Previous Article in Journal
Resolution Improvement for Coherent Illumination Microscopy via Incident Light Phase Modulation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Review of Non-Linear Optical Imaging Techniques for Cancer Detection

by
Francisco J. Ávila
Departamento de Física Aplicada, Facultad de Ciencias, Universidad de Zaragoza, 50009 Zaragoza, Spain
Optics 2024, 5(4), 416-433; https://doi.org/10.3390/opt5040031
Submission received: 10 September 2024 / Revised: 4 October 2024 / Accepted: 14 October 2024 / Published: 16 October 2024

Abstract

:
The World Health Organization (WHO) cancer agency predicts that more than 35 million cases of cancer will be experienced in 2050, a 77% increase over the 2022 estimate. Currently, the main cancers diagnosed are breast, lung, and colorectal. There is no standardized tool for cancer diagnoses; initially, clinical procedures are guided by the patient symptoms and usually involve biochemical blood tests, imaging, and biopsy. Label-free non-linear optical approaches are promising tools for tumor imaging, due to their inherent non-invasive biosafe contrast mechanisms and the ability to monitor collagen-related disorders, and biochemical and metabolic changes during cancer progression. In this review, the main non-linear microscopy techniques are discussed, according to three main contrast mechanisms: biochemical, metabolic, and structural imaging.

1. Introduction

The prognosis of cancer depends mainly on early detection, the type of tumor, treatments outcomes, and genetic factors [1]. Tumor markers help diagnose the type and stage of cancer, and to decide which treatment might be effective. Tumor markers also inform about aggressiveness, whether the treatment is effective, or whether the cancer has recurred. However, tumor markers are not universal for each type of cancer, and the same type of tumor marker can show various expression levels, even in two patients with the same stage and type of cancer.
In contrast, cancer cells share common characteristics such as activation, metabolism and dynamic nuclear activity [2]. Tumors require a microenvironment conducive to tumorigenesis. As a consequence, the extracellular matrix undergoes remodeling through tumor cell-matrix interaction that favors progression and metastasis [3].
Non-linear optical microscopy techniques, such as Raman techniques (i.e., stimulated coherent and anti-stokes Raman scattering) [4] and multiphoton microscopy [5], provide powerful tools for cancer detection and diagnosis. Their ability for deep tissue imaging with high-resolution and minimal photo-damage capabilities makes them invaluable in both research and clinical settings. These techniques enable the visualization of structural [6], biochemical [7], and molecular [8] changes in tissues, offering insights into cancer development, progression, and treatment response.
Traditional immunofluorescence microscopy and spectroscopy techniques in oncology require fluorescent dyes that may interfere with the metabolism of cancer cells [9].
Such limitations are overcome with the label-free nature of the reviewed techniques due to their inherent contrast mechanisms. Furthermore, the potential over other non-linear techniques lies in the ability to be combined with each other, providing joint biochemical, metabolic, and structural information on the same sample under study [10,11]. In this review, major advances in non-linear optical imaging modalities that provide chemical, structural, and metabolic information about tumor tissues are discussed.
In particular, the operating principles and applications in cancer detection for biochemical (spontaneous, stimulated and coherent anti-stokes Raman scattering), metabolic (two-photon excitation microscopy, lifetime fluorescence microscopy and transient absorption microscopy), and structural (second and third-harmonic generation microscopy) imaging techniques are analyzed.

2. Biochemical Imaging

2.1. Spontaneous Raman Scattering

Although spontaneous Raman scattering is not a non-linear but a linear effect, it has been added to this section for a better understanding of the principles of non-linear Raman techniques as described in Section 2.2 and Section 2.3.

2.1.1. Principle

Raman spectroscopy is an optical scattering imaging technique that provides non-destructive analysis of the chemical structure of biological tissues [4]. Raman scattering occurs upon the interaction of high-power lasers with the chemical bonds composing the biological sample. The Raman scattered light undergoes a wavelength shift of the incident light whose intensity and spectrum depend on the vibrational (chemical) bond of the specific molecules. Figure 1 schematizes the Raman scattering principle: During an elastic scattering event, the energy of the molecule would not change, and therefore, the scattered light retains the wavelength of the incident source (i.e., Rayleigh scattering). However, a Raman event is an inelastic scattering process in which an energy transfer occurs between the molecule and the scattered photons. If during the light-matter interaction the molecule is excited to a higher vibrational state (i.e., the molecule gains energy) the process is called Stokes Raman scattering. Conversely, if the molecule is relaxed to a lower vibrational state, the scattered photons gain energy with decreasing wavelength; this process is known as anti-Stokes Raman scattering [12].
Raman scattering shift is related to the vibrational state of the constituent atoms, where the interaction of the external electric field E = E 0 c o s ( ω 0 t ) (i.e., the laser source), with the specimen at the atomic scale induces a dipole moment ( P ) that defines the polarizability ( α ̿ ) as [13]:
α ̿ = P E
If the molecular polarizability is expanded into a Taylor series to the first order for a generalized coordinate r = r 0 cos ω 0 t , which is expressed as follows:
α ̿ r = α ̿ 0 + α ̿ r 0 +
The scattered electric field E s c a t t proportional to the induced dipole moment can be expressed as follows:
E s c a t t = α ̿ 0 E 0 cos ω 0 t + 1 / 2 α ̿ r 0 r 0 E 0 cos ω 0 ω t + cos ω 0 + ω t
In Equation (3), the first term corresponds to elastic (Rayleigh) scattering. The second term represents the Raman scattering effect, where the shift frequencies ( ω 0 ω ) and ( ω 0 + ω ) correspond to the Stokes and anti-Stokes Raman processes.

2.1.2. Applications in Cancer Detection

The contrast mechanism of Raman spectroscopy lies in the differences between the wavelength of incident and scattered light. Furthermore, Raman scattering is a label-free optical technology that is sensitive to changes in cells and macromolecules occurring during carcinogenesis; therefore, it is a powerful tool for tumor diagnosis at the molecular level [14,15].
Raman scattering imaging extends classical spectroscopy to 3D visualization of the molecular structure and morphology of tumors [16] and then to the classification of types and subtypes of malignant lesions in cancer diagnosis.
Raman scattering is able to distinguish between normal and cancerous tissues by identifying specific molecular biomarkers. Each tissue type exhibits a unique Raman spectrum based on its molecular constituents, allowing researchers to detect chemical changes indicative of malignancy, such as brain [17] and breast cancer [18,19].
Raman scattering has been employed in surgical settings to provide real-time diagnosis of tumor margins [20]. During cancer surgery, it is crucial to ensure complete removal of the tumor, which typically requires histopathological evaluation by a pathologist. Raman scattering allows for rapid assessment of tissue at the surgical site, potentially reducing waiting time and improving surgical outcomes [21].
The tumor microenvironment plays a critical role in cancer metabolism [22] and progression. Raman scattering can be used to map and analyze the molecular components of the tumor microenvironment, such as the extracellular matrix and infiltrating immune cells. This helps to understand how the environment surrounding the tumor contributes to its growth and resistance to treatment [22].
Raman scattering offers label-free chemical imaging, which has sparked interest in pharmacokinetic research due to the ability to track the biodynamics of anti-cancer drug delivery in real-time [21]. Figure 2a shows the confocal Raman imaging spectra from a human hepatocellular carcinoma cell line (HepG2) in the lipids, cytoplasm, and nucleus regions of the cell. Figure 2b compares the Raman spectra of HepG2 before (green spectrum) and after incubation (dark blue spectrum) with drug delivery polymers based on poly (lactic-co-glycolic acid) nanoparticles (PLGA NPs). The pink spectrum corresponds to the PLGA NPs in the dry state. The cells under study are shown in the insets.

2.2. Stimulated Raman Scattering

2.2.1. Principle

While spontaneous Raman scattering is an incoherent process, in stimulated Raman scattering (SRS) a pump and a Stokes photon interact with the tissue, resulting in the coherent generation of a new photon whose frequency is shifted by the vibrational energy of the molecules in the tissue [24].
SRS enhances the weak Raman signals by transferring energy from the pump to the Stokes beam, enabling the detection of molecular fingerprints with higher spatial resolution and faster acquisition times compared to conventional spontaneous Raman scattering [25].

2.2.2. Applications in Cancer Detection

SRS has emerged as a powerful tool for cancer detection due to its ability to provide high-resolution, label-free imaging of biological tissues. SRS takes advantage of the intrinsic vibrational properties of biomolecules, allowing for the identification and differentiation of cancerous cells from normal tissues based on their different molecular signatures [26]. Recently, the SRS technique has been improved by incorporating high-content SRS histology platforms [27] that have been shown to successfully map unsaturated and saturated lipids, extracellular matrix, cellular proteins, and water in breast tissue, and thereby provide both morphological and chemical information on unstained cancer breast tissues.
One of the major advantages of SRS microscopy in cancer detection is its ability to generate real-time in vivo images without the need for exogenous markers and the visualization of tumor margins.
This feature is particularly valuable in complex clinical procedures such as brain tumor surgery [28], ensuring complete removal of cancerous tissue while preserving healthy surrounding tissue [29]. This precise delineation is critical to reducing the likelihood of cancer recurrence and improving patient outcomes.
Recently, focal therapy has been proposed as a way to precisely remove the malignant mass in prostate cancer while preserving surrounding healthy tissue.
This has been made possible by the emergence of real-time histopathological capabilities of artificial intelligence-enhanced SRS [30]. As an example, Figure 3 (left column) shows stimulated Raman histology (SRH) images corresponding to a non-Hodgkin lymphoma specimen, non-small cell lung cancer brain metastasis (b), and glioblastoma specimens. After image acquisition, a trained convolutional neural network (CNN prediction) displays probability heat maps (middle column) corresponding to the tumor (red), low-quality regions (blue), and regions outside the tumor (green). If the probability maps and SRH are superimposed, the color-coded SRH images can help the surgeon to better interpret the histological information.

