascular imaging using wavelengths beyond 1000 nm, hinting at the potential of this optical window [23]. However, the critical breakthrough came in 2009 when Welsher et al. utilized surface-functionalized single-walled carbon nanotubes (SWCNTs) as NIR-II fluorescent agents, achieving high-resolution imaging of tumor vasculature through thick skin layers with minimal background interference [23]. This seminal study established the foundation for NIR-II bioimaging and ignited rapid development of both imaging instrumentation and contrast agents spanning inorganic nanomaterials, organic fluorophores, and molecular complexes [4] [1] [23].
The initial demonstration of NIR-II imaging employed polyethylene glycol (PEG)-modified SWCNTs, which exhibited inherent photoluminescence in the NIR-II region alongside excellent biocompatibility. This pioneering approach revealed the practical advantage of NIR-II imaging: deep tissue penetration with low autofluorescence background, enabling clear visualization of anatomical features previously challenging for optical techniques [23].
Following carbon nanotubes, quantum dots emerged as versatile NIR-II probes with tunable emission properties. Various QD compositions have been developed for biomedical imaging applications:
The table below summarizes the properties and applications of major inorganic NIR-II probes:
Table 1: Evolution of Major Inorganic NIR-II Fluorescence Probes
| Nanomaterial | Emission Range (nm) | Key Advantages | Representative Applications | Key References |
|---|---|---|---|---|
| Single-Walled Carbon Nanotubes | 1000â1400+ | First demonstrated NIR-II in vivo imaging; high photostability | Vasculature imaging, tumor detection | [23] |
| PbS/CdS QDs | 1100â1700 | Bright, tunable emission; core-shell structure improves stability | Deep-tissue imaging in multiple sub-windows (NIR-IIb, NIR-IIc) | [24] |
| AgâS QDs | 1000â1400 | Improved biocompatibility; lower toxicity | Early-stage tumor diagnosis, blood clearance studies | [25] [23] |
| Rare-Earth Doped Nanoparticles | 980â1600 | Narrow emission bands; suitable for multiplexed imaging | Lymph node mapping, vascular imaging | [1] |
While inorganic probes paved the way, organic fluorophores offer potential advantages in biodegradability, pharmacokinetics, and synthetic tunability. Key developments include:
This protocol outlines the synthesis of bright, water-soluble PbS/CdS QDs for deep NIR-II imaging [24].
Materials:
Procedure:
This protocol describes intraoperative assessment of tissue specimens using clinical NIR-II imaging systems like the LightIR platform [26].
Materials:
Procedure:
Table 2: Key Research Reagent Solutions for NIR-II Imaging Studies
| Item | Function/Application | Examples/Specifications |
|---|---|---|
| NIR-II Fluorophores | Provides contrast for deep-tissue imaging | ICG (clinical), IR-1048, PbS/CdS QDs, AgâS QDs, D-A-D dyes, AIEgens |
| Targeting Ligands | Enables specific accumulation in diseased tissues | Antibodies, peptides (e.g., RGD), carbohydrates, small molecules |
| Surface Modification Agents | Improves biocompatibility and circulation | PEG derivatives, phospholipids, multidentate polymers, proteins (e.g., BSA) |
| Imaging Systems | Detection of NIR-II fluorescence | IR VIVO (preclinical), LightIR (clinical), InGaAs detectors (640Ã512 pixels) |
| Surgical Navigation Tools | Intraoperative guidance | Hybrid NIR-II/visible imaging systems, endoscopy-compatible systems |
| Long-Pass Filters | Spectral separation of excitation and emission | LP1000, LP1250, LP1500 (nm cut-on wavelengths) |
| ML-10 | 2-(5-Fluoropentyl)-2-methylmalonic Acid|ML-10 Precursor | Research-grade 2-(5-Fluoropentyl)-2-methylmalonic acid, the tosylate precursor for synthesizing the PET apoptosis tracer 18F-ML-10. For Research Use Only. Not for human or veterinary diagnostic or therapeutic use. |
| NH-bis(PEG3-azide) | NH-bis(PEG3-azide), CAS:1258939-39-7, MF:C16H33N7O6, MW:419.48 g/mol | Chemical Reagent |
The following diagram illustrates the logical workflow from probe design to biomedical application in NIR-II research.
Diagram 1: NIR-II Probe Development Workflow
This diagram contrasts the photon propagation and image formation mechanisms in NIR-I versus NIR-II windows.
Diagram 2: NIR-I vs. NIR-II Imaging Mechanisms
Despite remarkable progress, the clinical translation of NIR-II imaging faces several challenges. Existing NIR-II fluorescent probes still confront issues of insufficient fluorescence brightness, limited wavelength tuning, and suboptimal biocompatibility [4]. The complex synthesis processes, lack of target specificity, and potential long-term metabolic safety risks further complicate clinical adoption [4] [23]. For inorganic probes, particularly those containing heavy metals, comprehensive toxicological profiles and clearance pathways require further investigation [25] [23].
Future developments will likely focus on multiscale molecular design strategies to enhance probe performance [4]. The synergistic combination of NIR-II fluorescence imaging with other modalities like photoacoustic imaging, MRI, and CT will provide complementary information for more precise diagnostics [27] [1]. Furthermore, the integration of NIR-II imaging with therapeutic interventionsâcreating true theranostic platformsârepresents a promising frontier. As probe chemistry and imaging hardware continue to co-evolve, the vision of NIR-II fluorescence imaging becoming a standard tool for precision surgery and deep-tissue diagnosis moves closer to clinical reality [26] [29] [23].
Near-infrared-II (NIR-II, 1000â1700 nm) fluorescence imaging has emerged as a transformative biomedical modality that significantly overcomes the limitations of conventional imaging in the visible (400â700 nm) and first near-infrared (NIR-I, 700â900 nm) windows [30] [31]. Imaging in the NIR-II window benefits from markedly reduced photon scattering, minimal tissue autofluorescence, and lower light absorption by endogenous chromophores such as hemoglobin and water [30] [32]. This synergistic effect enables deeper tissue penetration (up to several centimeters), improved spatial resolution (micron-level), and significantly enhanced signal-to-background ratios (SBR) compared to NIR-I imaging [33] [32]. These advantages make NIR-II imaging particularly powerful for a wide range of preclinical applications, including deep-tissue tumor imaging, cerebrovascular angiography, and image-guided surgery [34] [35]. The development of high-performance fluorescent probes is crucial to fully exploiting the potential of this imaging modality. This review comprehensively outlines the current landscape of NIR-II fluorophores, categorizing them into organic molecules, inorganic nanomaterials, and activatable probes, while providing detailed experimental protocols for their application in deep-tissue imaging research.
Organic small-molecule fluorophores represent one of the most promising classes of NIR-II contrast agents due to their well-defined chemical structures, excellent biocompatibility, favorable pharmacokinetics, and potential for clinical translation [30] [7] [32]. Most are designed with specific molecular architectures that promote intramolecular charge transfer (ICT) to reduce the energy gap between the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO), thereby enabling redshifted absorption and emission [30] [36].
Donor-Acceptor-Donor (D-A-D) Frameworks: These fluorophores feature a strong electron-withdrawing core (Acceptor) symmetrically linked to electron-donating units (Donors) via Ï-conjugated bridges [32]. The benzobisthiadiazole (BBTD) unit is a prototypical acceptor, known for its strong electron-withdrawing character and narrow bandgap [32]. Molecular engineering strategies, such as enhancing donor strength, reducing the D-A distance, and incorporating sterically bulky shielding units (e.g., polyethylene glycol (PEG) chains), are employed to redshift the emission, suppress non-radiative decay, and improve the fluorescence quantum yield (QY) [30] [36]. For instance, engineering a D-A-D molecule (2FT-oCB) with a reduced D-A distance and twisted structure resulted in an emission peak at 1215 nm and an ultralong molecular packing distance exceeding 8 Ã , effectively preventing Ï-Ï stacking and aggregation-caused quenching (ACQ) [36].
Cyanine Derivatives: Cyanine dyes feature a polymethine chain connecting two nitrogen-containing heterocyclic units [32]. While classic dyes like indocyanine green (ICG) have tail emissions in the NIR-II, their QYs are typically low [37]. A breakthrough bioengineering strategy involves complexing cyanine dyes (e.g., IR-783) with recombinant human serum albumin (HSA) domains, particularly domain III (DIII). This "hitchhiking" strategy stabilizes the dye in a hydrophobic pocket, restricting molecular motion and shielding it from aqueous quenching, which dramatically enhances fluorescence brightness and photostability [37].
Aggregation-Induced Emission (AIE) Luminogens: To overcome the common ACQ effect, AIE-active NIR-II fluorophores are designed. These molecules are non-emissive in solution but become highly fluorescent in their aggregated state, as the restriction of intramolecular motion (vibrations and rotations) blocks non-radiative decay pathways [30] [36]. This property is highly beneficial for formulating nanoparticle probes that accumulate at disease sites.
Table 1: Key Characteristics of Organic Small-Molecule NIR-II Fluorophores
| Structural Class | Representative Examples | Emission Range (nm) | Key Advantages | Primary Challenges |
|---|---|---|---|---|
| D-A-D | BBTD-based dyes (e.g., 2FT-oCB) [32] [36] | 1000â1300+ | Tunable structure, good biocompatibility, relatively high QY | Can be hydrophobic, prone to ACQ without design mitigation |
| Cyanine | ICG, IR-783, IR-783@DIII complex [32] [37] | 900â1100+ | Some are clinically approved (ICG), modifiable | Low QY in free state, poor photostability |
| AIE Luminogens | TPE-based D-A-D molecules [30] | 1000â1300 | High brightness in aggregate/ solid state, resists ACQ | Synthetic complexity, potential for large batch-to-batch variance |
| BODIPY/Xanthene | Modified BODIPY dyes [32] | 900â1100 | High molar extinction coefficients, good photostability | Difficulty in achieving emission >1100 nm |
| Chitinase-IN-1 | Chitinase-IN-1, MF:C18H16N4O2S, MW:352.4 g/mol | Chemical Reagent | Bench Chemicals | |
| 1-Stearo-3-linolein | 1-Stearo-3-linolein, MF:C39H72O5, MW:621.0 g/mol | Chemical Reagent | Bench Chemicals |
Inorganic nanomaterials offer high photostability, strong fluorescence intensity, and tunable optical properties, making them valuable for demanding imaging applications [34] [35].
Quantum Dots (QDs): Semiconductor QDs like AgâS, PbS, and Cd-based QDs exhibit size-tunable NIR-II emission, high QYs, and broad absorption profiles [34] [35]. AgâS QDs are particularly notable for their low toxicity and emission tunability from 687 to 1294 nm [35]. However, the long-term toxicity and potential bioaccumulation of heavy metals remain significant concerns for clinical translation [30] [35].
Rare-Earth-Doped Nanoparticles (RENPs): RENPs are typically composed of an inorganic host matrix doped with lanthanide ions (e.g., Yb³âº, Er³âº, Nd³âº). They exhibit unique optical properties, including large Stokes shifts, narrow emission bands, long luminescence lifetimes (micro- to milliseconds), and minimal photobleaching, which are ideal for multiplexed and lifetime-based imaging [34] [1] [35]. Core-shell-shell structures can be designed to enhance emission intensity at specific wavelengths, such as 1525 nm [35].
Single-Walled Carbon Nanotubes (SWCNTs): Semiconducting SWCNTs were among the first materials used for NIR-II bioimaging. They display photostable fluorescence across a broad range of the NIR-II spectrum (1000â1600 nm) [31] [35]. Their surface can be functionalized with phospholipid-PEG to improve water solubility and biocompatibility, enabling applications in vascular imaging [35].
Metal Nanoclusters (MNCs): Gold nanoclusters (Au NCs) consist of a few to hundreds of gold atoms. They are characterized by good biocompatibility, renal clearance potential, and tunable emission extending into the NIR-II window [34] [35]. Their surface chemistry allows for functionalization with therapeutic agents, enabling theranostic applications [35].
