Near-infrared window in biological tissue

Last updated

The near-infrared (NIR) window (also known as optical window or therapeutic window) defines the range of wavelengths from 650 to 1350 nanometre (nm) where light has its maximum depth of penetration in tissue. [1] Within the NIR window, scattering is the most dominant light-tissue interaction, and therefore the propagating light becomes diffused rapidly. Since scattering increases the distance travelled by photons within tissue, the probability of photon absorption also increases. Because scattering has weak dependence on wavelength, the NIR window is primarily limited by the light absorption of blood at short wavelengths and water at long wavelengths. The technique using this window is called NIRS. Medical imaging techniques such as fluorescence image-guided surgery often make use of the NIR window to detect deep structures.

Contents

Absorption properties of tissue components

The absorption coefficient () is defined as the probability of photon absorption in tissue per unit path length. [2] Different tissue components have different values. Moreover, is a function of wavelength. Discussed below are the absorption properties of the most important chromophores in tissue. The molar extinction coefficient () is another parameter that is used to describe photon absorption in tissue. By multiplying by the molar concentration and by ln(10), one can convert to .

Figure 1: The molar extinction coefficients of HbO2 and Hb. Fig 1 - The molar extinction coefficient of HbO2 and Hb.png
Figure 1: The molar extinction coefficients of HbO2 and Hb.

Blood

Blood consists of two different types of hemoglobin: oxyhemoglobin () is bound to oxygen, while deoxyhemoglobin () is unbound to oxygen. These two different types of hemoglobin exhibit different absorption spectra that are normally represented in terms of molar extinction coefficients, as shown in Figure 1. The molar extinction coefficient of Hb has its highest absorption peak at 420 nm and a second peak at 580 nm. Its spectrum then gradually decreases as light wavelength increases. On the other hand, shows its highest absorption peak at 410 nm, and two secondary peaks at 550 nm and 600 nm. As light wavelengths passes 600 nm, absorption decays much faster than Hb absorption. The points where the molar extinction coefficient spectra of and intersect are called isosbestic points.

By using two different wavelengths, it is possible to calculate the concentrations of oxyhemoglobin () and deoxyhemoglobin () as shown in the following equations:

Figure 2: The absorption spectrum of water. Fig 2 - The absorption spectrum of water.png
Figure 2: The absorption spectrum of water.

Here, and are the two wavelengths; and are the molar extinction coefficients of and , respectively; and are the molar concentrations of and in tissue, respectively. Oxygen saturation () can then be computed as

Water

Although water is nearly transparent in the range of visible light, it becomes absorbing over the near-infrared region. Water is a critical component since its concentration is high in human tissue. The absorption spectrum of water in the range from 250 to 1000 nm is shown in Figure 2. Although absorption is rather low in this spectral range, it still contributes to the overall attenuation of tissue.

Figure 3: The molar extinction coefficients of eumelanin and pheomelanin. Molar extinction coefficients of eumelanin and pheomelanin.png
Figure 3: The molar extinction coefficients of eumelanin and pheomelanin.

Other tissue components with less significant contributions to the total absorption spectrum of tissue are melanin and fat.

Figure 4: The absorption coefficient spectrum of fat. The absorption coefficient spectrum of fat.png
Figure 4: The absorption coefficient spectrum of fat.

Melanin

Melanin is a chromophore that exists in the human epidermal layer of skin responsible for protection from harmful UV radiation. When melanocytes are stimulated by solar radiation, melanin is produced. [7] Melanin is one of the major absorbers of light in some biological tissue (although its contribution is smaller than other components). There are two types of melanin: eumelanin which is black-brown and pheomelanin which is red-yellow. [8] The molar extinction coefficient spectra corresponding to both types are shown in Figure 3.

Fat

Fat is one of the major components in tissue that can comprise 10–40% of tissue. Although not many mammalian fat spectra are available, Figure 4 shows an example extracted from pig fat. [9]

Figure 5: The scattering coefficient spectrum of biological tissue. The scattering coefficient spectrum of biological tissue.png
Figure 5: The scattering coefficient spectrum of biological tissue.