2.3. Coherent Anti-Stokes Raman Scattering

2.3.1. Principle

Stimulated Raman scattering and coherent anti-Stokes Raman scattering (CARS) are non-linear optical methods used in vibrational spectroscopy but with different signal generation mechanisms and biomedical applications.
While in SRS, the signal is generated when the energy difference between a pump and a Stokes laser beam matches the vibrational frequency of a molecular bond, CARS generates a new frequency at the anti-Stokes wavelength after the interaction between the pump and Stokes beams. This new signal is a coherent light wave at a frequency that is the sum of the pump frequency and the vibrational frequency of the sample [31].
In CARS, the generated coherent signal is amplified along the propagation direction of the incident beams, allowing high-contrast three-dimensional biochemical imaging [32]. CARS provides stronger signals and the ability to suppress non-resonant fluorescence background signals, making it suitable for applications in biomedical imaging and materials science [33,34].
Furthermore, the incorporation of polarization units into CARS microscopes has made it possible to make the intensity of the anti-Stokes radiation sensitive to the directionality of molecular bonds and then add molecular orientation sensitivity to the regular chemical information of CARS microscopy [35]. These polarimetric approaches added new capabilities to CARS microscopy allowing for the determination of Raman polarization ratios and to study those polarization-sensitive features in CARS spectra from biological specimens [36].

2.3.2. Applications in Cancer Detection

CARS microscopy has emerged as a powerful tool in neuroscience and cancer research due to its ability to provide a coherent signal of vibrational signatures that are also characteristics of specific chemical bonds of biomolecules [37]. Lipid-rich tumor tissues have been related to cancer metastasis; however, if the role of lipids in carcinogenesis is not well understood, lipid-rich tumors cannot be distinguished from carcinoma. One of the potentials of CARS imaging is to visualize the intracellular lipid aggregation in tumor cells [38]. Furthermore, CARS microscopy has been employed to detect circulating tumor cells in the peripheral blood of metastatic patients [39]. Then, the subcellular information provided by CARS microscopy has helped to reveal metabolism in cancer progression [22] and anti-cancer drugs [40].
As an example, Figure 4 shows the ability of CARS microscopy to image cancer cells and their main sub-cellular components. Figure 4a maps the intracellular distribution of nucleic acid (blue), proteins (green), and lipids (red) from a HepG2 cancer cell line. The corresponding CARS spectra are shown in Figure 4b.

3. Metabolic Imaging

3.1. Two-Photon Excitation Fluorescence Microscopy

3.1.1. Principle

Multiphoton imaging arises from the second and third-order non-linear interaction between a high-power ultrafast infrared pulsed laser (traditionally titanium–sapphire lasers) and biological tissues [42]. When the addition of the energy of two or more photons allows the ground fluorophore to be excited to a higher energy excitation state, an electronic transition occurs in a manner similar to regular fluorescence [43] (single photon absorption). Biological molecules in the excited state then fall back to the ground state with spontaneous emission of a photon of lower energy than the sum of the incident photons due to the non-radiative vibrational relaxation process. If two photons are absorbed, the process is known as two-photon excitation fluorescence (TPEF) and has the ability to produce endogenous excitation fluorescence from biomolecules such as lipofuscin, melanin, flavin, and NADPH [44].
In multiphoton imaging, the excitation is confined to a focal volume where the scattered light is inversely proportional to the fourth power of the wavelength of the excitation source. Furthermore, the emission intensity of endogenous fluorophores is proportional to the square of the excitation source; these particular features provide the multiphoton imaging modality with intrinsic optical sectioning capabilities and high penetration power within scattering tissues [45].
Since the experimental demonstration of the suitability of two-photon excitation fluorescence (TPEF) for biological imaging in the 1990s [46], TPEF microscopy has been employed for deep tissue imaging [47], functional imaging in neurons [48], studying the dynamics of stromal cells interaction [49], or intravital subcellular imaging [50]. Subsequently, applications of TPEF microscopy in tumor research have been expanded to tumor microenvironment, metabolism, angiogenesis, and metastasis [51].

3.1.2. Applications in Cancer Detection

In some types of malignancies, the diagnosis often comes at a very late stage such as gastric cancer [52]. Diagnosis of gastric cancer is made by visual assessment of gastrointestinal endoscopy and late confirmation by histopathology. However, misinterpretation in the former (endoscopy) due to the lack of obvious signs by the naked eye, may cause a late impact in the latter (histopathology). In that sense, TPEF microscopy enables the discrimination of normal gastric tissue from cancer, adenoma, or ulcers due to the capabilities of two-photon imaging to visualize and measure enzymatic activity [53].
Cancer-associated fibroblasts (CAF) are cells within the tumor that promote cancer proliferation and extracellular matrix remodeling [54]. They are one of the major stromal cell populations in solid tumors and contribute to drug resistance [55]. In that regard, TPEF microscopy has demonstrated the potential to analyze the structure and spatial organization of gastric cancer cells and monitor spheroid growth. Figure 5 shows a single plane of TPEF images of a bicellular spheroid acquired six days after the start of growth.
Pathological assessment (i.e., histopathology) of treatment response requires tedious and time-consuming labeling processes. As a label-free imaging technique, TPEF microscopy enables the evaluation of histopathological changes induced by neoadjuvant therapy in the preoperative stage, revealing residual tumoral cells, fibrotic reactions, and inflammatory cell infiltration [56].
Tumor cells invade adjacent tissues by secreting degradative enzymes that reduce the pH of lysosomes, and then the extracellular pH of tumor tissues becomes acidic. TPEF microscopy has been shown to be pH-sensitive in the case of human colon cancer [57].
Photodynamic therapy (PDT) is a modern non-invasive light-matter-based interaction method that uses photosensitizers or light-activated drugs for the treatment of non-malignant diseases and various types of cancer [58].
Once the photosensitizers are activated by a selective wavelength, a process of selective destruction of malignant or abnormal cells begins.
One of the main limitations of PDT was the photodamage induced by single-photon excitation of the photosensitizers; however, the application of the concept of two-photon absorption made it possible to overcome this medical limitation due to the negligible photo-toxicity of the non-linear process that makes the technique suitable for the study of living cells [59].
The combination of two-photon excitation (TPE) and PDT gave rise to a new therapy concept (TPE-PDT) with greater penetration power in biological tissue and reduction of cell phototoxicity with promising therapeutic applications in the treatment of tumors [60].
PDT can destroy cancerous cells in inoperable types of cancer [61] or before tumors spread. PDT has demonstrated effectiveness in the treatment of colorectal [62], lung [63], breast, liver, and pancreatic cancer [64]. The main therapeutic mechanisms of PDT are immune response, vascular damage, and direct destruction of cancer cells.
TPEF lifetime microscopy measures the fluorescence decay of endogenous (or exogenous) fluorophore emission, allowing for label-free in-vivo imaging of metabolic dynamics [65] with promising applications in infectious, neurodegenerative, and cancer diseases [66].
In particular, the characterization of NAD(P)H fluorescence lifetime has been shown to be an intrinsic biomarker of the cellular metabolic state of living tissues, capable of tracking tumor cell dynamics [67].
In a recent publication reported by Karrobi et al. [68], lifetime TPEF microscopy was used to investigate changes in the cellular metabolism of breast spheroids, including non-cancerous epithelial breast (MCF-10A) and breast cancer (MD-MB-231) cell spheroid lines embedded in collagen. The findings revealed a greater shift in cancer spheroids towards oxidative phosphorylation and how they invaded collagen over time with stronger metabolic gradient modifications than MCF-10A spheroids.
Figure 6 shows an example of a time-integrated NAD(P)H two-photon excitation fluorescence image from a non-cancerous epithelial breast cell spheroid line (a) and the representative fluorescence decay signal (b).

3.2. Transient Absorption Microscopy

3.2.1. Principle

Transient absorption microscopy (TAM) is an advanced non-linear optical technique used in biology to study biodynamical processes at the subcellular scale with high spatiotemporal resolution. TAM works by probing changes in the absorption of light by a sample after being excited by a short laser pulse, allowing real-time observation of changes in the electronic states of molecules and providing information about fast processes such as energy transfer, charge separation, or molecular dynamics [69].
TAM can be employed for the study of photosynthetic processes, protein dynamics, and the behavior of chromophores and fluorophores within cells. For instance, it has been used to investigate how light energy is converted into chemical energy in photosynthesis, revealing key details about the lifetimes and pathways of excited states in complex biological systems [70].
Overall, transient absorption microscopy provides a unique window into the ultrafast dynamics of biological systems, offering detailed information crucial for unraveling the mechanisms underlying various biological functions [71].