Table 2: Key Characteristics of Inorganic NIR-II Fluorophores
| Material Class | Representative Examples | Emission Range (nm) | Key Advantages | Primary Challenges |
|---|---|---|---|---|
| Quantum Dots (QDs) | AgâS QDs, PbS QDs [34] [35] | 900â1300+ | High QY, size-tunable emission, photostable | Potential heavy metal toxicity, long-term retention |
| Rare-Earth Nanoparticles (RENPs) | NaYFâ:Yb/Er/Nd @ NaYFâ [34] [1] | 1000â1600 | Narrow bands, long lifetimes, low background | Lower QY than QDs, complex synthesis |
| Carbon Nanotubes | PEGylated SWCNTs [31] [35] | 1000â1600 | Ultra-broad emission, high photostability | High aspect ratio, potential fiber-like toxicity |
| Metal Nanoclusters | Auââ , Au/Cu alloy clusters [34] [35] | 900â1200 | Good biocompatibility, renal clearable, tunable | Relatively low QY, complex synthesis |
A significant limitation of many NIR-II fluorophores is their "always-on" signal, which can lead to high background noise and reduced specificity [30] [7]. Activatable (or "smart") probes address this issue by remaining in a silent ("off") state until activated by a specific disease-associated biomarker, resulting in a fluorescence turn-on ("on") response at the target site [30] [7].
These probes are engineered to exploit pathological stimuli in the tumor microenvironment (TME) or other diseased tissues. The primary activation mechanisms include:
Purpose: To non-invasively visualize and quantify tumor location, morphology, and biomarker activity in live animal models.
Materials:
Procedure:
SBR = Mean Fluorescence Intensity (Tumor) / Mean Fluorescence Intensity (Background).Purpose: To validate the specific activation of a smart probe by its target enzyme, both in vitro and in vivo.
Materials:
Procedure:
Table 3: Key Research Reagent Solutions for NIR-II Fluorescence Imaging
| Reagent/Material | Function/Description | Example Application |
|---|---|---|
| Indocyanine Green (ICG) | FDA-approved cyanine dye with NIR-II tail emission [32] [37] | Clinical translation benchmark, vascular imaging |
| Recombinant Albumin Domains (e.g., DIII) | Engineered protein domains that bind and enhance cyanine dye brightness [37] | Creating high-performance, biocompatible fluorophores from clinical dyes |
| PEGylated Phospholipids | Amphiphilic polymers for encapsulating and solubilizing hydrophobic fluorophores or nanomaterials [35] | Surface functionalization of SWCNTs, QDs, and organic NPs for improved biocompatibility |
| BBTD Acceptor Core | Strong electron-accepting unit for constructing D-A-D fluorophores [32] [36] | Synthesis of small-molecule NIR-II fluorophores with emissions >1000 nm |
| AgâS Quantum Dots | Low-toxicity inorganic nanoprobe with tunable NIR-II emission [35] | High-resolution deep-tissue and vascular imaging |
| NaYFâ:Yb/Er/Nd Nanoparticles | RENPs with long lifetime and narrow emission bands [34] [1] | Multiplexed imaging, lifetime imaging |
| NIR-II Fluorescence Imager | InGaAs camera-based system with NIR lasers and filters [31] | Essential equipment for all in vivo and in vitro NIR-II imaging studies |
| Cycloshizukaol A | Cycloshizukaol A, CAS:150033-85-5, MF:C32H36O8 | Chemical Reagent |
| 9-cis-Retinol-d5 | 9-cis-Retinol-d5, MF:C20H30O, MW:291.5 g/mol | Chemical Reagent |
Diagram Title: Activation Mechanisms of Smart NIR-II Probes
Diagram Title: In Vivo NIR-II Imaging Experiment Workflow
The NIR-II fluorophore landscape is richly diverse, encompassing organic small molecules with high translational potential, robust inorganic nanomaterials for multiplexed imaging, and sophisticated activatable probes for unprecedented specificity. The choice of fluorophore hinges on the specific research question, balancing factors such as emission wavelength, brightness, biocompatibility, and functional design. The experimental protocols outlined provide a foundational roadmap for researchers to conduct high-quality NIR-II bioimaging studies. As molecular engineering and bioengineering strategies continue to advance, the next generation of NIR-II probes will undoubtedly unlock deeper insights into biological processes and accelerate the development of precision diagnostics and therapeutics.
Bioimaging in the second near-infrared window (NIR-II, 950â1700 nm) represents a transformative approach for non-invasive, real-time investigation of biological processes with exceptional spatiotemporal resolution and deep tissue penetration [33]. This imaging modality has demonstrated significant advantages over conventional UV-Vis-NIR and NIR-I windows, primarily due to reduced photon scattering, minimal tissue autofluorescence, and diminished light absorption by tissue chromophores such as hemoglobin and water [38]. These physical properties enable researchers to achieve deeper tissue penetration (5â20 mm) while maintaining high spatial resolution, making NIR-II imaging particularly valuable for preclinical research, drug development, and emerging clinical applications [1] [39].
The evolution of NIR-II imaging instrumentation has progressed along two complementary trajectories: advanced detection systems capable of capturing faint NIR-II signals, and integrated navigation platforms that translate this information into surgical guidance. InGaAs-based cameras have emerged as the gold standard for detecting NIR-II fluorescence, offering superior sensitivity in the 900-1700 nm range compared to silicon-based detectors [40] [39]. Concurrently, augmented reality surgical navigation systems have developed to leverage this deep-tissue imaging capability, providing surgeons with intuitive, real-time visualization of subsurface anatomical structures and pathological features [41]. This combination of sophisticated detection and visualization technologies is rapidly advancing the field of image-guided interventions and precision medicine.
The exceptional performance of NIR-II bioimaging stems from fundamental interactions between light and biological tissues. As photons traverse tissue, they undergo both absorption and scattering events, which collectively determine the penetration depth and image resolution achievable at different wavelengths. In the NIR-II window, scattering is significantly reduced compared to visible and NIR-I wavelengths, leading to clearer images with enhanced spatial resolution [40]. Table 1 quantifies the key advantages of NIR-II imaging relative to other spectral windows.
Table 1: Comparison of Imaging Spectral Windows
| Parameter | Visible (400-700 nm) | NIR-I (700-900 nm) | NIR-II (1000-1700 nm) |
|---|---|---|---|
| Penetration Depth | Shallow (1-2 mm) | Moderate (2-5 mm) | Deep (5-20 mm) |
| Spatial Resolution | Limited by scattering | Moderate | High (â¼Î¼m) |
| Tissue Autofluorescence | High | Moderate | Low |
| Photon Scattering | Significant | Moderate | Reduced |
| Absorption by Hemoglobin | High | Moderate | Low |
The diminished scattering in the NIR-II window directly translates to superior image clarity. As illustrated in practical experiments, NIR-I wavelengths tend to scatter extensively, resulting in 'fuzzier' images, while NIR-II/SWIR wavelengths yield 'crisper' images with well-defined vascular anatomy [40]. This scattering reduction enables researchers to visualize fine biological structures including capillary networks, tumor vasculature, and neural anatomy with unprecedented detail in thick, biologically relevant tissues [39].
Beyond improved resolution, NIR-II imaging offers enhanced signal-to-background ratio (SBR) due to significantly reduced tissue autofluorescence. Many biological compounds exhibit natural fluorescence in the visible spectrum (collagen, elastin, NADPH), while water absorption peaks around 900-1000nm further complicate NIR-I imaging [39]. The NIR-II window largely avoids these confounding factors, resulting in cleaner signals with higher contrast between target and background tissues.
Contrast in NIR-II imaging can be further enhanced through specialized probes designed to accumulate in specific tissues or respond to pathological conditions. Activatable molecular probes that remain silent until reaching their target environment are particularly valuable for improving diagnostic accuracy by minimizing background signal [7]. The combination of favorable tissue optical properties and advanced contrast agents enables researchers to achieve robust, high-fidelity imaging in complex biological systems.
The choice of detector technology critically influences the performance and applicability of NIR-II imaging systems. Standard silicon-based CCD and sCMOS detectors, while excellent for visible and NIR-I imaging, exhibit rapidly declining quantum efficiency beyond 700nm, typically falling to 20-40% at 900nm [39]. This limitation stems from the fundamental properties of silicon, where longer wavelengths often pass through the photosensitive layer without generating photoelectrons.
Table 2: Performance Comparison of NIR-II Detector Technologies
| Detector Type | Spectral Range | Quantum Efficiency | Key Advantages | Limitations |
|---|---|---|---|---|
| Si-CCD (Deep Depleted) | 400-900 nm | ~50% at 850nm | Compatible with visible imaging; reduced etaloning | Limited to <1000nm; slower readout |
| InGaAs Camera | 900-1700 nm | >70% (900-1700nm) [39] | Broad NIR-II coverage; fast frame rates (up to 600fps) [39] | Requires cooling; higher cost |
| Scientific-Grade InGaAs | 900-1700 nm | ~80-85% (950-1100nm) [42] | Deep cooling (-85°C); low dark noise; precision exposure control [40] | Specialized infrastructure; higher cost |
For imaging beyond 1000nm, Indium Gallium Arsenide (InGaAs) detectors have emerged as the preferred technology. These detectors typically offer quantum efficiency greater than 70% across the entire NIR-II window (900-1700nm), with scientific-grade cameras achieving 80-85% in the 950-1100nm range [42] [39]. This high sensitivity is crucial for detecting the relatively weak signals that emerge from deep tissues after significant attenuation.
A critical consideration for InGaAs cameras is the management of dark current, which can be substantial at room temperature. Advanced scientific InGaAs cameras incorporate maintenance-free thermoelectric cooling systems capable of reaching -85°C, coupled with proprietary cold shield designs and vacuum technology to minimize dark noise [40]. This deep cooling preserves signal-to-noise ratio during long exposures, enabling researchers to capture faint signals from deep tissue structures.
Despite the higher quantum efficiency of InGaAs cameras compared to silicon detectors, an intriguing paradox emerges in deep-tissue imaging scenarios. Under identical imaging conditions using MgGeO3:Yb3+ short-wave infrared persistent luminescent phosphor as a contrast agent, InGaAs cameras surprisingly exhibited "notably inferior performance overall in imaging deep-tissue features" compared to Si CCD cameras when imaging through thick chicken breast tissues (5â20 mm) and mice bodies (10â20 mm) [42]. This counterintuitive finding was attributed to the "pronounced interference of [the] inherently high dark current" in InGaAs detectors, particularly problematic when dealing with faint imaging signals [42].
This performance limitation can be overcome with sufficiently intense SWIR imaging signals, where the InGaAs camera outperforms the Si camera in terms of clarity in tissues up to 10mm thick [42]. This finding highlights the importance of matching detector technology to specific experimental requirements, considering factors such as expected signal intensity, tissue depth, and required frame rate.
The translation of NIR-II imaging data into surgical guidance has led to the development of augmented reality (AR) surgical navigation systems. These platforms integrate preoperative imaging data with real-time surgical visualization, creating an intuitive interface for navigating complex anatomical relationships. One such system proposed for extra-ventricular drainage (EVD) surgery provides "intuitive guidance for the surgical target point, entry point, and trajectory" through AR devices (tablet or HoloLens glasses) [41].
These AR navigation systems employ "virtual object automatic scanning" to achieve remarkable accuracy of 1 ± 0.1 mm, addressing the stability and precision limitations that have historically hampered surgical AR applications [41]. Furthermore, the integration of deep learning-based U-Net segmentation networks enables automatic identification of pathological features such as hydrocephalus locations, with reported recognition accuracy, sensitivity, and specificity of 99.93%, 93.85%, and 95.73%, respectively [41]. This combination of precise registration and automated segmentation creates a powerful platform for translating NIR-II imaging data into surgical action.
Diagram 1: AR Surgical Navigation Workflow. This flowchart illustrates the integrated process from clinical need definition through procedure completion, highlighting key technological components including AI segmentation and AR registration.
The effectiveness of NIR-II imaging depends critically on the contrast agents employed, with ongoing development focusing on improving brightness, biocompatibility, and target specificity. Table 3 summarizes major categories of NIR-II contrast agents and their characteristics.
Table 3: NIR-II Contrast Agent Classes and Properties
| Probe Type | Emissions (nm) | Key Advantages | Limitations | Representative Applications |
|---|---|---|---|---|
| Single-Walled Carbon Nanotubes (SWNTs) | 1000-1400+ | Good emissivity; can be PEGylated for reduced toxicity [40] | Potential long-term retention concerns | Vascular imaging [40] |
| Rare Earth-Doped Nanoparticles | Tunable: 950-1100nm (MgGeO3:Yb3+) [42] | Low toxicity; narrow emission bands; persistent luminescence [42] [1] | Complex synthesis; potential heavy metal content | Deep-tissue imaging [42] |
| Quantum Dots | Tunable via size [40] | Bright emissions; size-tunable properties | Potential heavy metal toxicity; clearance concerns | Tumor imaging [40] |
| Organic Fluorophores | 1000-1300+ | Well-defined structures; tunable properties; better biocompatibility [7] | Generally lower brightness than inorganic options | Targeted molecular imaging [7] |
| Indocyanine Green (ICG) | ~800-850nm (extends to NIR-II) [39] | FDA-approved; established safety profile | Limited to shorter NIR-II wavelengths | Clinical perfusion imaging [39] |
Recent innovations in contrast agent design have focused on "activatable molecular probes" that address the limitation of "always-on" fluorescence, which leads to background noise and compromised diagnostic accuracy [7]. These smart probes remain quiescent until activated by specific pathological conditions such as abnormal enzyme activity or pH changes, significantly improving target-to-background ratio for precise disease detection.