Scattering properties of tissue components

Optical scattering occurs due to mismatches in refractive index of the different tissue components, ranging from cell membranes to whole cells. Cell nuclei and mitochondria are the most important scatterers. [11] Their dimensions range from 100 nm to 6 μm, and thus fall within the NIR window. Most of these organelles fall in the Mie scattering regime, and exhibit highly anisotropic forward-directed scattering. [12]

Light scattering in biological tissue is denoted by the scattering coefficient (), which is defined as the probability of photon scattering in tissue per unit path length. [13] Figure 5 shows a plot of the scattering spectrum. [14]

Effective attenuation coefficient

Attenuation of light in deep biological tissue depends on the effective attenuation coefficient (), which is defined as

where is the transport scattering coefficient defined as

where is the anisotropy of biological tissue, which has a representative value of 0.9. Figure 5 shows a plot of transport scattering coefficient spectrum in breast tissue, which has a wavelength dependence of . [15] The effective attenuation coefficient is the dominant factor for determining light attenuation at depth ≫ 1/ .

Estimation

The NIR window can be computed based on the absorption coefficient spectrum or the effective attenuation coefficient spectrum. A possible criterion for selecting the NIR window is given by the FWHM of the inverse of these spectra as shown in Figure 7.

In addition to the total concentration of hemoglobin, the oxygen saturation will define the concentration of oxy- and deoxyhemoglobin in tissue and so the total absorption spectrum. Depending on the type of tissue, we can consider different situations. Below, the total concentration of hemoglobin is assumed to be 2.3 mM.

Arteries absorption.png
Figure 6 (a): Spectra for arteries (SaO2 ≈ 98%).

Absorption coefficient: λmin = 686 nm; NIR window = (634–756) nm.

Effective attenuation coefficient: λmin = 690 nm; NIR window = (618–926) nm.
Veins absorption.png
Figure 6 (b): Spectra for veins (SvO2 ≈ 60%).

Absorption coefficient: λmin = 730 nm; NIR window = (664–932) nm.

Effective attenuation coefficient: λmin = 730 nm; NIR window = (630–1328) nm.
Breast absorption.png
Figure 6 (c): Spectra for breast tissue (StO2 ≈ 70%).

Absorption coefficient: λmin = 730 nm; NIR window = (656–916) nm.

Effective attenuation coefficient: λmin = 730 nm; NIR window = (626–1316) nm.

Absorption spectrum for arteries

In this case ≈ 98% (arterial oxygen saturation). Then oxyhemoglobin will be dominant in the total absorption (black) and the effective attenuation (magenta) coefficient spectra, as shown in Figure 6 (a). 'cite: Anisotropic diffusion filter for dorsal hand vein features extraction – Sarah Hachemi Benziane, Abdelkader Benyettou'

Absorption spectrum for veins

In this case ≈ 60% (venous oxygen saturation). Then oxyhemoglobin and deoxyhemoglobin will have similar contributions to the total absorption (black) and the effective attenuation (magenta) coefficient spectra, as shown in Figure 6 (b).

Figure 7: : Effective penetration depth in breast tissue (StO2 [?] 70%). Effective attenuation coefficient: lmin = 730 nm; NIR window = (626-1316) nm. Effective penetration depth.PNG
Figure 7: : Effective penetration depth in breast tissue (StO2 ≈ 70%). Effective attenuation coefficient: λmin = 730 nm; NIR window = (626–1316) nm.

Absorption spectrum for breast tissue

To define (tissue oxygen saturation) (or (tissue saturation index)), it is necessary to define a distribution of arteries and veins in tissue. an arterial-venous blood volume ratio of 20%/80% can be adopted. [16] Thus tissue oxygen saturation can be defined as = 0.2 x + 0.8 x ≈ 70%.

The total absorption (black) and the effective attenuation (magenta) coefficient spectra for breast tissue is shown in Figure 6 (c). In addition, the effective penetration depth is plotted in Figure 7.

See also

Related Research Articles

The Beer–Lambert law is commonly applied to chemical analysis measurements to determine the concentration of chemical species that absorb light. It is often referred to as Beer's law. In physics, the Bouguer–Lambert law is an empirical law which relates the extinction or attenuation of light to the properties of the material through which the light is travelling. It had its first use in astronomical extinction. The fundamental law of extinction is sometimes called the Beer–Bouguer–Lambert law or the Bouguer–Beer–Lambert law or merely the extinction law. The extinction law is also used in understanding attenuation in physical optics, for photons, neutrons, or rarefied gases. In mathematical physics, this law arises as a solution of the BGK equation.