3.2.2. Applications in Cancer Detection

In cancer detection, TAM is particularly useful for identifying molecular signatures that distinguish cancerous cells from healthy ones. For example, TAM can detect differences in the electronic states of chromophores and other biomolecules associated with cancer, such as mitochondrial redox, and then provide insights into the metabolic and structural changes that occur during neurodegenerative diseases, diabetes, or cancer [72]. Additionally, TAM has been used to study the dynamics of photosensitizers in photodynamic therapy (PDT), a cancer treatment that relies on light-activated compounds to kill cancer cells. By monitoring the transient absorption signals of these photosensitizers, researchers can optimize PDT treatment protocols and improve the efficacy of cancer therapy [73].
One of the most recent studies published by Xin and coworkers [74], showed the power of combining stimulated Raman scattering and transient absorption microscopy for real-time observation of subcellular components (i.e., proteins, lipids, and DNA). Their findings revealed induced oxidative damage in a single live cancer cell during PDT.

4. Structural Imaging

4.1. Second Harmonic Generation

Cancer cells are responsible for tumor growth and migration to other organs; however, the extracellular matrix undergoes remodeling that promotes the progression of metastasis and tumorigenesis [75]. The next subsection discusses the “gold standard” non-linear imaging modality to visualize the extracellular matrix primarily composed of collagen [76].

4.1.1. Principle

Multiphoton microscopy is sensitive to harmonic generations, where two or three incident photons can be instantaneously converted into a single photon with half or one-third of the excitation wavelength giving rise to second and third harmonic generation microscopy (SHG and THG), respectively [77]. Those frequency conversions are not absorptive processes and occur without energy dissipation as a result of excitation photons summation and phase matching induced by the non-linear susceptibility of the biological structures [78].
The contrast mechanism of SHG and THG microscopy, unlike TPEF which requires endogenous fluorophores, arises from the frequency conversion of the incident photons due to the non-linearity of the electromagnetic response of the specific molecular structures of biological tissues. According to the non-linear optics theory, the excited non-linear polarization ( P ) is related to the excitation laser source by means of the non-linear susceptibility, χ ^ (n) [79]:
P = ε 0 + χ ^ ( 1 )   ·   E + χ ^ ( 2 )   ·   E E + χ ^ ( 3 )   ·   E E E +
Therefore, the generation of harmonics depends not only on the need for ultrafast and intense pulsed laser but also on the non-linear susceptibility of the biological tissues. The SHG signal is provided only in non-centrosymmetric media such as fibrillar collagen [80]. In contrast, THG arises from sources such as lipid-water interfaces [81] and myelinated axons [82]. Figure 7 schematizes the main multiphoton imaging modalities, which provide label-free three-dimensional visualizations of biological structures at the cellular level with negligible phototoxicity. These features make multiphoton microscopy a potential tool for the analysis of tumor lesions.
The extracellular matrix is not only responsible for cell maintenance, but also for cell migration and proliferation [83]. In the extracellular matrix (ECM) of connective tissues, collagen is the most abundant protein and plays a fundamental role in functional structure and cellular processes [84].
Type-I fibrillar collagen is organized in a non-centrosymmetric molecular structure and possesses strong non-linear susceptibility. Nowadays, second harmonic generation (SHG) microscopy (see Figure 6b for the illustration of the SHG concept) is considered the gold standard imaging technique for visualizing and characterizing collagen-formed tissues [85]. The SHG signal is endogenous; therefore, the contrast mechanism arises from the structure of the sample, making this imaging technique sensitive to the molecular structure of the collagen-based connective tissue [86].
Beyond collagen-based tissues, SHG microscopy allows visualization of the structural proteins [86], membrane potential in neurons [87], or myosin filaments [88] when combining SHG and TPEF microscopy.
The SHG signal is a coherent process of tensorial nature in which forward (F-SHG) and backward (B-SHG) emissions coexist; the ratio of F/B-SHG depends on the structure of the susceptibility tensor [89]. Therefore, SHG imaging is a powerful optical microscopy technique to analyze collagen disorders (remodeling) associated to cancer and fibrous connective tissues [90].

4.1.2. Applications in Cancer Detection

In cancer progression, the extracellular matrix is remodeled alongside abnormal cell growth [91]. In particular, collagen undergoes fibrillation that induces increased stiffness in the ECM and promotes angiogenesis and cancer invasion [92]. Furthermore, changes in SHG directionality emission (i.e., the F/B-SHG ratio) have been found to be altered in the tumor bulk but not at the stroma interface [93], where high values of the F/B-SHG ratio describe highly organized ECM tissues.
Therefore, the characterization of spatial remodeling of collagen in the ECM is crucial as tumor invasion and metastasis are facilitated by alterations in the ECM [94].
SHG microscopy images from collagen fibers require quantitative analysis to first detect alteration in the spatial patterns and then numerically compute the collagen remodeling and degree of organization. The structure tensor, fast Fourier transform, Wavelet, or Hough transforms are well-established image transformation methods from which structural information can be extracted. An overview of the main image processing techniques for quantitative analysis of SHG imaging can be found in [95].
In particular, the structure tensor [96] has been demonstrated as a useful method to quantify the organization of collagen-based tissues and classify different spatial patterns. The Hough transform was successfully used to discriminate between normal and malignant tissues, analyzing the surrounding collagen of thyroid cancer nodule capsules [97].
The SHG process exhibits a strong polarization dependence [98]. The polarization-sensitive nature of SHG (P-SHG) microscopy has allowed the measurement of the second-order non-linear optical susceptibility tensor [99], providing information about the ultrastructure of collagen. As an example, Figure 8 compares the spatially-resolved components of the non-linear susceptibility tensor in normal and cancerous breast tissues, obtained using P-SHG.
Tokarz et al. [101] used P-SHG microscopy to calculate the components of the susceptibility tensor from histopathological samples of cancerous thyroid tissues. They were able to measure the molecular chirality of the malignant tissues and found cancer-related collagen disorders at molecular levels.
Polarimetric SHG microscopy reveals collagen orientation and ultrastructure that cannot be accessed by non-polarized SHG imaging. P-SHG has also been used to distinguish between breast, osteosarcoma, liver fibrosis, or melanoma tissues [102,103,104,105].
Therefore, pathologists can benefit from the invaluable ultrastructural information provided by polarization-sensitive SHG microscopy.

4.2. Third Harmonic Generation

4.2.1. Principle

While SHG microscopy arises mainly from non-centrosymmetric proteins such as Type-I collagen, the third-harmonic generation (THG) signal originates mainly from interfaces between water and lipid-rich structures such as lipid droplets or membrane lipid layers [106,107]. In this non-linear process, the frequency of the excitation light is tripled after the interaction with biological tissue (see Figure 7c) and provides information about the refractive index discontinuities of material interfaces via the third-order non-linear susceptibility tensor χ(3) [108].

4.2.2. Applications in Cancer Detection

THG offers structural information of single cells with submicron resolution and intracellular heterogeneities allowing discrimination between benign and tumorous tissues [109]. Gavgiotaki et al. [110] reported quantitative differentiation of malignant cells in benign and tumorous breast biopsies and found how disease severity altered cell morphological information that distinguished cell malignant grade.
Figure 9 shows THG imaging of cells isolated from healthy tissue (left column) and for different grades of breast cancer. The top and bottom rows compare nuclear irregularities and nucleoli heterogeneities. The ability of THG to detect variability in tumor cell size and shape in a non-invasive and label-free manner provides invaluable structural information from cancer biopsies at the subcellular level of resolution.

5. Discussion

In this article, the main methods of non-linear optical microscopy were analyzed, considering three different contrast mechanisms: biochemical, metabolic, and structural. Table 1 summarizes the reviewed technique, the principle, the contrast mechanisms, and the main applications, as well as the main advantages and disadvantages.
Cancer cells exhibit non-linear optical properties that can also be considered as a predictive indicator of cellular response induced by the treatment [111]. The emergence of biocompatible nanomaterials has allowed the conversion of macroscopic analysis of tumor tissues to molecular diagnosis tools in cancer research. Nanocluster enables low-toxicity, biocompatible tumor labeling that constitutes optical contrast biomarkers exhibiting non-linear optical properties [112] that can be analyzed with the reviewed optical methods at the cellular and molecular scales. Furthermore, nanoparticles, including gold nanoparticles, protein nanoparticles, cell membrane nanoparticles quantum dots, and others, can be controlled externally by optical methods, magnetism, and enzymes that make them behave like smart nanoparticles that are emerging as the new era of intelligent nanoparticle cancer treatment [113]. Studying the non-linear optical properties of cancer tissues allows for rapid estimation of cellular damage or response to treatments, spatial remodeling of the ECM in cancer growth, and the optimization of two-photon photodynamic therapy [114]. Furthermore, a recent study reported by Hoque et al. [115] reported a promising method to detect cancer by analyzing the non-linear optical properties of blood plasma. Their results achieved an accuracy of 92% in discriminating between normal and cancerous samples.