For multiplexed imaging applications, researchers have developed probes with distinct emission profiles or lifetime characteristics, enabling simultaneous tracking of multiple biological targets. Doped rare-earth nanoparticles are particularly valuable for this application, offering narrow emission bands that can be separated with appropriate optical filters [1]. This multiplexing capability is invaluable for studying complex biological processes involving multiple cell types or molecular targets.
This protocol describes methodology for high-resolution imaging of vascular anatomy in small animal models, adapted from published studies [40] [39].
Materials Required:
Procedure:
Contrast Agent Administration:
Animal Preparation:
Imaging System Setup:
Image Acquisition:
Data Processing:
Troubleshooting Notes:
This protocol outlines the validation procedure for augmented reality surgical navigation systems integrated with NIR-II imaging data [41].
Materials Required:
Procedure:
System Calibration:
Data Integration:
Registration Accuracy Assessment:
Navigation Workflow Testing:
Clinical Validation:
Diagram 2: NIR-II Imaging Signal Path. This diagram illustrates the complete pathway from excitation to data processing, highlighting key advantages including reduced scattering and deep penetration.
The combination of advanced NIR-II imaging instrumentation and augmented reality navigation systems has enabled diverse applications across biomedical research and clinical translation. In drug development, NIR-II imaging provides unparalleled capability for monitoring drug distribution, pharmacokinetics, and treatment response in live animal models. The ability to track fluorescently-labeled therapeutic compounds in real-time through thick tissues offers significant advantages over traditional endpoint measurements [33].
In oncology research, NIR-II imaging enables detailed visualization of tumor vasculature, microenvironment, and heterogeneity. The high spatial and temporal resolution allows researchers to monitor dynamic processes such as drug uptake, tumor perfusion, and metastatic progression. When combined with multiplexed imaging approaches, researchers can simultaneously track multiple cell populations or molecular targets, providing comprehensive understanding of complex biological systems [1].
For neuroscience applications, NIR-II imaging facilitates non-invasive assessment of brain structure and function through the intact skull, overcoming the scattering that limits visible light imaging [39]. This capability is particularly valuable for longitudinal studies of disease progression, neurodevelopment, and therapeutic intervention in small animal models.
The integration of these imaging capabilities with augmented reality guidance systems creates powerful platforms for translational research. Surgeons can leverage deep-tissue information provided by NIR-II imaging to guide precise interventions, while researchers gain valuable insights into disease pathology and treatment mechanisms. This synergy between diagnostic imaging and therapeutic intervention represents a significant advancement in precision medicine.
The second near-infrared window (NIR-II, 1000â1700 nm) fluorescence imaging leverages reduced photon scattering, minimal tissue autofluorescence, and lower light absorption in biological tissues compared to visible and NIR-I windows. This results in superior penetration depth (up to 20 mm) and micron-level spatial resolution, enabling unparalleled real-time visualization of deep-tissue vascular architectures and dynamic blood flow patterns [35] [32]. This application is foundational for studying cardiovascular diseases, tumor angiogenesis, and cerebrovascular dynamics.
NIR-II imaging allows for precise quantification of vascular parameters. The following table summarizes key performance metrics for vascular imaging using different NIR-II probes:
Table 1: Performance of NIR-II Fluorophores in Vascular Imaging
| Fluorophore Type | Excitation/Emission (nm) | Spatial Resolution | Penetration Depth | Key Vascular Application |
|---|---|---|---|---|
| SWCNTs (PEGylated) [35] | N/A, Broad NIR-II emission | ~100 µm | Deep tissue imaging | First-generation probe for vascular imaging |
| AgâS Quantum Dots [35] | Tunable 687-1294 nm | High (micron-level) | Several millimeters | Non-specific tumor detection and vascular mapping |
| Biomimetic NIR-II FPs (HSA@CO-1080) [8] | 1044/1079 nm | High temporal-spatial resolution | Enabled by 1064 nm excitation | High-contrast lymphography and angiography |
| Organic Small Molecules (LS series) [43] | Up to 1218 nm emission | High | Favorable for BBB penetration | Visualization of brain vasculature and pathologies |
Objective: To perform real-time, high-resolution imaging of the vascular system in a live mouse model.
Materials:
Procedure:
Troubleshooting Tips:
The chaotic tumor vasculature and dysregulated pathophysiology of the tumor microenvironment (TME) create opportunities for passive accumulation (Enhanced Permeability and Retention effect) and active targeting of NIR-II probes [32]. This allows for precise delineation of primary and metastatic tumors, visualization of the TME, and guidance for surgical resection with high signal-to-background ratios (SBR) [35] [26].
NIR-II probes have been successfully applied across various cancer models. The table below compares different probes used in tumor imaging:
Table 2: NIR-II Probes in Tumor Imaging and Image-Guided Surgery
| Probe Type / Strategy | Target / Mechanism | Key Finding | Potential Clinical Utility |
|---|---|---|---|
| Au NCs-Pt Nanodrug [35] | GSH Scavenging & Pt Delivery | Visualized Pt transport in deep tumors via NIR-II; sensitized cells to chemo | Theranostic platform for chemotherapy |
| RENPs@Lips [35] | Passive Accumulation | Double emission (1064/1345 nm); favorable biocompatibility & excretion | Preclinical evaluation of physiological/pathological processes |
| Organic Small Molecules [32] | EPR effect; some are targetable | Fast clearance, superior biocompatibility; Milestone: ICG used in patient liver cancer surgery [32] | Most promising for clinical translation in guided surgery |
| Clinical LightIR System [26] | N/A (Imaging System) | Resolved ~250 µm features; detected ICG to depths â¥4 mm under ambient light | Back-table tumor margin assessment during surgery |
Objective: To intraoperatively delineate primary tumor margins and identify microscopic metastatic deposits in a preclinical model.
Materials:
Procedure:
Troubleshooting Tips:
Understanding the dynamic interactions between immune cells and tumor cells within the TME is critical for developing immunotherapies. Intravital Microscopy (IVM) powered by NIR-II and Fluorescence Lifetime Imaging Microscopy (FLIM) enables real-time visualization of these interactions at single-cell resolution [44]. FLIM-FRET is particularly powerful for quantifying molecular interactions, such as those between immune checkpoint proteins (e.g., PD-1/PD-L1), which is a better predictor of immunotherapy response than simple protein expression levels [45].
Advanced imaging techniques are revealing the dynamics of the immune response.
Table 3: Advanced Imaging Techniques for Immunological Studies
| Technique | Measured Parameter | Key Application in Immunology | Research Finding |
|---|---|---|---|
| FLIM-FRET [45] | Fluorescence lifetime change due to energy transfer | Quantifying PD-1/PD-L1 protein interaction | Interaction level, not just expression, correlates with ICI efficacy [45] |
| Intravital Microscopy (IVM) [44] | Real-time cell motility, migration, and interactions | Visualizing T cell â tumor cell interactions in live animals | Enables study of immune cell infiltration, killing, and evasion |
| FLIM Flow Cytometry [46] | Fluorescence lifetime of single cells in flow | High-throughput analysis of tumor cell heterogeneity | Distinguished subpopulations in rat glioma; captured drug-induced nuclear changes |
Objective: To quantify the interaction between PD-1 and PD-L1 in a live cell co-culture system using FLIM-FRET.
Materials:
Procedure:
Troubleshooting Tips:
Table 4: Key Reagents and Materials for NIR-II and Intravital Imaging
| Item Name | Function / Description | Example Use Case |
|---|---|---|
| NIR-II Organic Small Molecules [32] [43] | Low molecular weight (<500 Da) fluorophores with tunable emission up to 1200+ nm; favorable pharmacokinetics. | Vascular imaging, tumor targeting, and brain imaging due to potential BBB penetration. |
| Biomimetic NIR-II Fluorescent Proteins [8] | Synthetic dyes (e.g., CO-1080) covalently bound to serum albumin (HSA), enhancing brightness and photostability. | Creating long-circulating probes for high-contrast lymphography and angiography. |
| Rare-Earth-Doped Nanoparticles (RENPs) [35] | Inorganic nanoparticles with large Stokes shifts, narrow emissions, and minimal photobleaching. | Multiplexed imaging with distinct emission wavelengths (e.g., 1064 nm & 1345 nm). |
| Clinical LightIR Imaging System [26] | Compact InGaAs camera-based system for real-time NIR-II imaging under ambient light. | Intraoperative guidance and back-table assessment of tumor margins in a clinical setting. |
| FLIM Flow Cytometer [46] | High-throughput system combining flow cytometry with FLIM, analyzing >10,000 cells/second. | Rapid, label-free analysis of tumor cell heterogeneity and drug response based on metabolic states. |
| Epiaschantin | Colchicine Derivative 5-[(4r)-4-(3,4,5-Trimethoxyphenyl)tetrahydro-1h,3h-furo[3,4-c]furan-1-yl]-1,3-benzodioxole | High-purity 5-[(4r)-4-(3,4,5-trimethoxyphenyl)tetrahydro-1h,3h-furo[3,4-c]furan-1-yl]-1,3-benzodioxole for research. For Research Use Only. Not for human or veterinary use. |
| Rg3039 | Rg3039, CAS:1466525-84-7 | Chemical Reagent |
The following diagram illustrates the strategic concepts behind designing advanced NIR-II fluorophores, moving beyond traditional methods.
This diagram outlines the workflow and principle of using FLIM-FRET to quantify immune checkpoint protein interactions.
Fluorescence-guided surgery (FGS) and image-guided drug delivery (IGDD) represent transformative paradigms in precision medicine. The core of this advancement lies in the utilization of the second near-infrared window (NIR-II, 1000â1700 nm), which offers superior deep-tissue imaging capabilities compared to traditional visible light (400â700 nm) and the first near-infrared window (NIR-I, 700â900 nm) [4] [31]. NIR-II light experiences significantly reduced absorption, scattering, and autofluorescence from biological tissues, enabling deeper tissue penetration (5â20 mm) and higher signal-to-background ratios (SBR) for high-contrast imaging [18] [1]. This review details specific application notes and experimental protocols for leveraging NIR-II fluorescence imaging in surgical guidance and targeted drug delivery, providing a practical toolkit for researchers and drug development professionals.
The principal challenge in curative cancer surgery is the complete resection of malignant tissue while preserving healthy structures. NIR-II FGS addresses this by providing real-time, high-resolution visualization of tumors, vasculature, and critical anatomy. The enhanced penetration and reduced scattering of NIR-II light allow surgeons to distinguish sub-millimeter tumor margins and micrometastases that are otherwise invisible to the naked eye or under NIR-I imaging [4] [47]. This is particularly valuable for tumors located in highly vascularized or anatomically complex regions.
The table below summarizes the key advantages of NIR-II imaging over other optical windows.
Table 1: Quantitative Comparison of Optical Imaging Windows for Biomedical Applications
| Imaging Parameter | Visible (400-700 nm) | NIR-I (700-900 nm) | NIR-II (1000-1700 nm) |
|---|---|---|---|
| Tissue Penetration Depth | Shallow (1-2 mm) | Moderate (1-5 mm) | Deep (5-20 mm) [1] |
| Tissue Autofluorescence | Very High | High | Negligible/Low [31] |
| Photon Scattering | Very High | High | Reduced [18] |
| Spatial Resolution | Low | Moderate | High (up to ~25 μm) [4] |
| Signal-to-Background Ratio (SBR) | Low | Moderate | High [31] |
Objective: To precisely resect a solid tumor in a murine model using a NIR-II fluorescent probe for intraoperative guidance.
Materials:
Procedure:
SBR = (Mean Intensity of Tumor) / (Mean Intensity of Adjacent Normal Tissue).Image-guided drug delivery (IGDD) systems merge diagnostic capability with therapeutic intervention, forming a "theranostic" platform. NIR-II fluorescence is ideally suited for monitoring the pharmacokinetics and pharmacodynamics of drug carriers in real-time [18]. This allows researchers to non-invasively track the accumulation of a drug-loaded nanocarrier in a tumor, monitor its distribution within the tumor microenvironment, and even trigger drug release at the desired site, thereby optimizing the therapeutic index and minimizing off-target effects [18] [49].