In physics, the cross section is a measure of the probability that a specific process will take place in a collision of two particles. For example, the Rutherford cross-section is a measure of probability that an alpha particle will be deflected by a given angle during an interaction with an atomic nucleus. Cross section is typically denoted σ (sigma) and is expressed in units of area, more specifically in barns. In a way, it can be thought of as the size of the object that the excitation must hit in order for the process to occur, but more exactly, it is a parameter of a stochastic process.

<span class="mw-page-title-main">Refractive index</span> Ratio of the speed of light in vacuum to that in the medium

In optics, the refractive index of an optical medium is a dimensionless number that gives the indication of the light bending ability of that medium.

<span class="mw-page-title-main">Optical depth</span> Physics concept

In physics, optical depth or optical thickness is the natural logarithm of the ratio of incident to transmitted radiant power through a material. Thus, the larger the optical depth, the smaller the amount of transmitted radiant power through the material. Spectral optical depth or spectral optical thickness is the natural logarithm of the ratio of incident to transmitted spectral radiant power through a material. Optical depth is dimensionless, and in particular is not a length, though it is a monotonically increasing function of optical path length, and approaches zero as the path length approaches zero. The use of the term "optical density" for optical depth is discouraged.

The propagation constant of a sinusoidal electromagnetic wave is a measure of the change undergone by the amplitude and phase of the wave as it propagates in a given direction. The quantity being measured can be the voltage, the current in a circuit, or a field vector such as electric field strength or flux density. The propagation constant itself measures the dimensionless change in magnitude or phase per unit length. In the context of two-port networks and their cascades, propagation constant measures the change undergone by the source quantity as it propagates from one port to the next.

Circular dichroism (CD) is dichroism involving circularly polarized light, i.e., the differential absorption of left- and right-handed light. Left-hand circular (LHC) and right-hand circular (RHC) polarized light represent two possible spin angular momentum states for a photon, and so circular dichroism is also referred to as dichroism for spin angular momentum. This phenomenon was discovered by Jean-Baptiste Biot, Augustin Fresnel, and Aimé Cotton in the first half of the 19th century. Circular dichroism and circular birefringence are manifestations of optical activity. It is exhibited in the absorption bands of optically active chiral molecules. CD spectroscopy has a wide range of applications in many different fields. Most notably, UV CD is used to investigate the secondary structure of proteins. UV/Vis CD is used to investigate charge-transfer transitions. Near-infrared CD is used to investigate geometric and electronic structure by probing metal d→d transitions. Vibrational circular dichroism, which uses light from the infrared energy region, is used for structural studies of small organic molecules, and most recently proteins and DNA.

<span class="mw-page-title-main">Wavenumber</span> Spatial frequency of a wave

In the physical sciences, the wavenumber, also known as repetency, is the spatial frequency of a wave, measured in cycles per unit distance or radians per unit distance. It is analogous to temporal frequency, which is defined as the number of wave cycles per unit time or radians per unit time.

<span class="mw-page-title-main">Mie scattering</span> Scattering of an electromagnetic plane wave by a sphere

In electromagnetism, the Mie solution to Maxwell's equations describes the scattering of an electromagnetic plane wave by a homogeneous sphere. The solution takes the form of an infinite series of spherical multipole partial waves. It is named after German physicist Gustav Mie.

A Newtonian fluid is a fluid in which the viscous stresses arising from its flow are at every point linearly correlated to the local strain rate — the rate of change of its deformation over time. Stresses are proportional to the rate of change of the fluid's velocity vector.

Absorbance is defined as "the logarithm of the ratio of incident to transmitted radiant power through a sample ". Alternatively, for samples which scatter light, absorbance may be defined as "the negative logarithm of one minus absorptance, as measured on a uniform sample". The term is used in many technical areas to quantify the results of an experimental measurement. While the term has its origin in quantifying the absorption of light, it is often entangled with quantification of light which is “lost” to a detector system through other mechanisms. What these uses of the term tend to have in common is that they refer to a logarithm of the ratio of a quantity of light incident on a sample or material to that which is detected after the light has interacted with the sample.