6. Conclusions

Cancer cells induce non-linear optical properties in cell-matrix interaction [116] that can be measured from multimodal optical approaches that also provide invaluable monitoring of anti-cancer drugs and support the application of photodynamic therapies and the use of nanomaterials as an emerging paradigm for cancer treatment.
Non-linear optical techniques reviewed in this paper, such as multiphoton microscopy and Raman approaches, are technically sophisticated and their complexity often requires specialized equipment and expertise, limiting their clinical application. Those technical limitations added to elevated costs as well as required training and infrastructure make the implementation of these techniques in hospitals difficult. Thus, non-linear techniques are still confined to research labs and not accessible to clinical practices for diagnostic procedures.
Other limitations of non-linear optical techniques for cancer detection are the penetration depth, the lack of standardization of clinical guidelines, and in-vivo applications, in which motion artifacts, blood flow, and complexity of the tumor environment can alter the image quality and then the diagnostic accuracy.
Those limitations may be substantially addressed by focusing on the recommendations summarized in Table 2.
To conclude, by addressing these limitations and focusing on the recommendations, non-linear optical techniques could become a truly promising, more practical, and effective tool for cancer detection implemented in clinical settings in the near future.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The author declares no conflicts of interest.

References

  1. Gospodarowicz, M.; O’Sullivan, B. Prognostic factors in cancer. Semin. Surg. Oncol. 2003, 21, 13–18. [Google Scholar] [CrossRef] [PubMed]
  2. DeBerardinis, R.J.; Chandel, N.S. Fundamentals of cancer metabolism. Sci. Adv. 2016, 2, e1600200. [Google Scholar] [CrossRef] [PubMed]
  3. Winkler, J.; Abisoye-Ogunniyan, A.; Metcalf, K.J.; Werb, Z. Concepts of extracellular matrix remodelling in tumour progression and metastasis. Nat. Commun. 2020, 11, 5120. [Google Scholar] [CrossRef] [PubMed]
  4. Jones, R.R.; Hooper, D.C.; Zhang, L.; Wolverson, D.; Valev, V.K. Raman Techniques: Fundamentals and Frontiers. Nanoscale Res. Lett. 2019, 14, 231. [Google Scholar] [CrossRef] [PubMed]
  5. Larson, A. Multiphoton microscopy. Nat. Photon. 2011, 5, 1. [Google Scholar] [CrossRef]
  6. Schenke-Layland, K. Non-invasive multiphoton imaging of extracellular matrix structures. J. Biophotonics 2008, 1, 451–462. [Google Scholar] [CrossRef]
  7. Hu, F.; Shi, L.; Min, W. Biological imaging of chemical bonds by stimulated Raman scattering microscopy. Nat. Methods 2019, 16, 830–842. [Google Scholar] [CrossRef]
  8. Van Steenbergen, V.; Boesmans, W.; Li, Z.; de Coene, Y.; Vints, K.; Baatsen, P.; Dewachter, I.; Ameloot, M.; Clays, K.; Berghe, P.V. Molecular understanding of label-free second harmonic imaging of microtubules. Nat. Commun. 2019, 10, 3530. [Google Scholar] [CrossRef]
  9. Szeto, H.H.; Schiller, P.W.; Zhao, K.; Luo, G. Fluorescent dyes alter intracellular targeting and function of cell-penetrating tetrapeptides. FASEB J. 2004, 19, 118–120. [Google Scholar] [CrossRef]
  10. Jhan, J.-W.; Chang, W.-T.; Chen, H.-C.; Lee, Y.T.; Wu, M.F.; Chen, C.H.; Liau, I. Integrated multiple multi-photon imaging and Raman spectroscopy for charac-terizing structure-constituent correlation of tissues. Opt. Express 2008, 16, 16431–16441. [Google Scholar] [CrossRef]
  11. Quansah, E.; Shaik, T.A.; Çevik, E.; Wang, X.; Höppener, C.; Meyer-Zedler, T.; Krafft, C. Investigating biochemical and structural changes of glycated collagen using mul-timodal multiphoton imaging, Raman spectroscopy, and atomic force microscopy. Anal. Bioanal. Chem. 2023, 415, 6257–6267. [Google Scholar] [CrossRef] [PubMed]
  12. Kauffmann, T.H.; Kokanyan, N.; Fontana, M.D. Use of Stokes and anti-Stokes Raman scattering for new applications. J. Raman Spectrosc. 2018, 50, 418–424. [Google Scholar] [CrossRef]
  13. Griffiths, D.J. Introduction to Electrodynamics, 3rd ed.; Pearson Education Dorling Kindersley: Glenview, IL, USA, 2007; ISBN 81-7758-293-3. [Google Scholar]
  14. Qi, Y.F.; Liu, Y.H.; Liu, D.M. Research progress on application of Raman spectroscopy in tumor diagnosis. Laser Optoelectron. Prog. 2020, 57, 220001. [Google Scholar]
  15. Ilchenko, O.; Pilhun, Y.; Kutsyk, A. Towards Raman imaging of centimeter scale tissue areas for real-time opto-molecular vis-ualization of tissue boundaries for clinical applications. Light Sci. Appl. 2022, 11, 143. [Google Scholar] [CrossRef] [PubMed]
  16. Auner, G.W.; Koya, S.K.; Huang, C.; Broadbent, B.; Trexler, M.; Auner, Z.; Elias, A.; Curtin Mehne, K.; Brusatori, M.A. Applications of Raman spectroscopy in cancer diagnosis. Cancer Metastasis Rev. 2018, 37, 691–717. [Google Scholar] [CrossRef]
  17. Hollon, T.; Orringer, D.A. Label-free brain tumor imaging using Raman-based methods. J. Neuro-Oncol. 2021, 151, 393–402. [Google Scholar] [CrossRef]
  18. Haka, A.S.; Shafer-Peltier, K.E.; Fitzmaurice, M.; Crowe, J.; Dasari, R.R.; Feld, M.S. Diagnosing breast cancer by using Raman spectroscopy. Proc. Natl. Acad. Sci. USA 2005, 102, 12371–12376. [Google Scholar] [CrossRef]
  19. Hanna, K.; Krzoska, E.; Shaaban, A.M.; Muirhead, D.; Abu-Eid, R.; Speirs, V. Raman spectroscopy: Current applications in breast cancer diagnosis, challenges and future prospects. Br. J. Cancer 2021, 126, 1125–1139. [Google Scholar] [CrossRef] [PubMed]
  20. Santos, I.P.; Barroso, E.M.; Bakker Schut, T.C.; Caspers, P.J.; van Lanschot, C.G.F.; Choi, D.-H.; van der Kamp, M.F.; Smits, R.W.H.; van Doorn, R.; Verdijk, R.M.; et al. Raman spectroscopy for cancer detection and cancer surgery guidance: Translation to the clinics. Analyst 2017, 142, 3025–3047. [Google Scholar] [CrossRef]
  21. Kong, K.; Rowlands, C.J.; Varma, S.; Perkins, W.; Leach, I.H.; Koloydenko, A.A.; Williams, H.C.; Notingher, I. Diagnosis of tumors during tissue-conserving surgery with integrated autofluo-rescence and Raman scattering microscopy. Proc. Natl. Acad. Sci. USA 2013, 110, 15189–15194. [Google Scholar]
  22. Xu, J.; Yu, T.; Zois, C.E.; Cheng, J.-X.; Tang, Y.; Harris, A.L.; Huang, W.E. Unveiling Cancer Metabolism through Spontaneous and Coherent Raman Spectroscopy and Stable Isotope Probing. Cancers 2021, 13, 1718. [Google Scholar] [CrossRef] [PubMed]
  23. Romero, G.; Rojas, E.; Estrela-Lopis, I.; Donath, E.; Moya, S.E. Spontaneous confocal Raman microscopy--a tool to study the uptake of nanopar-ticles and carbon nanotubes into cells. Nanoscale Res. Lett. 2011, 6, 429. [Google Scholar] [CrossRef] [PubMed]
  24. Nandakumar, P.; Kovalev, A.; Volkmer, A. Vibrational imaging based on stimulated Raman scattering microscopy. New J. Phys. 2009, 11, 033026. [Google Scholar] [CrossRef]
  25. Freudiger, C.W.; Min, W.; Saar, B.G.; Lu, S.; Holtom, G.R.; He, C.; Tsai, J.C.; Kang, J.X.; Xie, X.S. Label-free biomedical imaging with high sensitivity by stimulated Raman scat-tering microscopy. Science 2008, 322, 1857–1861. [Google Scholar] [CrossRef]
  26. Zhu, M.; Chen, X.; Chi, M.; Wu, Y.; Zhang, M.; Gao, S. Spontaneous-stimulated Raman co-localization dual-modal analysis approach for efficient identification of tumor cells. Talanta 2024, 277, 126297. [Google Scholar] [CrossRef]
  27. Ni, H.; Dessai, C.P.; Lin, H.; Wang, W.; Chen, S.; Yuan, Y.; Ge, X.; Ao, J.; Vild, N.; Cheng, J.-X. High-content stimulated Raman histology of human breast cancer. Theranostics 2024, 14, 1361–1370. [Google Scholar] [CrossRef]