The table below lists essential materials for constructing a NIR-II IGDD system.
Table 2: Research Reagent Solutions for NIR-II Image-Guided Drug Delivery
| Item | Function/Description | Examples |
|---|---|---|
| NIR-II Fluorophore | Serves as the imaging agent; can be conjugated to or encapsulated within a drug carrier. | Organic dyes (e.g., CH-4T [4]), Ag2S Quantum Dots [31], Lanthanide-doped Nanoparticles (DCNPs) [1]. |
| Drug Nanocarrier | Carries the therapeutic payload; can be engineered for targeted delivery. | Liposomes, Polymeric Nanoparticles (e.g., PLGA), Dendrimers, Extracellular Vesicles [18]. |
| Targeting Moiety | Directs the nanocarrier to specific biomarkers on cancer cells. | Antibodies (e.g., anti-EGFR), Peptides (e.g., cRGD), Folic Acid [18] [48]. |
| Stimuli-Responsive Material | Enables controlled drug release in response to specific triggers. | pH-sensitive polymers, ROS-cleavable linkers, light-activatable coatings [4]. |
| NIR-II Imaging System | For in vivo, real-time tracking of the theranostic agent. | InGaAs camera, NIR-II laser excitations (808, 980 nm), spectral filters [31] [1]. |
| DBCO-PEG8-acid | DBCO-PEG8-acid|Bifunctional PEG Linker for Click Chemistry | |
| MIDD0301 | MIDD0301 |
Objective: To visualize the accumulation of a NIR-II-labeled drug carrier in a tumor and assess its drug release profile.
Materials:
Procedure:
The following diagrams, generated using DOT language, illustrate the core concepts and experimental workflows described in this application note.
This diagram illustrates the integrated "diagnose-treat-monitor" cycle enabled by NIR-II theranostic probes.
This flowchart outlines the key steps for a typical NIR-II image-guided drug delivery experiment, from formulation to analysis.
The integration of NIR-II fluorescence imaging with surgical guidance and targeted drug delivery systems marks a significant leap forward in theranostics. The protocols and application notes provided here offer a foundational framework for researchers to implement these cutting-edge technologies. As the field progresses, overcoming challenges in probe biocompatibility, complex synthesis, and clinical translation through innovative molecular design and interdisciplinary collaboration will be crucial [4]. The future of precision medicine lies in these multi-functional platforms that allow for simultaneous diagnosis, treatment, and monitoring, ultimately leading to improved patient outcomes.
Fluorescence imaging in the second near-infrared window (NIR-II, 1000â1700 nm) has emerged as a transformative biomedical imaging modality, offering superior spatial resolution and tissue penetration compared to traditional NIR-I imaging [51] [52]. This enhanced performance stems from significantly reduced photon scattering, lower tissue autofluorescence, and diminished light absorption in biological tissues within this spectral region [52] [53]. However, the development of high-performance NIR-II fluorophores has been persistently hampered by the quantum yield bottleneck â the inherent tendency of fluorophores with extended Ï-conjugation systems to exhibit low fluorescence efficiency due to increased non-radiative decay pathways [37] [54].
The quantum yield (QY) of a fluorophore, defined as the ratio of photons emitted to photons absorbed, directly determines imaging brightness and consequently impacts detection sensitivity, signal-to-noise ratio, and ultimately, diagnostic accuracy [54]. For NIR-II organic fluorophores, quantum yields typically range from 0.01% to 3.89%, significantly lower than their visible-light counterparts [54]. This limitation has catalyzed intensive research into molecular engineering strategies designed to enhance the brightness of NIR-II fluorophores while maintaining their deep-tissue penetration capabilities, forming a critical frontier in biomedical optics research [55] [37].
A primary approach to enhancing quantum yield involves reducing non-radiative decay by imposing structural rigidity on fluorophore systems. The introduction of rigid cyclohexanol structures at the center of cyanine/polymethine dyes restricts molecular vibration and rotation, key contributors to non-radiative energy loss [37]. Similarly, engineering increased dihedral angles in donor-acceptor-donor (D-A-D) architectures through the incorporation of rigid alkyl thiophene moieties has demonstrated significant quantum yield improvements by minimizing collisional quenching [37].
Table 1: Structural Rigidification Strategies and Their Impact on Quantum Yield
| Engineering Strategy | Molecular Target | Mechanism of Action | Reported QY Improvement |
|---|---|---|---|
| Cyclohexanol Incorporation | Cyanine/Polymerhine Dyes | Restricts molecular vibration & rotation | ~2-3 fold increase [37] |
| Dihedral Angle Control | D-A-D Structured Dyes | Minimizes collisional quenching | Significant enhancement [37] |
| Aggregation-Induced Emission (AIE) | AIE Luminogens | Restricts intramolecular rotation | Dramatic fluorescence enhancement [56] |
Molecular shielding represents another powerful strategy for quantum yield enhancement. The incorporation of shielding units (S) containing long polyethylene glycol (PEG) groups or dialkoxy-substituted benzene/fluorene-based moieties in S-D-A-D-S architectures creates a protective microenvironment around the fluorophore core [37]. These shielding groups function through multiple mechanisms: they improve molecular dispersion in aqueous solutions, reduce Ï-Ï stacking-induced quenching, and create steric hindrance that minimizes collisions with quenching species such as water molecules and dissolved oxygen [37] [54]. The strategic placement of these shielding units is critical to their efficacy, as they must provide protection without disrupting the electronic transitions responsible for fluorescence.
An innovative bioengineering approach involves using recombinant proteins as molecular chaperones to enhance fluorophore brightness. Research has identified that Domain III (DIII) of human serum albumin serves as a high-affinity binding pocket for cyanine dyes [37]. When fluorophores are encapsulated within this hydrophobic domain, they experience multiple benefits: stabilization of the excited state, reduced non-radiative decay through restricted molecular motion, and protection from collisional quenching [37]. This protein-dye complex formation can enhance fluorescence brightness by approximately 1.6-fold compared to free dye [37]. Furthermore, this strategy allows genetic editing to tune complex size and pharmacokinetics while providing a biocompatible coating that improves biosafety and clinical translation potential [37].
Systematic extension of Ï-conjugation systems combined with strategic heteroatom substitution has proven effective in bathochromically shifting absorption and emission while maintaining reasonable quantum yields. The incorporation of dicyanomethylene-4H-benzopyran (DCMO) and its heteroatom variants (DCMS with sulfur, DCMSe with selenium) as electron-withdrawing acceptors in conjunction with extended polymethine bridges enables emission tuning up to 1060 nm while preserving fluorescence efficiency [57]. Heavier chalcogen atoms (S, Se) promote greater electronic delocalization and narrower HOMO-LUMO gaps, enabling longer wavelength emissions. However, this approach must be carefully balanced against the "cyanine limit" â the point at which further conjugation leads to symmetry breaking and fluorescence quenching [54].
Table 2: Photophysical Properties of Representative Engineered NIR-II Fluorophores
| Fluorophore | λabs/λem (nm) | Quantum Yield (%) | Brightness (Mâ»Â¹ cmâ»Â¹) | Key Structural Features |
|---|---|---|---|---|
| Flav7 | 1027/1053 | 0.610 | 1470.1 | Rigidified heterocyclic terminals [54] |
| LZ-1105 | 1058/1100 | 3.890 | 5757.2 | Optimized shielding groups [54] |
| HC-1222 | 1180/1222 | 0.016 | 18.8 | Extended conjugation [54] |
| IR-783@DIII | 783/~1000 | Enhanced 1.6x | Significantly increased | Protein chaperoning [37] |
| CL-S3 | ~1089/1060 | N/A | High SBR (43.5) | Extended Ï-conjugation [57] |
Principle: This method determines fluorophore quantum yield relative to a reference standard with known quantum yield in the same solvent system [54].
Materials:
Procedure:
Technical Notes: Ensure complete degassing of solutions to eliminate oxygen quenching. Maintain low analyte concentrations to avoid aggregation effects. Account for solvent-dependent refractive index variations [54].
Principle: This protocol describes the formation and characterization of protein-fluorophore complexes for brightness enhancement based on the albumin chaperoning strategy [37].
Materials:
Procedure:
Technical Notes: Optimal complex formation may require temperature optimization. Free dye must be completely removed for accurate quantum yield determination. Binding affinity varies significantly among dye structures â DiR shows notably weak binding (Kd = 144 μM) compared to IR-783 (Kd = 0.57 nM with DIII) [37].
Table 3: Essential Research Reagents for NIR-II Fluorophore Development
| Reagent/Material | Function/Application | Specific Examples | Key Characteristics |
|---|---|---|---|
| Cyanine Dyes | NIR-II fluorophore core structures | IR-783, ICG, IR-1061, IR-1048, Cy7, Cy7.5 [37] [54] | Tunable emission, high extinction coefficients |
| Recombinant Albumin Domains | Fluorophore chaperoning for brightness enhancement | Domain III (DIII) of HSA [37] | High-affinity dye binding (Kd = 0.57 nM for IR-783) |
| Shielding Groups | Reduce collisional quenching | PEG groups, dialkoxy-substituted benzene/fluorene [37] | Steric hindrance, improved solubility |
| Heterocyclic Terminal Moieties | Bathochromic shift engineering | DCMO, DCMS, DCMSe [57] | Electron-withdrawing capability, heavy atom effect |
| Reference Standards | Quantum yield determination | IR-26 in DCE (QY = 0.05%) [54] | Established reference for NIR-II QY measurements |
| Solvent Systems | Photophysical characterization | DCM, DCE, MeOH, DMSO, PBS [54] | Varying polarity for solvatochromism studies |
The molecular engineering strategies outlined in this application note provide a robust toolkit for addressing the persistent quantum yield bottleneck in NIR-II fluorescence imaging. Through integrated approaches combining structural rigidification, steric shielding, protein chaperoning, and rational Ï-conjugation design, researchers can systematically enhance fluorophore brightness while maintaining the deep-tissue penetration advantages of the NIR-II window.
The ongoing development of these engineering strategies continues to push the boundaries of in vivo imaging capabilities. Recent innovations in direct NIR-II chemiluminescence platforms [57] and genetically engineered protein-fluorophore complexes [37] represent promising avenues that may circumvent traditional quantum yield limitations entirely. As these molecular engineering approaches mature, they will undoubtedly accelerate the clinical translation of NIR-II fluorescence imaging, ultimately enhancing diagnostic precision and therapeutic monitoring capabilities across a spectrum of disease states.
The experimental protocols and materials detailed herein provide a foundation for researchers to implement these brightness enhancement strategies in their own NIR-II fluorophore development workflows, contributing to the collective effort to overcome the quantum yield bottleneck in deep-tissue bioimaging.
Fluorescence imaging in the second near-infrared window (NIR-II, 1000â1700 nm) has emerged as a transformative modality for deep-tissue bioimaging, offering significant advantages over traditional visible (400â700 nm) and NIR-I (700â900 nm) imaging. The longer wavelengths in the NIR-II window experience reduced photon scattering, lower tissue autofluorescence, and decreased light absorption by chromogenic biomolecules, resulting in millimeter- to centimeter-level penetration depths with micrometer spatial resolution [58] [14] [51]. This exceptional performance has positioned NIR-II fluorescence imaging as an indispensable tool for preclinical research and clinical applications, particularly in cancer diagnosis, image-guided surgery, and therapeutic monitoring.
However, the benefits across the broad NIR-II spectrum are not uniform. The NIR-IIa (1000â1300 nm) and NIR-IIb (1500â1700 nm) sub-regions are particularly valuable for deep-tissue imaging due to an optimal trade-off between reduced scattering and water absorption [14] [32]. As water absorption peaks sharply around 1450 nm, imaging in the NIR-IIb window can leverage rising light absorption to attenuate scattering photons, thereby further suppressing background interference [36]. Consequently, developing bright fluorophores with emissions red-shifted beyond 1000 nm, particularly into the NIR-IIa and NIR-IIb regions, represents a critical frontier in advancing deep-tissue imaging capabilities. This application note details molecular design strategies, experimental protocols, and reagent solutions to achieve these red-shifted emissions for enhanced biomedical imaging.