<span class="mw-page-title-main">Transmittance</span> Effectiveness of a material in transmitting radiant energy

In optical physics, transmittance of the surface of a material is its effectiveness in transmitting radiant energy. It is the fraction of incident electromagnetic power that is transmitted through a sample, in contrast to the transmission coefficient, which is the ratio of the transmitted to incident electric field.

<span class="mw-page-title-main">Functional near-infrared spectroscopy</span> Optical technique for monitoring brain activity

Functional near-infrared spectroscopy (fNIRS) is an optical brain monitoring technique which uses near-infrared spectroscopy for the purpose of functional neuroimaging. Using fNIRS, brain activity is measured by using near-infrared light to estimate cortical hemodynamic activity which occur in response to neural activity. Alongside EEG, fNIRS is one of the most common non-invasive neuroimaging techniques which can be used in portable contexts. The signal is often compared with the BOLD signal measured by fMRI and is capable of measuring changes both in oxy- and deoxyhemoglobin concentration, but can only measure from regions near the cortical surface. fNIRS may also be referred to as Optical Topography (OT) and is sometimes referred to simply as NIRS.

<span class="mw-page-title-main">Absorption cross section</span> Mmeasures the probability of an absorption process

In physics, absorption cross section is a measure for the probability of an absorption process. More generally, the term cross section is used in physics to quantify the probability of a certain particle-particle interaction, e.g., scattering, electromagnetic absorption, etc. Typical absorption cross section has units of cm2⋅molecule−1. In honor of the fundamental contribution of Maria Goeppert Mayer to this area, the unit for the two-photon absorption cross section is named the "GM". One GM is 10−50 cm4⋅s⋅photon−1.

The linear attenuation coefficient, attenuation coefficient, or narrow-beam attenuation coefficient characterizes how easily a volume of material can be penetrated by a beam of light, sound, particles, or other energy or matter. A coefficient value that is large represents a beam becoming 'attenuated' as it passes through a given medium, while a small value represents that the medium had little effect on loss. The (derived) SI unit of attenuation coefficient is the reciprocal metre (m−1). Extinction coefficient is another term for this quantity, often used in meteorology and climatology. Most commonly, the quantity measures the exponential decay of intensity, that is, the value of downward e-folding distance of the original intensity as the energy of the intensity passes through a unit thickness of material, so that an attenuation coefficient of 1 m−1 means that after passing through 1 metre, the radiation will be reduced by a factor of e, and for material with a coefficient of 2 m−1, it will be reduced twice by e, or e2. Other measures may use a different factor than e, such as the decadic attenuation coefficient below. The broad-beam attenuation coefficient counts forward-scattered radiation as transmitted rather than attenuated, and is more applicable to radiation shielding. The mass attenuation coefficient is the attenuation coefficient normalized by the density of the material.

<span class="mw-page-title-main">Mass attenuation coefficient</span> Property of materials

The mass attenuation coefficient, or mass narrow beam attenuation coefficient of a material is the attenuation coefficient normalized by the density of the material; that is, the attenuation per unit mass. Thus, it characterizes how easily a mass of material can be penetrated by a beam of light, sound, particles, or other energy or matter. In addition to visible light, mass attenuation coefficients can be defined for other electromagnetic radiation, sound, or any other beam that can be attenuated. The SI unit of mass attenuation coefficient is the square metre per kilogram. Other common units include cm2/g and L⋅g−1⋅cm−1. Mass extinction coefficient is an old term for this quantity.

When an electromagnetic wave travels through a medium in which it gets attenuated, it undergoes exponential decay as described by the Beer–Lambert law. However, there are many possible ways to characterize the wave and how quickly it is attenuated. This article describes the mathematical relationships among:

Surface-extended X-ray absorption fine structure (SEXAFS) is the surface-sensitive equivalent of the EXAFS technique. This technique involves the illumination of the sample by high-intensity X-ray beams from a synchrotron and monitoring their photoabsorption by detecting in the intensity of Auger electrons as a function of the incident photon energy. Surface sensitivity is achieved by the interpretation of data depending on the intensity of the Auger electrons instead of looking at the relative absorption of the X-rays as in the parent method, EXAFS.

Free carrier absorption occurs when a material absorbs a photon, and a carrier is excited from an already-excited state to another, unoccupied state in the same band. This intraband absorption is different from interband absorption because the excited carrier is already in an excited band, such as an electron in the conduction band or a hole in the valence band, where it is free to move. In interband absorption, the carrier starts in a fixed, nonconducting band and is excited to a conducting one.