  28. Ji, M.; Lewis, S.; Camelo-Piragua, S.; Ramkissoon, S.H.; Snuderl, M.; Venneti, S.; Fisher-Hubbard, A.; Garrard, M.; Fu, D.; Wang, A.C.; et al. Detection of human brain tumor infiltration with quantitative stimulated Raman scattering microscopy. Sci. Transl. Med. 2015, 7, 309ra163. [Google Scholar] [CrossRef]
  29. Saar, B.G.; Freudiger, C.W.; Reichman, J.; Stanley, C.M.; Holtom, G.R.; Xie, X.S. Video-Rate Molecular imaging in vivo with stimulated raman scattering. Science 2010, 330, 1368–1370. [Google Scholar] [CrossRef] [PubMed]
  30. Reinecke, D.; von Spreckelsen, N.; Mawrin, C.; Ion-Margineanu, A.; Fürtjes, G.; Jünger, S.T.; Khalid, F.; Freudiger, C.W.; Timmer, M.; Ruge, M.I.; et al. Novel rapid intraoperative qualitative tumor detection by a residual con-volutional neural network using label-free stimulated Raman scattering microscopy. Acta Neuropathol. Commun. 2022, 10, 109. [Google Scholar] [CrossRef]
  31. Evans, C.L.; Xie, X.S. Coherent Anti-Stokes Raman Scattering Microscopy: Chemical Imaging for Biology and Medicine. Annu. Rev. Anal. Chem. 2008, 1, 883–909. [Google Scholar] [CrossRef]
  32. Zumbusch, A.; Holtom, G.R.; Xie, X.S. Three-dimensional vibrational imaging by coherent anti-Stokes Raman scattering. Phys. Rev. Lett. 1999, 82, 4142–4145. [Google Scholar] [CrossRef]
  33. Gachet, D.; Billard, F.; Rigneault, H. Background-free coherent anti-Stokes Raman spectroscopy near transverse interfaces: A vectorial study. J. Opt. Soc. Am. B 2008, 25, 1655–1666. [Google Scholar] [CrossRef]
  34. Heuke, S.; Rigneault, H. Coherent Stokes Raman scattering microscopy (CSRS). Nat. Commun. 2023, 14, 3337. [Google Scholar] [CrossRef] [PubMed]
  35. Vito, G.M.; Bifone, A.; Piazza, V. Rotating-polarization CARS microscopy: Combining chemical and molecular orientation sen-sitivity. Opt. Express 2012, 20, 29369–29377. [Google Scholar] [CrossRef]
  36. Li, S.; Li, Y.; Yi, R.; Liu, L.; Qu, J. Coherent Anti-Stokes Raman Scattering Microscopy and Its Applications. Front. Phys. 2020, 8, 598420. [Google Scholar] [CrossRef]
  37. McCullagh, E.A.; Poleg, S.; Stich, D.; Moldovan, R.; Klug, A. Coherent Anti-Stokes Raman Spectroscopy (CARS) Application for Imaging Myelination in Brain Slices. J. Vis. Exp. 2022, 185, e64013. [Google Scholar] [CrossRef]
  38. Le, T.T.; Huff, T.B.; Cheng, J.-X. Coherent anti-Stokes Raman scattering imaging of lipids in cancer metastasis. BMC Cancer 2009, 9, 42. [Google Scholar] [CrossRef]
  39. Mitra, R.; Chao, O.; Urasaki, Y.; Goodman, O.B.; Le, T.T. Detection of lipid-rich prostate circulating tumour cells with coherent anti-stokes raman scattering microscopy. BMC Cancer 2012, 12, 540. [Google Scholar] [CrossRef] [PubMed]
  40. Zeng, J.; Zhao, W.; Yue, S. Coherent Raman Scattering Microscopy in Oncology Pharmacokinetic Research. Front. Pharmacol. 2021, 12, 630167. [Google Scholar] [CrossRef]
  41. Vernuccio, F.; Vanna, R.; Ceconello, C.; Bresci, A.; Manetti, F.; Sorrentino, S.; Ghislanzoni, S.; Lambertucci, F.; Motiño, O.; Martins, I.; et al. Full-Spectrum CARS Microscopy of Cells and Tissues with Ultrashort White-Light Continuum Pulses. J. Phys. Chem. B 2023, 127, 4733–4745. [Google Scholar] [CrossRef]
  42. Williams, R.M.; Zipfel, W.R.; Webb, W.W. Multiphoton microscopy in biological research. Curr. Opin. Chem. Biol. 2001, 5, 603–608. [Google Scholar] [CrossRef]
  43. Stringari, C.; Abdeladim, L.; Malkinson, G.; Mahou, P.; Solinas, X.; Lamarre, I.; Brizion, S.; Galey, J.-B.; Supatto, W.; Legouis, R.; et al. Multicolor two-photon imaging of endogenous fluorophores in living tissues by wavelength mixing. Sci. Rep. 2017, 7, 3792. [Google Scholar] [CrossRef] [PubMed]
  44. Vergen, J.; Hecht, C.; Zholudeva, L.V.; Marquardt, M.M.; Hallworth, R.; Nichols, M.G. Metabolic imaging using two-photon excited NADH intensity and fluorescence lifetime imaging. Microsc. Microanal. 2012, 18, 761–770. [Google Scholar] [CrossRef] [PubMed]
  45. Centonze, V.E.; White, J.G. Multiphoton excitation provides optical sections from deeper within scattering specimens than Confocal imaging. Biophys. J. 1998, 75, 2015–2024. [Google Scholar] [CrossRef]
  46. Denk, W.; Strickler, J.H.; Webb, W.W. Two-Photon Laser Scanning Fluorescence Microscopy. Science 1990, 248, 73–76. [Google Scholar] [CrossRef]
  47. Helmchen, F.; Denk, W. Deep tissue two-photon microscopy. Nat. Methods 2005, 2, 932–940. [Google Scholar] [CrossRef]
  48. Denk, W.; Delaney, K.; Gelperin, A.; Kleinfeld, D.; Strowbridge, B.; Tank, D.; Yuste, R. Anatomical and functional imaging of neurons using 2-photon laser scanning microscopy. J. Neurosci. Methods 1994, 54, 151–162. [Google Scholar] [CrossRef] [PubMed]
  49. Bousso, P.; Bhakta, N.R.; Lewis, R.S.; Robey, E. Dynamics of thymocyte-stromal cell interactions visualized by two-photon mi-croscopy. Science 2002, 296, 1876–1880. [Google Scholar] [CrossRef]
  50. Lu, Z.; Zuo, S.; Shi, M.; Fan, J.; Xie, J.; Xiao, G.; Yu, L.; Wu, J.; Dai, Q. Long-term intravital subcellular imaging with confocal scanning light-field microscopy. Nat. Biotechnol. 2024, 38802562. [Google Scholar] [CrossRef]
  51. Jun, L.I.U. Two-photon microscopy in pre-clinical and clinical cancer research. Front. Optoelectron. 2015, 8, 141–151. [Google Scholar]
  52. Alzeeb, G.; Dubreuil, M.; Arzur, D.; Rivet, S.; Corcos, L.; Le Grand, Y.; Le Jossic-Corcos, C. Gastric cancer multicellular spheroid analysis by two-photon microscopy. Biomed. Opt. Express 2022, 13, 3120–3130. [Google Scholar] [CrossRef] [PubMed]
  53. Noh, C.-K.; Lim, C.S.; Lee, G.H.; Cho, M.K.; Lee, H.W.; Roh, J.; Kim, Y.B.; Lee, E.; Park, B.; Kim, H.M.; et al. A Diagnostic Method for Gastric Cancer Using Two-Photon Microscopy with Enzyme-Selective Fluorescent Probes: A Pilot Study. Front. Oncol. 2021, 11, 634219. [Google Scholar] [CrossRef]
  54. Yang, D.; Liu, J.; Qian, H.; Zhuang, Q. Cancer-associated fibroblasts: From basic science to anticancer therapy. Exp. Mol. Med. 2023, 55, 1322–1332. [Google Scholar] [CrossRef]
  55. Franchi-Mendes, T.; Lopes, N.; Brito, C. Heterotypic Tumor Spheroids in Agitation-Based Cultures: A Scaffold-Free Cell Model That Sustains Long-Term Survival of Endothelial Cells. Front. Bioeng. Biotechnol. 2021; 9, 649949. [Google Scholar]
  56. Li, L.; Hong, S.; Kang, D.; Huang, X.; Zhang, S.; Zhan, Z.; Zhou, Y.; Chen, J. Two-photon imaging reveals histopathological changes in gastric tumor microenvironment induced by neoadjuvant treatment. Biomed. Opt. Express 2023, 14, 5085–5096. [Google Scholar] [CrossRef]
  57. Hong, S.T.; Kim, T.H.; Choi, J.W.; Park, S.J.; Kwon, S.A.; Paik, K.C.; Han, M.S.; Kim, E.S.; Chun, H.J.; Heo, J.N.; et al. Two-Photon Probes for pH: Detection of Human Colon Cancer using Two-Photon Microscopy. Anal Chem. 2017, 89, 9830–9835. [Google Scholar] [CrossRef] [PubMed]
  58. Kwiatkowski, S.; Knap, B.; Przystupski, D.; Saczko, J.; Kędzierska, E.; Knap-Czop, K.; Kotlińska, J.; Michel, O.; Kotowski, K.; Kulbacka, J. Photodynamic therapy-mechanisms, photosensitizers and combinations. Biomed. Pharmacother. 2018, 106, 1098–1107. [Google Scholar] [CrossRef]
  59. Benninger, R.K.; Piston, D.W. Two-photon excitation microscopy for the study of living cells and tissues. Curr. Protoc. Cell Biol. 2013, 59, 4.11.1–4.11.24. [Google Scholar] [CrossRef]
  60. Juvekar, V.; Lee, D.J.; Park, T.G.; Samanta, R.; Kasar, P.; Kim, C.; Rotermund, F.; Kim, H.M. Two-photon excitation photosensitizers for photodynamic therapy: From small-molecules to nano-complex systems. Coord. Chem. Rev. 2024, 506, 215711. [Google Scholar] [CrossRef]
  61. Jung, H.S.; Kim, H.J. Definitive surgery and intraoperative photodynamic therapy for locally advanced non-small cell lung cancer: A case report. World J. Surg. Oncol. 2022, 20, 265. [Google Scholar] [CrossRef]