The D-A-D molecular architecture, featuring an electron-accepting (A) core symmetrically flanked by electron-donating (D) units, provides a highly tunable platform for achieving NIR-II emission.
Cyanine dyes, characterized by a polymethine chain connecting two heterocyclic units, are another major class of NIR-II fluorophores.
Engineering fluorophores with aggregation-induced emission (AIE) properties is particularly effective for designing bright nanofluorophores. AIE luminogens (AIEgens) are non-emissive in solution but become highly fluorescent in their aggregated state, which is ideal for nanoparticle formulation. Molecular design for NIR-II AIEgens focuses on creating twisted D-A-D structures full of molecular rotors. The restriction of intramolecular motion in the aggregate state blocks non-radiative pathways, forcing the energy to be released radiatively, thereby enhancing fluorescence in the NIR-II window [36].
The following tables summarize the photophysical properties of state-of-the-art NIR-II fluorophores developed using the strategies above, providing a comparative overview of their performance.
Table 1: Photophysical Properties of Representative D-A-D NIR-II Fluorophores
| Molecule Name | Acceptor Core | Donor Unit | Absorption Peak (nm) | Emission Peak (nm) | Molar Extinction Coefficient (Mâ»Â¹cmâ»Â¹) | Reference |
|---|---|---|---|---|---|---|
| 2TT-(o)CB | BBTD | Triphenylamine-based alkylthiophene | 695 | ~1000 | 1.1 Ã 10â´ | [36] |
| 2MTT-(o)CB | BBTD | Methoxy-triphenylamine | 736 | ~1050 | 1.3 Ã 10â´ | [36] |
| 2MPT-(o)CB | BBTD | Methoxy-diphenylamine | 860 | ~1150 | 1.8 Ã 10â´ | [36] |
| 2FT-(o)CB | BBTD | Fluorene-based diamine | 829 | 1215 | 2.3 Ã 10â´ | [36] |
Table 2: Performance of Other NIR-II Fluorophore Classes
| Fluorophore Class/Name | Type | Absorption Peak (nm) | Emission Peak (nm) | Key Feature | Reference |
|---|---|---|---|---|---|
| Flav7 | Cyanine | N/A | >1000 | QY of 0.53 | [58] |
| HSA@CO-1080 | Protein-seeking cyanine | 1044 | 1079 | 22-fold fluorescence enhancement | [8] |
| NDP (Activated) | Naphthalene diimide | 1066 | 1138, 1316 | "Off-on-off" probe | [59] |
This protocol is essential for characterizing newly synthesized NIR-II fluorophores in solution [36] [59].
Most organic NIR-II dyes are hydrophobic and require nanoparticle formulation for biological applications [36] [59].
This protocol validates the deep-tissue imaging performance of the formulated NIR-II probes [36].
Table 3: Key Reagents for Developing and Testing Red-Shifted NIR-II Probes
| Reagent / Material | Function / Application | Examples / Notes |
|---|---|---|
| BBTD Acceptor Core | Strong electron-acceptor for D-A-D dyes; enables emission >1200 nm. | Central unit in 2FT-(o)CB and related molecules [36]. |
| Triphenylamine Donors | Electron-donating units with tunable strength via substituents. | Used in 2TT-(o)CB; methoxy groups enhance donating power [36]. |
| PSMA Polymer | Amphiphilic polymer for nanoprecipitation; provides colloidal stability. | Polystyrene-co-maleic anhydride; forms small, stable nanoparticles [36]. |
| F-127-D-Gal Polymer | Functionalized copolymer for active targeting and biocompatibility. | Pluronic F-127 conjugated with D-Galactose for liver-targeting [59]. |
| Human Serum Albumin (HSA) | Endogenous protein for creating biomimetic NIR-II fluorescent proteins. | Covalently binds protein-seeking dyes, enhancing brightness & biocompatibility [8]. |
| 1064 nm Laser | Optimal excitation source for long-wavelength NIR-II dyes. | Deeper penetration than 808 nm; reduces tissue autofluorescence [8] [59]. |
| InGaAs Camera | Essential detector for NIR-II light. | Requires cooling (e.g., liquid Nâ) for low-noise imaging in 1000-1700 nm range. |
The following diagram illustrates the integrated workflow for developing and applying red-shifted NIR-II fluorophores, from molecular design to biological validation.
Diagram 1: Integrated R&D Workflow for Red-Shifted NIR-II Fluorophores
The strategic red-shifting of emission wavelengths into the NIR-IIa and NIR-IIb windows is paramount for unlocking the full potential of deep-tissue fluorescence imaging. The molecular design strategies outlinedâprimarily focused on intensifying donor-acceptor interactions in D-A-D architectures, engineering cyanine dyes, and leveraging AIE effectsâprovide a robust roadmap for developing brighter, longer-wavelength fluorophores. The accompanying experimental protocols and essential reagent toolkit offer a practical foundation for researchers to synthesize, characterize, and validate new NIR-II probes. As these strategies continue to evolve, they will accelerate the translation of NIR-II fluorescence imaging from a powerful research tool into a clinical reality for sensitive disease diagnosis and precise image-guided interventions.
The efficacy of fluorescence imaging agents, particularly those operating in the second near-infrared window (NIR-II, 900-1700 nm), is fundamentally governed by their pharmacokinetic (PK) profile. Optimizing this profileâspecifically the balance between circulation time, target specificity, and renal clearanceâis paramount for achieving high signal-to-background ratios (SBR) in deep-tissue imaging. The NIR-II window offers significant advantages over traditional NIR-I imaging, including reduced photon scattering, lower tissue autofluorescence, and deeper tissue penetration, but these benefits can only be fully leveraged with contrast agents possessing tailored pharmacokinetic properties [31] [36]. Effective PK optimization enables precise delineation of biological structures and disease targets, such as tumors, with reported SBR values exceeding 100 at tissue depths of 4-6 mm [36].
The core challenge in molecular probe design lies in managing the interplay of three key parameters:
Achieving an optimal balance requires a deep understanding of how a molecule's physicochemical propertiesâsuch as size, hydrophilicity, and molecular geometryâinfluence its pharmacokinetic pathway. For instance, a highly hydrophilic agent may exhibit rapid renal clearance, reducing background signal but potentially limiting its target accumulation time. Conversely, a more lipophilic agent may have a longer circulation time but suffer from high non-specific uptake in off-target tissues, leading to increased background noise [63].
Objective: To quantitatively determine the renal clearance of a NIR-II fluorescent agent in a pre-clinical murine model, providing a critical parameter for understanding its pharmacokinetic profile.
Principle: Renal clearance (CLR) is calculated by measuring the amount of fluorescent agent excreted in the urine over a specific interval and relating it to the plasma concentration-time profile. The most accurate method uses the area under the curve (AUC) to avoid assumptions about linear plasma decay [64].
Materials:
Procedure:
Diagram 1: Renal Clearance Determination Workflow
Objective: To characterize the comprehensive PK profile and tissue biodistribution of a NIR-II contrast agent, enabling the evaluation of its circulation time and target specificity.
Principle: This protocol involves administering the agent, collecting blood for PK analysis over time, and subsequently harvesting organs to measure agent accumulation, thereby quantifying its distribution and clearance pathways [60].
Materials:
Procedure:
Diagram 2: Key Pharmacokinetic Pathways
| Reagent / Material | Function / Application in PK Studies | Key Characteristics & Examples |
|---|---|---|
| NIR-II Fluorophores | Core imaging agent for tracking PK/BD. | Small Organic Dyes (e.g., IRDye 680LT): Often show renal clearance. Targeted Agents (e.g., ABY-029): Affibody- or antibody-conjugated for specificity. D-A-D Molecules (e.g., 2FT-oCB): Engineered for bright emission >1200 nm [60] [36]. |
| Calibration Matrix | Creating tissue-specific calibration curves for accurate concentration recovery. | Homogenates of relevant tissues (liver, tumor, spleen, etc.) and whole blood. Critical for accounting for tissue optical properties [60]. |
| Borosilicate Capillaries | Standardized containers for fluorescence measurement of homogenates. | Provide a standardized path length for excitation light, ensuring consistent and quantitative fluorescence readings from small sample volumes [60]. |
| Para-Aminohippuric Acid (PAH) | Reference compound for measuring effective renal plasma flow. | Actively secreted by renal tubules; used to quantify renal secretory function and blood flow [64]. |
| Inulin / Iothalamate | Reference marker for measuring Glomerular Filtration Rate (GFR). | Inert polysaccharide filtered by glomeruli but not secreted or reabsorbed; gold standard for GFR measurement [64]. |
| Imaging Agent / Window | Emission Peak (nm) | Key Pharmacokinetic & Imaging Properties | Primary Application Rationale |
|---|---|---|---|
| IRDye 680LT | ~680 nm (NIR-I) / Can be used in NIR-II with filters | Pharmacokinetically similar, untargeted control; rapid renal clearance. | Paired-agent imaging (as untargeted control); baseline for vascular imaging [60]. |
| ABY-029 | ~800 nm (NIR-I/II border) | EGFR-targeted Affibody molecule; pharmacokinetics influenced by targeting. | Fluorescence-guided surgery; demonstrates active targeting vs. passive distribution [60]. |
| 2FT-oCB | 1215 nm | Engineered D-A-D molecule; ultralong molecular packing (>8 Ã ) inhibits Ï-Ï stacking, enhancing brightness [36]. | Deep-tissue imaging with high SBR; emission matched to low-scattering windows. |
| SIDAG | ~800 nm (NIR-I/II border) | Hydrophilic derivative of ICG; shifted PK toward renal elimination; increased acute tolerance [63]. | Improved tumor contrast over ICG; model for engineering hydrophilicity to modulate clearance. |
| NIR-IIb Window (1500-1700 nm) | N/A | Greatly reduced photon scattering and negligible autofluorescence. | High-contrast imaging of deep structures; requires bright agents with emission in this range [31] [36]. |
The following reagents and tools are indispensable for conducting the experiments outlined in these protocols.
| Category | Item | Specific Function |
|---|---|---|
| Contrast Agents | IRDye 680LT (carboxylate) | Untargeted, renally-cleared reference agent for PK studies [60]. |
| ABY-029 (Anti-EGFR Affibody) | Targeted agent for studying the impact of receptor binding on PK and distribution [60]. | |
| Engineered D-A-D Fluorophores (e.g., 2FT-oCB) | Bright, long-wavelength emitters for high-contrast NIR-IIb imaging [36]. | |
| Analytical Tools | Tissue Homogenization Buffer | Medium for creating uniform tissue suspensions for quantitative fluorescence analysis. |
| Tissue-Specific Calibration Curves | Essential standard curves for converting fluorescence signal to accurate concentration in different tissue matrices [60]. | |
| Reference Standards | Para-Aminohippuric Acid (PAH) | Gold standard for measuring renal plasma flow and studying active secretion [64]. |
| Inulin or Iothalamate | Gold standard for measuring Glomerular Filtration Rate (GFR) [64]. |
The delivery of therapeutic agents to solid tumors is a formidable challenge in oncology, primarily due to a series of biological barriers that significantly limit drug efficacy. The Enhanced Permeability and Retention (EPR) effect, first discovered by Maeda and colleagues, has long been considered the cornerstone of nanoparticle-based cancer therapy [65]. This pathophysiological phenomenon enables macromolecules and nanoparticles to selectively accumulate in tumor tissue due to leaky vasculature and defective lymphatic drainage [65] [66]. However, the journey from intravenous administration to intracellular drug release is fraught with obstacles, including rapid clearance by the Mononuclear Phagocyte System (MPS), heterogeneous EPR effects across tumor types, and poor penetration through the dense tumor microenvironment [65] [67] [68].
The clinical translation of nanomedicines has been disappointing despite promising preclinical results. With over 50,000 manuscripts published on nanoparticles for cancer by the end of 2021, only 15 antitumor nanomedicines have gained clinical approval worldwide [67]. This stark contrast highlights the critical need to understand and navigate the biological barriers that impede effective tumor targeting. This Application Note provides detailed protocols and strategies to overcome these barriers, with particular emphasis on leveraging NIR-II fluorescence imaging to monitor and validate drug delivery processes in real-time.
The MPS, comprising macrophages in the liver, spleen, and bone marrow, represents the first major barrier to nanoparticle delivery. Following intravenous administration, nanotherapeutics are rapidly recognized and cleared by the MPS, significantly reducing their circulation time and tumor accumulation [67] [68]. Protein corona formation â the adsorption of plasma proteins onto nanoparticle surfaces â further enhances MPS recognition and uptake [68].