In chemistry, the molar absorption coefficient or molar attenuation coefficient is a measurement of how strongly a chemical species absorbs, and thereby attenuates, light at a given wavelength. It is an intrinsic property of the species. The SI unit of molar absorption coefficient is the square metre per mole, but in practice, quantities are usually expressed in terms of M−1⋅cm−1 or L⋅mol−1⋅cm−1. In older literature, the cm2/mol is sometimes used; 1 M−1⋅cm−1 equals 1000 cm2/mol. The molar absorption coefficient is also known as the molar extinction coefficient and molar absorptivity, but the use of these alternative terms has been discouraged by the IUPAC.

Spectroscopic optical coherence tomography (SOCT) is an optical imaging and sensing technique, which provides localized spectroscopic information of a sample based on the principles of optical coherence tomography (OCT) and low coherence interferometry. The general principles behind SOCT arise from the large optical bandwidths involved in OCT, where information on the spectral content of backscattered light can be obtained by detection and processing of the interferometric OCT signal. SOCT signal can be used to quantify depth-resolved spectra to retrieve the concentration of tissue chromophores, characterize tissue light scattering, and/or used as a functional contrast enhancement for conventional OCT imaging.

References

  1. Smith, Andrew M.; Mancini, Michael C.; Nie, Shuming (2009). "Bioimaging: Second window for in vivo imaging". Nature Nanotechnology. 4 (11): 710–711. Bibcode:2009NatNa...4..710S. doi:10.1038/nnano.2009.326. ISSN   1748-3387. PMC   2862008 . PMID   19898521.
  2. LV. Wang and HI. Wu, Biomedical Optics. Wiley. ISBN   978-0-471-74304-0, 2007.
  3. Molar extinction coefficients of oxy and deoxyhemoglobin compiled by Scott Prahl. URL: http://omlc.ogi.edu/spectra/hemoglobin.
  4. G. M. Hale, and M. R. Querry, Optical constants of water in the 200 nm to 200 μm wavelength region, Appl. Opt., 12, 555–563, 1973.
  5. Extinction coefficient of melanin by Steven Jacques. URL: http://omlc.ogi.edu/spectra/melanin/extcoeff.html.
  6. R.L.P. van Veen, H.J.C.M. Sterenborg, A. Pifferi, A. Torricelli, and R. Cubeddu, OSA Annual BIOMED Topical Meeting, 2004.
  7. T. Vo-Dinh, Biomedical Photonics Handbook. Taylor & Francis, Inc. ISBN   0-8493-1116-0, 2002.
  8. George Zonios and Aikaterini Dimou, Ioannis Bassukas, Dimitrios Galaris, and Argyrios Ysolakidis and Efthimios Kaxiras, J. Biomed. Opt., Vol.13, 014017, 2008.
  9. R.L.P. van Veen, H.J.C.M. Sterenborg, A. Pifferi, A. Torricelli, and R. Cubeddu, OSA Annual BIOMED Topical Meeting, 2004.
  10. S. Jacques, C. Newman, D. Levy, and A. von Eschenbach. Univ. of Texas M. D. Anderson Cancer Center, 1987.
  11. LV. Wang and HI. Wu, Biomedical Optics. Wiley. ISBN   978-0-471-74304-0, 2007.
  12. T. Vo-Dinh, Biomedical Photonics Handbook. Taylor & Francis, Inc. ISBN   0-8493-1116-0, 2002.
  13. LV. Wang and HI. Wu, Biomedical Optics. Wiley. ISBN   978-0-471-74304-0, 2007.
  14. S. Jacques, C. Newman, D. Levy, and A. von Eschenbach. Univ. of Texas M. D. Anderson Cancer Center, 1987.
  15. S. Srinivasan, B. Pogue, S. Jiang, H. Dehghani, C. Kogel, S. Soho, J. Gibson, T. Tosteson, S. Poplack, and K. Paulsen, K D 2003, Proc Natl Acad. Sci. USA 100 12349 54.
  16. S. Nioka, S. Wen, J. Zhang, J. Du, X. Intes, Z. Zhao, and B. Chance, Simulation study of breast tissue hemodynamics during pressure perturbation. Oxygen Transport to Tissue XXVI 566, 17–22, 2006.