  62. Rodrigues, J.A.; Correia, J.H. Photodynamic Therapy for Colorectal Cancer: An Update and a Look to the Future. Int. J. Mol. Sci. 2023, 24, 12204. [Google Scholar] [CrossRef]
  63. Xu, C.; Law, S.K.; Leung, A.W.N. Comparison of the Differences between Two-Photon Excitation, Upconversion, and Con-ventional Photodynamic Therapy on Cancers in In Vitro and In Vivo Studies. Pharmaceuticals 2024, 17, 663. [Google Scholar] [CrossRef] [PubMed]
  64. Starkey, J.R.; Rebane, A.K.; Drobizhev, M.A.; Meng, F.; Gong, A.; Elliott, A.; McInnerney, K.; Spangler, C.W. New Two-Photon Activated Photodynamic Therapy Sensitizers Induce Xenograft Tumor Regressions after Near-IR Laser Treatment through the Body of the Host Mouse. Clin. Cancer Res. 2008, 14, 6564–6573. [Google Scholar] [CrossRef] [PubMed]
  65. Bower, A.J.; Li, J.; Chaney, E.J.; Marjanovic, M.; Spillman, D.R.; Boppart, S.A. High-speed imaging of transient metabolic dynamics using two-photon fluorescence lifetime imaging microscopy. Optica 2018, 5, 1290–1296. [Google Scholar] [CrossRef] [PubMed]
  66. Ranawat, H.; Pal, S.; Mazumder, N. Recent trends in two-photon auto-fluorescence lifetime imaging (2P-FLIM) and its biomedical applications. Biomed. Eng. Lett. 2019, 9, 293–310. [Google Scholar] [CrossRef]
  67. Liang, W.; Chen, D.; Guan, H.; Park, H.C.; Li, K.; Li, A.; Li, M.J.; Gannot, I.; Li, X. Label-Free Metabolic Imaging In Vivo by Two-Photon Fluorescence Lifetime Endomi-croscopy. ACS Photonics 2022, 9, 4017–4029. [Google Scholar] [CrossRef]
  68. Karrobi, K.; Tank, A.; Fuzail, M.A.; Kalidoss, M.; Tilbury, K.; Zaman, M.; Ferruzzi, J.; Roblyer, D. Fluorescence Lifetime Imaging Microscopy (FLIM) reveals spatial-metabolic changes in 3D breast cancer spheroids. Sci. Rep. 2023, 13, 3624. [Google Scholar] [CrossRef]
  69. Sundström, V. Femtobiology. Annu. Rev. Phys. Chem. 2008, 59, 53–77. [Google Scholar] [CrossRef]
  70. Berera, R.; van Grondelle, R.; Kennis, J.T.M. Ultrafast transient absorption spectroscopy: Principles and application to photosynthetic systems. Photosynth. Res. 2009, 101, 105–118. [Google Scholar] [CrossRef]
  71. Zhu, Y.; Cheng, J.-X. Transient absorption microscopy: Technological innovations and applications in materials science and life science. J. Chem. Phys. 2020, 152, 020901. [Google Scholar] [CrossRef]
  72. Wang, E.; Whitcomb, L.A.; Chicco, A.J.; Wilson, J.W. Transient absorption spectroscopy and imaging of redox in muscle mitochondria. Biomed. Opt. Express 2022, 13, 2103–2116. [Google Scholar] [CrossRef]
  73. Zhang, S.; Wang, J.; Wang, Z.; Shao, M.; Zhang, C.; Chen, X.; Sun, J.; Kwok, R.T.K.; Lam, J.W.Y.; Tang, B.Z. Study of transient absorption spectroscopy of a D–π–A structure aggregation-induced emission luminogen and its photodynamic therapy application. J. Mater. Chem. B 2024, 12, 8349–8356. [Google Scholar] [CrossRef] [PubMed]
  74. Xin, L.; Luo, Z.; Liu, X.; Huang, Z. Unveiling the Spatiotemporal and Dose Responses within a Single Live Cancer Cell to Pho-toswitchable Upconversion Nanoparticle Therapeutics Using Hybrid Hyperspectral Stimulated Raman Scattering and Transient Absorption Microscopy. Anal. Chem. 2024, 96, 6148–6157. [Google Scholar] [CrossRef]
  75. Tanzer, M.L. Current concepts of extracellular matrix. J. Orthop. Sci. 2006, 11, 326–331. [Google Scholar] [CrossRef] [PubMed]
  76. Rozario, T.; DeSimone, D.W. The extracellular matrix in development and morphogenesis: A dynamic view. Dev. Biol. 2010, 341, 126–140. [Google Scholar] [CrossRef]
  77. James, D.S.; Campagnola, P.J. Recent Advancements in Optical Harmonic Generation Microscopy: Applications and Perspectives. BME Front. 2021, 2021, 3973857. [Google Scholar] [CrossRef]
  78. LaComb, R.; Nadiarnykh, O.; Townsend, S.S.; Campagnola, P.J. Phase Matching Considerations in Second Harmonic Generation from Tissues: Effects on Emission Directionality, Conversion Efficiency and Observed Morphology. Opt. Commun. 2007, 281, 1823–1832. [Google Scholar] [CrossRef]
  79. Boyd, R.W. Nonlinear Optics; Academic Press: Cambridge, MA, USA, 1992. [Google Scholar]
  80. Williams, R.M.; Zipfel, W.R.; Webb, W.W. Interpreting second-harmonic generation images of collagen I fibrils. Biophys. J. 2005, 88, 1377–1386. [Google Scholar] [CrossRef] [PubMed]
  81. Weigelin, B.; Bakker, G.-J.; Friedl, P. Third harmonic generation microscopy of cells and tissue organization. J. Cell Sci. 2016, 129, 245–255. [Google Scholar] [CrossRef]
  82. Farrar, M.J.; Wise, F.W.; Fetcho, J.R.; Schaffer, C.B. In Vivo imaging of myelin in the vertebrate central nervous system using third harmonic generation microscopy. Biophys. J. 2011, 100, 1362–1371. [Google Scholar] [CrossRef]
  83. Popova, N.V.; Jücker, M. The Functional Role of Extracellular Matrix Proteins in Cancer. Cancers 2022, 14, 238. [Google Scholar] [CrossRef]
  84. Kular, J.K.; Basu, S.; Sharma, R.I. The extracellular matrix: Structure, composition, age-related differences, tools for analysis and applications for tissue engineering. J. Tissue Eng. 2014, 5, 2041731414557112. [Google Scholar] [CrossRef]
  85. Aghigh, A.; Bancelin, S.; Rivard, M.; Pinsard, M.; Ibrahim, H.; Légaré, F. Second harmonic generation microscopy: A powerful tool for bio-imaging. Biophys. Rev. 2023, 15, 43–70. [Google Scholar] [CrossRef] [PubMed]
  86. Mohler, W.; Millard, A.C.; Campagnola, P.J. Second harmonic generation imaging of endogenous structural proteins. Methods 2003, 29, 97–109. [Google Scholar] [CrossRef]
  87. Jiang, J.; Yuste, R. Second-Harmonic generation imaging of membrane potential with photon counting. Microsc. Microanal. 2008, 14, 526–531. [Google Scholar] [CrossRef] [PubMed]
  88. Wallace, S.J.; Morrison, J.L.; Botting, K.J.; Kee, T.W. Second-harmonic generation and two-photon-excited autofluorescence microscopy of cardiomyocytes: Quantification of cell volume and myosin filaments. J. Biomed. Opt. 2008, 13, 064018. [Google Scholar] [CrossRef] [PubMed]
  89. Liu, J.; Jin, Y.; Liu, H.; Wang, F.; Deng, X. The forward and backward second-harmonic generation from crystallized collagen fibre with tightly focused linearly polarized beams. J. Opt. 2012, 14, 055301. [Google Scholar] [CrossRef]
  90. Keikhosravi, A.; Bredfeldt, J.S.; Sagar, A.K.; Eliceiri, K.W. Second-harmonic generation imaging of cancer. Methods Cell Biol. 2014, 123, 531–546. [Google Scholar]
  91. Mierke, C.T. The matrix environmental and cell mechanical properties regulate cell migration and contribute to the invasive phenotype of cancer cells. Rep Prog Phys. 2019, 82, 064602. [Google Scholar] [CrossRef]
  92. Song, K.; Yu, Z.; Zu, X.; Li, G.; Hu, Z.; Xue, Y. Collagen Remodeling along Cancer Progression Providing a Novel Opportunity for Cancer Diagnosis and Treatment. Int. J. Mol. Sci. 2022, 23, 10509. [Google Scholar] [CrossRef]
  93. Desa, D.E.; Wu, W.; Brown, R.M.; Brown, E.B., 4th; Hill, R.L.; Turner, B.M.; Brown, E.B., 3rd. Second-Harmonic Generation Imaging Reveals Changes in Breast Tumor Collagen Induced by Neoadjuvant Chemotherapy. Cancers 2022, 14, 857. [Google Scholar] [CrossRef]
  94. Venning, F.A.; Wullkopf, L.; Erler, J.T. Targeting ECM Disrupts Cancer Progression. Front. Oncol. 2015, 5, 224. [Google Scholar] [CrossRef] [PubMed]
  95. Nejim, Z.; Navarro, L.; Morin, C.; Badel, P. Quantitative analysis of second harmonic generated images of collagen fibers: A review. Res. Biomed. Eng. 2022, 39, 273–295. [Google Scholar] [CrossRef]
  96. Ávila, F.J.; Bueno, J.M. Analysis and quantification of collagen organization with the structure tensor in second harmonic microscopy images of ocular tissues. Appl. Opt. 2015, 54, 9848–9854. [Google Scholar] [CrossRef] [PubMed]