The EPR effect allows nanoparticles ranging from 40-400 nm to extravasate through the defective tumor vasculature with fenestrations of 100-780 nm and accumulate in the tumor interstitium due to impaired lymphatic drainage [65] [66]. This phenomenon can lead to drug concentrations in tumors that are 10-100 times higher than those achieved with conventional free drug administration [65].
However, the EPR effect exhibits significant heterogeneity across different tumor types, patients, and even regions within the same tumor [65] [67]. This variability stems from differences in vascular density, endothelial cell receptor expression, vascular maturation, extracellular matrix composition, tumor cell density, hypoxia, and interstitial fluid pressure [65]. Pancreatic ductal adenocarcinoma, for instance, presents particularly strong stromal barriers that limit EPR-based targeting [65].
The tumor microenvironment presents multiple additional barriers including:
Table 1: Key Biological Barriers and Their Impact on Nanodrug Delivery
| Biological Barrier | Main Challenges | Consequences for Nanodrugs |
|---|---|---|
| MPS Clearance | Opsonization, phagocytic uptake | Short circulation half-life, reduced bioavailability |
| EPR Heterogeneity | Inter- and intra-tumoral variability in vascular permeability | Inconsistent tumor accumulation, median of only 0.7% injected dose reaches tumors [65] |
| Tumor Microenvironment | High IFP, dense ECM, acidic pH | Limited penetration and distribution, reduced cellular uptake |
| Cellular Uptake | Endosomal entrapment, inefficient drug release | Limited drug bioavailability intracellularly |
Second near-infrared window (NIR-II, 1000-1700 nm) fluorescence imaging has emerged as a powerful tool for monitoring nanoparticle distribution and tumor accumulation in real-time. This technology provides significant advantages over traditional imaging modalities for validating navigation strategies across biological barriers.
NIR-II fluorescence imaging offers:
The improved performance in the NIR-II window stems from reduced scattering and the constructive role of moderate water absorption in suppressing multiply scattered photons that contribute to background noise [5]. This absorption preferentially depletes background signals while preserving ballistic photon signals, enhancing image contrast [5].
Recent research has identified sub-windows within the NIR-II spectrum with optimized imaging characteristics:
Table 2: NIR-II Fluorescent Probes for Tracking Tumor Targeting
| Probe Type | Emission Range (nm) | Advantages | Limitations |
|---|---|---|---|
| Indocyanine Green (ICG) | 800-1200 | FDA-approved, rapid liver clearance (within 10 h) [58] | Lack of tumor targeting, not activatable |
| Organic Cyanines | 1000-1500 | Tunable emission, commercial availability | Compromised stability and QY with extended polymethine chain [58] |
| Quantum Dots (PbS/CdS) | 1000-1500+ | High brightness, tunable emission, narrow bandwidth [1] [5] | Potential toxicity concerns |
| Lanthanide Nanoparticles | 1000-2000 | Narrow emission bands, long luminescence lifetimes, minimal photobleaching [1] | Complex synthesis |
| D-A-D Dyes | 1000-1300 | High stability, large Stokes shift [1] | Requires conjugation for optimal performance |
Purpose: To monitor nanoparticle circulation, tumor accumulation, and clearance in real-time using NIR-II fluorescence imaging.
Materials:
Procedure:
Notes: For optimal SBR, utilize the "off-peak" imaging method with 1400 nm long-pass detection, which can exceed the performance of strict NIR-IIb imaging [5]. The surgical window for optimal tumor resection guidance typically occurs when SBR reaches ~11.5 [58].
Purpose: To augment nanoparticle tumor accumulation by modulating tumor vasculature and microenvironment.
Materials:
Procedure:
Notes: The combination of pharmacological and physical methods often yields synergistic effects. Monitor for potential increases in systemic toxicity with vasoactive agents.
Purpose: To simultaneously track multiple nanoparticle formulations or target different cellular populations to assess tumor heterogeneity.
Materials:
Procedure:
Notes: Lifetime-based multiplexing can complement spectral multiplexing for more complex experimental designs [1].
Table 3: Essential Research Reagents for Barrier Navigation Studies
| Reagent Category | Specific Examples | Key Functions | Applications |
|---|---|---|---|
| NIR-II Fluorophores | ICG, Flav7, IR-26, CH-4T [58] [1] | Real-time tracking, surgical guidance | Pharmacokinetic studies, tumor margin delineation |
| Nanoparticle Platforms | PEGylated liposomes, PLGA nanoparticles, dendrimers, gold nanoparticles [65] [66] | Drug encapsulation, prolonged circulation, EPR exploitation | Therapeutic delivery, combination therapies |
| EPR Enhancers | Bradykinin, nitric oxide donors, prostaglandins [65] [66] | Vasodilation, increased vascular permeability | Augmenting tumor accumulation |
| TME Modulators | ECM-degrading enzymes, TGF-β inhibitors [65] [67] | Reduced interstitial pressure, decreased matrix density | Improving tumor penetration |
| Surface Modifiers | PEG, CD47-derived peptides, "Self" peptides [68] | Reduced opsonization, MPS evasion | Prolonging circulation half-life |
The following diagram illustrates the sequential biological barriers and strategic approaches to overcome them:
Successful tumor targeting requires addressing the entire Circulation, Accumulation, Penetration, Internalization, and Release (CAPIR) cascade [65]. The complementary integration of passive targeting (EPR effect) with active targeting strategies is essential for improving therapeutic outcomes. This includes:
Navigating biological barriers requires a multifaceted approach that addresses the sequential challenges of MPS clearance, EPR heterogeneity, and tumor penetration. The integration of NIR-II fluorescence imaging provides an invaluable tool for developing and validating these strategies in real-time. By combining optimized nanoparticle design with EPR enhancement techniques and leveraging advanced imaging technologies, researchers can significantly improve the tumor targeting efficiency of nanotherapeutics.
The future of barrier navigation lies in personalized approaches that account for inter- and intra-tumoral heterogeneity, with treatment strategies tailored to individual patient characteristics and real-time monitoring of therapeutic response. As NIR-III imaging (2080-2340 nm) emerges, further improvements in imaging fidelity will enable even more precise evaluation of nanoparticle trafficking and tumor targeting [5].
Within the broader thesis on Near-Infrared Window II (NIR-II, 1000â1700 nm) fluorescence imaging for deep tissue penetration research, this document provides detailed application notes and protocols for validating probe and system efficacy across the research pipeline. Compared to conventional visible (400â760 nm) and NIR-I (760â900 nm) imaging, NIR-II bioimaging offers superior spatial-temporal resolution and deeper tissue penetration (up to several centimeters) due to reduced photon scattering, low tissue autofluorescence, and minimal light absorption by biological components [33] [5]. These advantages make it an indispensable tool for fundamental and preclinical research, enabling real-time, non-invasive, and multi-dimensional investigations of in vivo biological processes [33] [4]. The following sections outline standardized protocols and key data for efficacy validation from in vitro assays through clinical patient studies.
In vitro validation establishes the foundational optical and biochemical properties of NIR-II probes, serving as a critical prerequisite for in vivo applications.
This protocol assesses the activity and selectivity of bioconjugates between NIR-II fluorophores and targeting ligands (e.g., antibodies, proteins) [33].
Key Reagents:
Methodology:
This procedure determines the blood circulation half-life and organ-specific accumulation of NIR-II probes in small animals [33].
Key Reagents:
Methodology:
Table 1: Key quantitative parameters from in vitro NIR-II assays.
| Parameter | Experimental Result | Implication |
|---|---|---|
| Microarray P/N Ratio | Increased from 1.1 (pre-DGU) to 5.3 (post-DGU) for Erb@IR-FGP [33] | DGU purification significantly improves signal-to-background ratio and probe selectivity. |
| Blood Half-Life | ~24 minutes for IR-BGP6 [33] | Indicates rapid clearance, which can be advantageous for reducing background signal. |
| Renal Excretion | ~91% of IR-BGP6 excreted in urine within first 10 hours [33] | Confirms primary clearance pathway, important for biocompatibility and toxicity assessment. |
| Spatial Resolution | ~125 µm achieved by preclinical IR VIVO system [26] | Benchmarks the high-resolution capability of NIR-II imaging systems. |
| Detection Sensitivity | ICG concentrations as low as 30 nM (NIR-I) and 300 nM (NIR-II) [26] | Highlights the sensitivity limits for fluorophore detection. |
In vivo protocols translate validated probes to living systems for functional and anatomical imaging.
This protocol images vascular structures and tumor margins, leveraging the deep penetration of NIR-II light [69] [5].
Key Reagents:
Methodology:
This protocol acquires three-dimensional information on tumor vasculature using a binocular stereo vision principle [69].
Key Reagents:
Methodology:
For studies requiring long-term observation (months), genetically encoded probes are ideal [70].
Key Reagents:
Methodology:
Table 2: Key quantitative parameters from in vivo NIR-II imaging.
| Parameter | Experimental Result | Implication |
|---|---|---|
| Tissue Penetration Depth | Up to ~10 mm in brain tissue using NIR-IIb/IIx windows [5] | Enables visualization of deep-seated structures and tumors. |
| Temporal Resolution | Real-time imaging capability (frames per second) [33] | Allows for monitoring of dynamic processes like blood flow. |
| Spatial Resolution | ~1.3 mm depth in mouse brain microscopy [5] | Represents the deepest in vivo NIR-II fluorescence microscopy in mice brain. |
| Signal-to-Noise Ratio (SNR) | ~5x higher in NIR-II bioluminescence vs NIR-II fluorescence [71] | Elimination of excitation light leads to a dramatically cleaner signal. |
| Tumor-to-Normal Tissue (T/N) Ratio | Up to 83.4 for ATP-responsive metastases tracing [71] | Enables high-contrast detection of malignant tissue against healthy background. |
Translating NIR-II imaging to clinical settings requires validation against current standards of care in patient studies.
This protocol is for the ex vivo assessment of freshly excised tumor specimens to rapidly identify positive margins, a critical step in fluorescence-guided surgery (FGS) [26] [72].
Key Reagents:
Methodology:
This procedure tests the feasibility of NIR-II imaging in a real-world surgical environment without the need for room blackout [26].
Key Reagents:
Methodology:
Table 3: Key quantitative parameters from clinical NIR-II imaging studies.
| Parameter | Experimental Result | Implication |
|---|---|---|
| Spatial Resolution (Clinical System) | ~250 µm resolved by LightIR system [26] | Sufficient resolution for identifying fine surgical margins. |
| Depth Penetration (Clinical) | ICG detected to depths â¥4 mm [26] | Allows visualization of subsurface tumors not visible to the naked eye. |
| Contrast-to-Noise (aCNR) | Varies by cancer type; SWIR potential in PSCC, NIR better in HNSCC [72] | Highlights that optimal imaging window can be cancer-specific. |
| Ambient Light Operation | LightIR system successfully visualized subsurface targets without blackout [26] | Critical for practical integration into surgical workflow. |
Table 4: Key reagents and materials for NIR-II fluorescence imaging experiments.
| Item Category | Specific Examples | Function & Application |
|---|---|---|
| NIR-II Fluorophores | Organic small molecules (IR-FGP, IR-BGP6, IR-1048), Polymer dots (IR-TPE Pdots), Inorganic QDs (PbS/CdS), Rare-earth nanoparticles, Genetic encoded probes (iRFPs) [33] [70] [69] | Serve as contrast agents. Choice depends on required brightness, biocompatibility, and targeting needs. |
| Targeting Ligands | Antibodies (e.g., Cetuximab, anti-PD-L1), Peptides, Proteins | Conjugated to fluorophores to enable specific binding to cellular receptors (e.g., EGFR) on target tissues [33] [72]. |
| Imaging Systems | Preclinical (IR VIVO), Clinical/Intraoperative (LightIR), Custom-built setups with InGaAs cameras [26] [69] | Detect NIR-II fluorescence. Systems vary in portability, sensitivity, and suitability for in vivo vs. clinical use. |
| Key Optical Components | Laser Diodes (e.g., 793 nm, 808 nm), Long-Pass Filters (e.g., LP1100, LP1250, LP1400), InGaAs Cameras [26] [69] | Components for excitation (lasers), emission filtering (to block excitation light), and detection (cameras). |
| Animal Models | Wild-type mice, Transgenic models (e.g., iRFP713flox/flox), Tumor-bearing models (e.g., MC38, MTCQ1) [33] [70] [69] | Provide the in vivo context for evaluating probe performance and imaging biological processes. |
The following diagram outlines the core logical pathway for validating NIR-II imaging efficacy from basic research to clinical application.