  97. Bueno, J.M.; Ávila, F.J.; Hristu, R.; Stanciu, S.G.; Eftimie, L.; Stanciu, G.A. Objective analysis of collagen organization in thyroid nodule capsules using second harmonic generation microscopy images and the Hough transform. Appl. Opt. 2020, 59, 6925. [Google Scholar] [CrossRef]
  98. Alizadeh, M.; Ghotbi, M.; Loza-Alvarez, P.; Merino, D. Comparison of Different Polarization Sensitive Second Harmonic Generation Imaging Techniques. Methods Protoc. 2019, 2, 49. [Google Scholar] [CrossRef]
  99. Yu, W.; Li, X.; Wang, B.; Li, Y.; Shen, B.; Qu, J.; Liu, L. Nonlinear polarization tensor measurement with a vectorial complex field in second-harmonic-generation microscopy. Phys. Rev. A 2023, 107, 013505. [Google Scholar] [CrossRef]
  100. Golaraei, A.; Kontenis, L.; Cisek, R.; Tokarz, D.; Done, S.J.; Wilson, B.C.; Barzda, V. Changes of collagen ultrastructure in breast cancer tissue determined by second-harmonic generation double Stokes-Mueller polarimetric microscopy. Biomed. Opt. Express 2016, 7, 4054–4068. [Google Scholar] [CrossRef]
  101. Tokarz, D.; Cisek, R.; Joseph, A.; Asa, S.L.; Wilson, B.C.; Barzda, V. Characterization of pathological thyroid tissue using polarization-sensitive second harmonic generation microscopy. Mod. Pathol. 2020, 100, 1280–1287. [Google Scholar] [CrossRef]
  102. Hompland, T.; Erikson, A.; Lindgren, M.; Lindmo, T.; Davies, C.d.L. Second-Harmonic Generation in Collagen as a Potential Cancer Diagnostic Parameter. J. Biomed. Opt. 2008, 13, 054050. [Google Scholar] [CrossRef]
  103. Rouède, D.; Schaub, E.; Bellanger, J.-J.; Ezan, F.; Scimeca, J.-C.; Baffet, G.; Tiaho, F. Determination of Extracellular Matrix Collagen Fibril Architectures and Pathological Remodeling by Polarization Dependent Second Harmonic Microscopy. Sci. Rep. 2017, 7, 12197. [Google Scholar] [CrossRef]
  104. Ambekar, R.; Lau, T.-Y.; Walsh, M.; Bhargava, R.; Toussaint, K.C. Quantifying Collagen Structure in Breast Biopsies Using Second-Harmonic Generation Imaging. Biomed. Opt. Express 2012, 3, 2021–2035. [Google Scholar] [CrossRef] [PubMed]
  105. Lin, J.; Pan, S.; Zheng, W.; Huang, Z. Polarization-Resolved Second-Harmonic Generation Imaging for Liver Fibrosis Assessment without Labeling. Appl. Phys. Lett. 2013, 103, 173701. [Google Scholar] [CrossRef]
  106. Watanabe, T.; Thayil, A.; Jesacher, A.; Grieve, K.; Debarre, D.; Wilson, T.; Booth, M.; Srinivas, S. Characterisation of the dynamic behaviour of lipid droplets in the early mouse embryo using adaptive harmonic generation microscopy. BMC Cell Biol. 2010, 11, 38. [Google Scholar] [CrossRef] [PubMed]
  107. Müller; Squier; Wilson; Brakenhoff 3D microscopy of transparent objects using third-harmonic generation. J. Microsc. 1998, 191, 266–274. [CrossRef] [PubMed]
  108. Chu, S.W.; Chen, S.Y.; Chern, G.W.; Tsai, T.H.; Chen, Y.C.; Lin, B.L.; Sun, C.K. Studies of chi(2)/chi(3) tensors in submicron-scaled bio-tissues by polarization harmonics optical microscopy. Biophys. J. 2004, 86, 3914–3922. [Google Scholar] [CrossRef] [PubMed]
  109. Gavgiotaki, E.; Filippidis, G.; Tsafas, V.; Bovasianos, S.; Kenanakis, G.; Georgoulias, V.; Tzardi, M.; Agelaki, S.; Athanassakis, I. Third Harmonic Generation microscopy distinguishes malignant cell grade in human breast tissue biopsies. Sci. Rep. 2020, 10, 11055. [Google Scholar] [CrossRef]
  110. Gavgiotaki, E.; Filippidis, G.; Markomanolaki, H.; Kenanakis, G.; Agelaki, S.; Georgoulias, V.; Athanassakis, I. Distinction between breast cancer cell subtypes using third harmonic generation microscopy. J. Biophotonics 2016, 10, 1152–1162. [Google Scholar] [CrossRef]
  111. Ardakani, A.A.; Ghader, A.; Asgari, H.; Keshavarz, M.; Tazehmahalleh, F.E.; Ara, M.H.M.; Malekzadeh, M.; Ghaznavi, H.; Shakeri-Zadeh, A. The capability of nonlinear optical characteristics as a predictor for cellular uptake of nanoparticles and cell damage. Photodiagn. Photodyn. Ther. 2019, 27, 442–448. [Google Scholar] [CrossRef]
  112. Combes, G.F.; Vučković, A.-M.; Bakulić, M.P.; Antoine, R.; Bonačić-Koutecky, V.; Trajković, K. Nanotechnology in Tumor Biomarker Detection: The Potential of Liganded Nanoclusters as Nonlinear Optical Contrast Agents for Molecular Diagnostics of Cancer. Cancers 2021, 13, 4206. [Google Scholar] [CrossRef]
  113. Sun, L.; Liu, H.; Ye, Y.; Lei, Y.; Islam, R.; Tan, S.; Tong, R.; Miao, Y.-B.; Cai, L. Smart nanoparticles for cancer therapy. Signal Transduct. Target. Ther. 2023, 8, 418. [Google Scholar] [CrossRef]
  114. Aneesa, V.M.; Safna, K.P.; Hussan, S.; Lekshmi, S.; Babu, T.D.; Muraleedharan, K. Analysis of non-linear optical properties of phytochemical photosensitizers in cancer photodynamic therapy by quantum computational. Results Chem. 2024, 8, 101580. [Google Scholar] [CrossRef]
  115. Hoque, E.; Biswas, M.K.; Hossain, M.; Bhattacharja, S.; Hasan, K.M.E.; Sharafuddin, S.M.; Das, S.K.; Haque, Y. Nonlinear optical phase shift in blood plasmas for neoplasia diagnosis. Opt. Express 2023, 31, 23056–23065. [Google Scholar] [CrossRef] [PubMed]
  116. Ghader, A.; Behruzi, M.; Ara, M.H.M.; Ghaznavi, H.; Ardakani, A.A. Nonlinear optical response of cancer cells following conventional and nano-technology based treatment strategies: Results of chemo-, thermo- and radiation therapies. Photodiagn. Photodyn. Ther. 2021, 37, 102686. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Raman scattering principle.
Figure 1. Raman scattering principle.
Optics 05 00031 g001
Figure 2. (a): Raman spectra of HepG2 cell line in lipids (blue), cytoplasm (green), and nucleus (red) regions. (b): Comparison of Raman spectra before (green) and after incubation of HepG2 with PLGA nanoparticles (blue). The pink spectrum corresponds to the isolated PLGA NPs. Image reproduced from [23].
Figure 2. (a): Raman spectra of HepG2 cell line in lipids (blue), cytoplasm (green), and nucleus (red) regions. (b): Comparison of Raman spectra before (green) and after incubation of HepG2 with PLGA nanoparticles (blue). The pink spectrum corresponds to the isolated PLGA NPs. Image reproduced from [23].
Optics 05 00031 g002
Figure 3. SRH images corresponding to a non-Hodgkin lymphoma specimen (A), non-small cell lung cancer brain metastasis (B), and glioblastoma (C) specimens (left column). Heat maps of CNN prediction algorithm (middle column) and overlay of the SRH images with heat maps are shown in the right column. Scale bar: 100 μm. Image reproduced from [30].
Figure 3. SRH images corresponding to a non-Hodgkin lymphoma specimen (A), non-small cell lung cancer brain metastasis (B), and glioblastoma (C) specimens (left column). Heat maps of CNN prediction algorithm (middle column) and overlay of the SRH images with heat maps are shown in the right column. Scale bar: 100 μm. Image reproduced from [30].
Optics 05 00031 g003
Figure 4. (a): CARS image of HepG2 cell line showing lipids (red), proteins (green), and nucleic acid (blue) regions. (b): associated CARS spectra. Image reproduced from [41].
Figure 4. (a): CARS image of HepG2 cell line showing lipids (red), proteins (green), and nucleic acid (blue) regions. (b): associated CARS spectra. Image reproduced from [41].
Optics 05 00031 g004
Figure 5. TPEF image of CAF (middle); monocellular HGT1 (labeled with eGFP) spheroids (left) and bicellular HGT1/CAF (right). Scale bar: 100 μm. Image reproduced from [55].
Figure 5. TPEF image of CAF (middle); monocellular HGT1 (labeled with eGFP) spheroids (left) and bicellular HGT1/CAF (right). Scale bar: 100 μm. Image reproduced from [55].
Optics 05 00031 g005