This diagram visually compares the key performance metrics between NIR-I and NIR-II imaging windows, based on data from preclinical and clinical studies.
Within the broader thesis on NIR-II fluorescence imaging for deep-tissue penetration research, this document establishes detailed application notes and protocols for the quantitative assessment of multi-organ dysfunction and therapeutic response. Traditional methods for monitoring organ function, such as pathological analysis and the measurement of biomarkers like creatinine, are often invasive, slow, and lack the specificity for early detection of dysfunction [73] [74]. Consequently, there is a pressing need for non-invasive, real-time technologies that enable the precise longitudinal monitoring of multiple organs simultaneously.
Fluorescence imaging in the second near-infrared window (NIR-II, 1000â1700 nm) has emerged as a powerful solution to this challenge. Due to reduced light scattering, absorption, and negligible tissue autofluorescence in this spectral region, NIR-II imaging offers superior tissue penetration depth (up to several centimeters) and a high spatiotemporal resolution compared to imaging in the visible or first near-infrared (NIR-I) windows [52] [75] [76]. These advantages make it an ideal modality for the real-time, quantitative assessment of physiological processes and drug-induced pathologies across deep tissues [77]. This protocol details the application of a multiplexed NIR-II fluorescent bioimaging method for the simultaneous evaluation of liver, kidney, and gastrointestinal function, providing a novel approach for enhancing the evaluation of drug side effects and therapeutic interventions [73].
The quantitative assessment of multi-organ dysfunction requires high-fidelity data, which is enabled by the unique physical properties of the NIR-II window.
The following table summarizes the quantitative benefits of NIR-II imaging compared to traditional optical windows.
Table 1: Comparative Advantages of NIR-II Imaging for Quantitative Assessment
| Characteristic | Visible/NIR-I Imaging | NIR-II Imaging | Impact on Quantitative Assessment |
|---|---|---|---|
| Tissue Penetration | Limited (1-3 mm) [76] | Deep (~1 cm) [52] | Enables imaging of deep organs like liver and kidneys. |
| Spatial Resolution | Lower (blurred by scattering) | Higher (< 10 µm through skull) [77] | Allows precise localization of signal to specific organ structures. |
| Signal-to-Noise Ratio | Lower due to autofluorescence | Significantly higher [52] [77] | Improves accuracy of signal quantification and detection sensitivity. |
| Tissue Scattering | High | Reduced (decreases with λ^α) [52] | Minimizes image distortion, leading to more accurate measurements. |
The successful implementation of this protocol relies on a suite of specialized reagents and equipment. The core components are NIR-II fluorescent probes, which can be broadly categorized into inorganic and organic types, each with distinct properties suitable for different applications.
Table 2: Essential Research Reagents and Materials for NIR-II Imaging
| Item Category | Specific Examples | Function and Key Characteristics |
|---|---|---|
| NIR-II Fluorophores | Hemicyanine Dyes (HDs) [73], Organic Small Molecules (CH1055, Y6CT) [75] [78], Lanthanide-Doped Nanoparticles (DCNPs) [76], Quantum Dots (AgâS QDs) [76] | Function as contrast agents. HDs and small molecules offer synthetic tunability and better biocompatibility; inorganic probes like DCNPs and QDs offer high photostability and quantum yield. |
| Targeting Moieties | Antibodies, Peptides, Ligands | Conjugated to fluorophores to enable specific binding to cellular biomarkers or organ structures, enhancing localization and signal specificity. |
| Molecular Scaffolds | A-D-A type conjugated molecules (e.g., Y6CT) [78], Donor-Acceptor-Donor (D-A-D) structures [75] | Form the core of organic NIR-II fluorophores, allowing for precise tuning of absorption/emission wavelengths and brightness through chemical synthesis. |
| Nanoparticle Matrix | DSPE-mPEG2000 [78] | Used to encapsulate hydrophobic organic dyes into nanoparticles (e.g., Y6CT-NPs), improving aqueous solubility, biocompatibility, and circulation time. |
| Activation Probes | Activatable Probes (e.g., DPM-HD3-CO) [79] | "Turn-on" in response to specific biomarkers (e.g., carbon monoxide), providing high specificity for detecting pathological events. |
| Imaging Instrumentation | InGaAs Camera [52] [75] | Essential detector for capturing light in the 900-1700 nm range. Requires cooling to reduce thermal noise for high-sensitivity NIR-II imaging. |
| Excitation Sources | Lasers (808 nm), White-light sources (Laparoscopic LED) [78] | Activate the NIR-II fluorophores. Lasers are common, but clinical white-light sources offer homogeneous illumination and safety for translation. |
This section provides a detailed, step-by-step methodology for applying multiplexed NIR-II imaging to assess drug-induced multi-organ dysfunction, using cisplatin and aristolochic acids as model compounds [73].
Objective: To simultaneously monitor the real-time dysfunction of the liver, kidneys, and gastrointestinal tract in a live animal model following drug administration.
Materials:
Procedure:
Validation: Correlate the in vivo imaging findings with gold-standard ex vivo analyses, such as histopathological examination (H&E staining) and transcriptomics of the extracted organs [73].
Objective: To non-invasively monitor the response of a diseased organ to a therapeutic intervention using NIR-II imaging.
Materials:
Procedure:
The quantitative data derived from NIR-II imaging experiments should be systematically organized to facilitate comparison and interpretation. The following table provides a template for summarizing key pharmacokinetic parameters from a multi-organ dysfunction study.
Table 3: Quantitative Pharmacokinetic Parameters from NIR-II Imaging of Organ Function
| Organ | Experimental Group | Time-to-Peak (TTP, min) | Signal Half-Life (tâ/â, min) | AUC (a.u.) | Interpretation |
|---|---|---|---|---|---|
| Liver | Control | 5.2 ± 0.8 | 45.3 ± 5.1 | 10500 ± 850 | Normal uptake and metabolism |
| Cisplatin-Treated | 8.7 ± 1.1 | 92.5 ± 9.8 | 21800 ± 1200 | Impaired metabolic function | |
| Kidney | Control | 3.5 ± 0.5 | 25.1 ± 3.2 | 5800 ± 450 | Normal filtration and clearance |
| Cisplatin-Treated | 6.8 ± 0.9 | 65.4 ± 7.2 | 15200 ± 1100 | Acute kidney injury | |
| Stomach | Control | - | - | - | Low background signal |
| AA-I-Treated | 15.2 ± 2.5 | >120 | 9500 ± 800 | Gastric emptying disorder [73] |
The following diagrams illustrate the core experimental workflow and molecular design principles underlying the NIR-II imaging protocol.
Diagram 1: Experimental Workflow for NIR-II Organ Dysfunction Assessment
Diagram 2: NIR-II Imaging Principle and Advantage
Fluorescence imaging in the second near-infrared window (NIR-II, 1000â1700 nm) has emerged as a transformative modality for in vivo biological imaging, addressing critical limitations of both conventional clinical imaging systems and earlier optical techniques. This application note provides a structured comparison of NIR-II fluorescence imaging against established clinical modalities (MRI, CT, PET) and first near-infrared window (NIR-I) fluorescence imaging, with emphasis on their respective capabilities for deep-tissue imaging research. The content is framed within the broader thesis that NIR-II imaging offers superior performance for specific preclinical and potential clinical applications requiring high spatial resolution, deep penetration, and real-time visualization capabilities.
The fundamental advantage of NIR-II imaging stems from reduced photon scattering and minimal tissue autofluorescence at longer wavelengths, which collectively enhance penetration depth and improve signal-to-background ratios (SBR). Recent research has further refined our understanding of optimal sub-windows within the NIR-II spectrum, including regions with higher water absorption (1400-1500 nm and 1880-2080 nm) that paradoxically improve imaging contrast by preferentially attenuating scattered background photons [24]. For researchers and drug development professionals, understanding these comparative advantages is crucial for selecting appropriate imaging methodologies for specific experimental and diagnostic challenges.
Table 1: Head-to-Head Comparison of Key Imaging Modalities
| Parameter | NIR-II Fluorescence | NIR-I Fluorescence | MRI | CT | PET |
|---|---|---|---|---|---|
| Spatial Resolution | â¼Âµm range (â¼25-50 µm in vivo) [32] | â¼Âµm range (limited by scattering) [32] | â¼Âµm (preclinical) to mm (clinical) [80] | â¼Âµm (preclinical) to mm (clinical) [80] | 1-2 mm (preclinical) to 5-7 mm (clinical) [80] |
| Penetration Depth | Up to several centimeters [32] [34] | <1 cm [32] | Unlimited | Unlimited | Unlimited |
| Temporal Resolution | Excellent (seconds to minutes) [80] | Excellent (seconds to minutes) [80] | Minutes to hours [32] | Excellent (seconds) [32] | Minutes to hours [32] |
| Sensitivity | High (nM-pM) [80] | High (nM-pM) [80] | Low (µM-mM) [80] | Low | High (pM-fM) [80] |
| Ionizing Radiation | No [32] | No [32] | No | Yes [32] | Yes [32] |
| Quantitative Capability | Moderate | Moderate | Excellent (functional) | Excellent (anatomical) | Excellent (metabolic) |
| Molecular Specificity | Excellent [32] | Excellent [32] | Good with targeted agents | Poor | Excellent [32] |
| Real-time Imaging | Yes [34] | Yes [34] | No | Yes | No |
| Cost & Accessibility | Moderate | Low | High [32] | Moderate [32] | High [32] |
| Contrast Agent Requirements | Required (fluorophores) [32] | Required (fluorophores) [32] | Optional (gadolinium, iron oxide) | Iodine-based required | Radiolabeled tracers required |
Table 2: NIR-II Sub-window Characteristics for Deep-Tissue Imaging
| Spectral Window | Wavelength Range (nm) | Key Characteristics | Optimal Applications |
|---|---|---|---|
| NIR-IIa | 1300-1400 [24] | Reduced scattering, low water absorption | Deep-tissue vascular imaging [24] |
| NIR-IIx | 1400-1500 [24] | Moderate water absorption improves contrast | High-contrast imaging in aqueous tissues [24] |
| NIR-IIb | 1500-1700 [24] | Further reduced scattering | Superficial high-resolution imaging |
| NIR-IIc | 1700-1880 [24] | Low scattering, emerging applications | Imaging in adipose-rich environments [24] |
| Extended NIR-II | 1880-2080 [24] | High water absorption for maximal contrast | Ultra-high contrast for complex backgrounds [24] |
NIR-II Fluorescence Imaging excels in high-resolution real-time visualization of molecular targets with exceptional sensitivity. The technique leverages reduced scattering of longer wavelengths, enabling deeper penetration (up to several centimeters) compared to NIR-I while maintaining micron-scale resolution [32] [34]. Recent evidence demonstrates that previously avoided spectral regions near water absorption peaks (â¼1450 nm and â¼1930 nm) actually provide superior contrast by preferentially attenuating longer-path scattered photons [24]. This paradigm shift enables high-contrast imaging in the 1880-2080 nm window when sufficiently bright fluorophores are employed.
Clinical Modalities (MRI, CT, PET) provide whole-body capabilities with complementary strengths. MRI offers excellent soft-tissue contrast without radiation exposure but suffers from low sensitivity and slow imaging times [32]. CT provides exceptional bone visualization and rapid acquisition but involves ionizing radiation and poor soft-tissue contrast [80]. PET delivers unparalleled sensitivity for metabolic processes but requires radioactive tracers and offers limited spatial resolution [80] [32].
NIR-I Fluorescence Imaging (700-900 nm) maintains advantages in instrument availability and established dye chemistry but is fundamentally limited by significant tissue scattering and autofluorescence, constraining penetration depth to <1 cm with compromised resolution in deep tissues [32].
Principle: Exploits the high water absorption around 1930 nm to preferentially attenuate scattered photons, dramatically improving signal-to-background ratio in vascular imaging [24].
Materials:
Procedure:
Key Considerations: The bright fluorescence of PbS/CdS QDs must overcome significant water absorption attenuation. Optimal results require emission matching the 1880-2080 nm window [24].
Principle: Creates biomimetic NIR-II fluorescent proteins through covalent binding of protein-seeking dyes to serum albumin, enhancing brightness and photostability for prolonged imaging sessions [8].
Materials:
Procedure:
Key Considerations: The hydrophobic cavity of HSA restricts molecular motion of covalently bound dyes, reducing non-radiative decay and enhancing fluorescence intensity up to 22-fold [8].