Figure 6. Lifetime TPEF image of a normal breast cell (a) and the normalized fluorescence decay signal (b). I(t), IRF(t)m, and F(t) correspond to the intensity of the decay signal, the impulse response function, and a model fitted, respectively. Reproduced from [68].
Figure 6. Lifetime TPEF image of a normal breast cell (a) and the normalized fluorescence decay signal (b). I(t), IRF(t)m, and F(t) correspond to the intensity of the decay signal, the impulse response function, and a model fitted, respectively. Reproduced from [68].
Optics 05 00031 g006
Figure 7. Illustration of the two-photon excitation fluorescence (a), second harmonic generation (b), and third harmonic generation (c) non-linear processes.
Figure 7. Illustration of the two-photon excitation fluorescence (a), second harmonic generation (b), and third harmonic generation (c) non-linear processes.
Optics 05 00031 g007
Figure 8. Second-order susceptibility values of normal (a) and tumor breast (b) tissues. Scale bar: 20 μm. Images reproduced from [100].
Figure 8. Second-order susceptibility values of normal (a) and tumor breast (b) tissues. Scale bar: 20 μm. Images reproduced from [100].
Optics 05 00031 g008
Figure 9. THG images of cells isolated from healthy tissue (left column) and for different grades of breast cancer (third to fourth columns). The top (A) and bottom (B) panels compare the morphology of the nucleus and nucleoli, respectively. Red and white arrows indicate the irregular nucleus and nucleoli, respectively. Scale bar: 2 μm. Images reproduced from [109].
Figure 9. THG images of cells isolated from healthy tissue (left column) and for different grades of breast cancer (third to fourth columns). The top (A) and bottom (B) panels compare the morphology of the nucleus and nucleoli, respectively. Red and white arrows indicate the irregular nucleus and nucleoli, respectively. Scale bar: 2 μm. Images reproduced from [109].
Optics 05 00031 g009
Table 1. Summary of the reviewed techniques, principles, contrast mechanisms, applications, and key features. RS: spontaneous Raman scattering; SRS: stimulated Raman scattering; CARS: coherent anti-Stokes Raman scattering; TPEF: two-photon excitation fluorescence; FLIM: fluorescence lifetime imaging microscopy; TAM: transient absorption microscopy; SHG: second harmonic generation; THG: third-harmonic generation.
Table 1. Summary of the reviewed techniques, principles, contrast mechanisms, applications, and key features. RS: spontaneous Raman scattering; SRS: stimulated Raman scattering; CARS: coherent anti-Stokes Raman scattering; TPEF: two-photon excitation fluorescence; FLIM: fluorescence lifetime imaging microscopy; TAM: transient absorption microscopy; SHG: second harmonic generation; THG: third-harmonic generation.
TechniquePrincipleContrast MechanismApplicationsKey Features
RSInelastic scattering of light, where energy is transferred to molecular vibrations.Chemical composition based on vibrational modes.Chemical imaging, material science, cancer diagnostics.
-
Weak signal, requires longer acquisition times.
-
Molecular specificity without labels.
-
Sensitive to chemical bonds.
SRSEnhanced Raman signal by stimulating Raman-active modes with pump and probe lasers.Chemical composition based on vibrational modesFast imaging of live cells, tissue imaging.
-
Higher sensitivity than spontaneous Raman.
-
Rapid, real-time imaging.
-
Requires synchronized lasers.
CARSNon-linear process where anti-Stokes shifted light is generated, enhancing Raman signal.Chemical composition based on vibrational modesLabel-free imaging, lipid mapping, biomedical imaging.
-
High sensitivity, fast imaging.
-
Non-resonant background can interfere with signals.
-
Complex instrumentation required.
TPEFNon-linear excitation where two photons are absorbed simultaneously to excite a fluorophore.Fluorescence emission from excited states Deep tissue imaging, neuroscience, cellular imaging.
-
Allows deep tissue penetration.
-
Reduces photodamage and photobleaching.
-
Requires pulsed lasers.
FLIMMeasures the decay time of fluorescence after two-photon excitation.Temporal decay of fluorescence signal, independent of intensity.Cellular metabolism, protein interactions, cancer diagnostics.
-
Provides functional information.
-
Requires complex lifetime analysis.
-
Useful for multiparametric imaging.
TAMMeasures the absorption of a probe pulse following excitation by a pump pulse.Differential absorption signal revealing electronic and vibrational dynamics.Study of ultrafast dynamics, excited state lifetimes, and charge transfer in materials and biological systems.
-
High temporal resolution.
-
Can study non-fluorescent molecules.
-
Requires ultrafast lasers and sophisticated analysis tools.
SHGNon-linear optical process where two photons combine to form a single photon with twice the energy.Structural information from non-centrosymmetric molecules (e.g., collagen).Imaging of collagen, muscle fibers, and other non-centrosymmetric structures.
-
Label-free imaging.
-
Specific to non-centrosymmetric structures.
-
No photobleaching or phototoxicity.
THGNon-linear optical process where three photons combine to form a single photon with three times the energy.Interface-sensitive imaging of boundaries between different refractive indices.Imaging of interfaces, cellular structures, and developmental biology.
-
Label-free, subcellular resolution.
-
Sensitive to interfaces and material properties.
-
Limited to specific sample types with clear interfaces.
Table 2. Main limitation and recommendations for future improvements of non-linear optical approaches for cancer detection.
Table 2. Main limitation and recommendations for future improvements of non-linear optical approaches for cancer detection.
LimitationRecommendation
High-costsCost reduction strategies, including the development of simplified designs based on cost-effective materials, open-sources manufacturing processes.
Technological equipmentDevelopment of robust, affordable, and user-friendly systems and data analysis software. Advancements in miniaturizing equipment.
Tissue penetration Enhance tissue penetration depth with the implementation of adaptive optics and the combination with multimodal approaches.
Data AnalysisIncorporating artificial intelligence for automatic image processing, pattern recognition, and data analysis to enhance the differentiation of cancerous tissues.
In vivo trialsMore in vivo studies and clinical trials are needed to demonstrate the efficacy and reliability of non-linear optical techniques in real-clinic cancer detection, to ensure broad applicability.
Multi-modal imagingCombining the contrast mechanisms of the reviewed techniques with other imaging techniques such as PET or MRI could offer a more comprehensive diagnostic approach.
Standardization protocolsIt is critical to establish standardized protocols for the use of non-linear optical techniques in cancer detection. This includes guidelines for image acquisition, data analysis, and interpretation of results.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ávila, F.J. A Review of Non-Linear Optical Imaging Techniques for Cancer Detection. Optics 2024, 5, 416-433. https://doi.org/10.3390/opt5040031

AMA Style

Ávila FJ. A Review of Non-Linear Optical Imaging Techniques for Cancer Detection. Optics. 2024; 5(4):416-433. https://doi.org/10.3390/opt5040031

Chicago/Turabian Style

Ávila, Francisco J. 2024. "A Review of Non-Linear Optical Imaging Techniques for Cancer Detection" Optics 5, no. 4: 416-433. https://doi.org/10.3390/opt5040031

Article Metrics

Back to TopTop