Figure 1: Strategic approaches for developing high-performance NIR-II fluorophores through molecular engineering and brightness enhancement techniques [81] [32].
Figure 2: Mechanism demonstrating how water absorption in specific NIR-II windows (1400-1500 nm and 1880-2080 nm) preferentially attenuates scattered background photons, enhancing image contrast [24].
Table 3: Key Research Reagents for NIR-II Fluorescence Imaging
| Reagent Category | Specific Examples | Function & Characteristics | Application Context |
|---|---|---|---|
| Inorganic Nanoparticles | PbS/CdS QDs [24], AgâS QDs [82] | Bright, photostable emitters with tunable NIR-II emission | Deep-tissue imaging (1880-2080 nm window) [24] |
| Organic Small Molecules | D-A-D fluorophores [32], Cyanines [32] | Well-defined structures, favorable biocompatibility | Molecular-targeted imaging, rapid clearance studies |
| Biomimetic Systems | HSA@CO-1080 [8], Protein-seeking dyes | Enhanced brightness via restricted molecular motion | Long-circulating vascular imaging, lymphography |
| Surface Modification Agents | PEG derivatives, targeting peptides | Improve biocompatibility and target specificity | Tumor-specific imaging, pharmacokinetic optimization |
| Detection Systems | InGaAs cameras [34], Spectral filters | Capture NIR-II photons with high sensitivity | System setup for various NIR-II sub-windows |
NIR-II fluorescence imaging represents a powerful modality that addresses fundamental limitations of both conventional clinical imaging and earlier optical techniques. Its unique combination of high spatial resolution, deep-tissue penetration, and real-time imaging capability positions it as an invaluable tool for preclinical research and emerging clinical applications.
The strategic selection of NIR-II sub-windows, particularly those leveraging water absorption characteristics (1400-1500 nm and 1880-2080 nm), enables unprecedented contrast for visualizing complex biological structures and processes. Continued development of brighter fluorophores through molecular engineering strategies like intramolecular covalent bond locking [81] and biomimetic protein-dye complexes [8] will further expand the capabilities of NIR-II imaging.
For researchers and drug development professionals, integrating NIR-II fluorescence imaging with established modalities like MRI and PET offers a multimodal approach that leverages the respective strengths of each technology. This synergistic strategy promises to accelerate both fundamental biological discovery and the translation of novel diagnostic and therapeutic agents.
Fluorescence imaging in the second near-infrared window (NIR-II, 1000-1700 nm) represents a paradigm shift in biomedical imaging, offering unprecedented capabilities for deep-tissue visualization with high spatial resolution. This technology leverages the favorable optical properties of biological tissues in this spectral range, including reduced photon scattering, lower tissue autofluorescence, and minimized light absorption by key tissue components such as hemoglobin and water [31] [14]. These intrinsic advantages enable NIR-II fluorescence imaging to achieve millimeter-to-centimeter penetration depths with micron-scale resolution, far surpassing the capabilities of traditional NIR-I imaging (700-900 nm) [31] [58]. The burgeoning applications span from intraoperative surgical guidance and tumor margin delineation to vascular imaging and therapeutic monitoring, positioning NIR-II technology at the forefront of next-generation diagnostic and therapeutic platforms [26] [58].
Despite these remarkable advantages, the clinical translation of NIR-II imaging agents faces significant challenges centered on biocompatibility, toxicity, and regulatory compliance. The limited pool of effective NIR-II fluorophores, coupled with concerns regarding potential long-term accumulation and immunogenic responses, presents substantial barriers to clinical adoption [26] [34]. Many promising NIR-II probes exhibit suboptimal pharmacokinetics, inadequate target-to-background ratios, and insufficient brightness in the longer-wavelength NIR-II sub-windows (NIR-IIa, 1300-1400 nm; NIR-IIb, 1500-1700 nm) where imaging performance is optimal [14] [36]. Furthermore, the high cost of specialized imaging equipment, particularly InGaAs cameras required for NIR-II detection, has limited accessibility and widespread clinical application [31] [26]. This application note provides a comprehensive framework for addressing these translational challenges, with structured protocols for evaluating the safety and efficacy of NIR-II contrast agents within a regulatory context.
The biological interactions of NIR-II imaging agents vary significantly based on their material composition, surface chemistry, and structural characteristics. A thorough understanding of these properties is essential for rational probe design and clinical development.
Table 1: Comparative Analysis of NIR-II Fluorophore Classes and Their Properties
| Fluorophore Class | Examples | Key Advantages | Biocompatibility Concerns | Current Status |
|---|---|---|---|---|
| Organic Small Molecules | ICG, Cyanine dyes, BODIPY, Donor-Acceptor-Donor (D-A-D) dyes | Tunable pharmacokinetics, renal clearance, well-defined chemical structures [58] [83] | Moderate quantum yield, potential photobleaching, variable stability in biological environments [14] [58] | ICG FDA-approved for NIR-I; NIR-II analogs in preclinical development [58] |
| Inorganic Nanomaterials | Quantum Dots (QDs), Rare-Earth-Doped Nanoparticles (RENPs), Single-Walled Carbon Nanotubes (SWCNTs) | High photostability, strong fluorescence intensity, tunable optical properties [31] [34] | Long-term accumulation potential, metal ion leakage, undefined clearance pathways [26] [34] | Primarily preclinical; significant toxicity concerns require resolution |
| Hybrid Materials | Dye-loaded nanoparticles, polymer-fluorophore conjugates | Enhanced brightness, multimodal capabilities, tailored pharmacokinetics [34] | Complex manufacturing, batch-to-batch variability, uncertain regulatory pathway | Early-stage development with promising preclinical results |
The biological safety of NIR-II imaging agents is governed by a set of Critical Quality Attributes (CQAs) that must be carefully evaluated:
Size and Surface Characteristics: Nanoparticle size directly influences biodistribution, cellular uptake, and clearance pathways. Materials smaller than 5-6 nm typically undergo renal clearance, while larger particles may accumulate in the reticuloendothelial system (liver and spleen) [14]. Surface charge (zeta potential) affects protein corona formation and subsequent immune recognition, with highly positive or negative surfaces generally exhibiting increased cytotoxicity [34].
Chemical Stability and Degradation Products: Fluorophores must demonstrate stability in physiological environments to prevent release of potentially toxic components. For inorganic nanomaterials, this includes assessment of metal ion leaching under acidic conditions (e.g., lysosomal environments) [34]. Degradation pathways should be clearly elucidated, with toxicological profiling of all major degradation products.
Photophysical Properties and Light-Tissue Interactions: High quantum yield enables lower dosing for effective imaging, directly reducing potential toxicity [14]. Additionally, photostability minimizes generation of reactive oxygen species (ROS) that could cause tissue damage, particularly important for applications requiring prolonged illumination [36].
Robust assessment protocols are essential for generating the comprehensive safety and efficacy data required for regulatory submissions.
Objective: To evaluate baseline cytotoxicity, cellular uptake, and inflammatory responses induced by NIR-II imaging agents.
Materials:
Methodology:
Data Analysis: Calculate IC50 values for cytotoxicity and no-observed-adverse-effect-level (NOAEL) concentrations. Correlation analysis between physicochemical properties (size, charge) and biological responses guides material optimization.
Objective: To assess systemic toxicity, organ accumulation, and clearance pathways in relevant animal models.
Materials:
Methodology:
Data Analysis: Establish maximum tolerated dose (MTD) and no-observed-adverse-effect-level (NOAEL). Calculate key pharmacokinetic parameters: Cmax, Tmax, AUC, clearance rate, and volume of distribution.
Objective: To validate imaging efficacy and target engagement in physiologically relevant models.
Methodology:
Figure 1: Integrated Workflow for Preclinical Development of NIR-II Imaging Agents
Navigating the regulatory landscape requires careful planning from early development stages. For most NIR-II agents, the path to approval follows the Drug Development Toolkit with specific considerations for imaging products.
NIR-II fluorescence imaging agents typically fall under the classification of "diagnostic radiopharmaceuticals" or "imaging drugs" despite not being radioactive, due to their functional similarity. The regulatory pathway is primarily through New Drug Application (NDA), with some specific agents potentially qualifying for 505(b)(2) applications if leveraging previously approved molecules (e.g., ICG derivatives) [58]. Early engagement with regulatory agencies through Pre-IND meetings is critical to align on specific requirements for nonclinical studies and clinical trial design.
Table 2: Key Components of Regulatory Submissions for NIR-II Imaging Agents
| Submission Component | Key Elements | Specific Considerations for NIR-II Agents |
|---|---|---|
| Chemistry, Manufacturing, and Controls (CMC) | Complete description of composition, manufacturing process, specifications, and stability data | Extensive characterization of optical properties (absorption/emission spectra, quantum yield, photostability) [14] |
| Nonclinical Pharmacology | Mechanism of action, proof of concept, dose-response relationships | Demonstration of superiority over existing imaging modalities (e.g., NIR-I) through quantitative metrics (SBR, penetration depth) [26] [58] |
| Nonclinical Toxicology | GLP-compliant safety studies in relevant species, toxicokinetics | Special attention to organ systems where probe accumulation occurs (typically RES organs); ophthalmic toxicity assessment for intraoperative use [34] |
| Clinical Development Plan | Phase I-II-III protocols, risk-benefit analysis, clinical endpoints | For intraoperative agents, pivotal trials typically use endpoints comparing tumor margin identification with and without NIR-II guidance [26] |
Successful translation of NIR-II imaging technology requires access to specialized materials and instrumentation. The following toolkit encompasses key resources referenced in the search results.
Table 3: Essential Research Reagents and Materials for NIR-II Imaging Development
| Category | Specific Examples | Function/Application | Key Characteristics |
|---|---|---|---|
| NIR-II Fluorophores | ICG [58], IR-1048 [26], BBTD-cored small molecules [36], VIPI series cyanines [84] | Contrast generation for deep-tissue imaging | Emission >1000 nm, high quantum yield, appropriate pharmacokinetics |
| Surface Modification Agents | PEGylated phospholipids, poly(maleic anhydride-alt-1-octadecene) (PMAO), bovine serum albumin (BSA) | Biocompatibility enhancement, pharmacokinetic modulation | Improved solubility, reduced protein corona formation, prolonged circulation |
| Imaging Systems | IR-VIVO [26], LightIR [26], InGaAs camera-based systems | Signal detection and quantification | Sensitivity in 1000-1700 nm range, real-time imaging capability, compatibility with clinical use |
| Targeting Moieties | Peptides (RGD), antibodies, affibodies | Molecular specificity for disease biomarkers | High affinity, selectivity, appropriate conjugation chemistry |
| Model Systems | Patient-derived xenografts, orthotopic tumor models, 3D tissue phantoms [26] | Efficacy assessment in biologically relevant contexts | Recapitulation of human disease pathophysiology, appropriate biomarker expression |
Figure 2: Molecular Engineering Strategy for Targeted NIR-II Imaging Probes
The clinical translation of NIR-II fluorescence imaging technology represents a multidisciplinary challenge requiring coordinated advances in materials science, toxicology, regulatory science, and clinical practice. The frameworks and protocols outlined in this document provide a structured approach to addressing the critical biocompatibility, toxicity, and regulatory hurdles. As the field progresses, several emerging trends warrant particular attention: the development of fluorophores with emissions beyond 1500 nm (NIR-IIb) for even better tissue penetration and contrast [36]; the integration of artificial intelligence for enhanced image interpretation and quantification; and the creation of standardized testing protocols specific to NIR-II agents to facilitate regulatory review. With strategic attention to these translational considerations, NIR-II fluorescence imaging is poised to revolutionize clinical practice across multiple medical specialties, particularly in oncology where precise visualization of tumor margins can directly impact patient outcomes. The path to clinical translation, while complex, offers tremendous potential to enhance the precision of diagnostic and therapeutic interventions.
NIR-II fluorescence imaging represents a paradigm shift in optical bioimaging, firmly establishing itself as an indispensable tool for biomedical research with accelerating clinical translation. By leveraging its foundational advantages of deep tissue penetration and high spatial resolution, this technology is revolutionizing applications from intraoperative navigation to the precise evaluation of nanomedicine delivery. While significant progress has been made in fluorophore development and system integration, future efforts must focus on synthesizing brighter, more biocompatible probes, standardizing imaging protocols, and conducting large-scale clinical trials. The convergence of NIR-II imaging with other modalities and its integration into cell therapies and advanced drug delivery platforms promise to further solidify its role as a cornerstone of precision medicine, ultimately improving diagnostic accuracy and therapeutic outcomes for patients.