Fact-checked by Grok 2 weeks ago

Spectral radiance

Spectral radiance is a fundamental quantity in that quantifies the emitted, reflected, transmitted, or received by a surface per projected area, per , and per or at a specific point and . It is typically denoted as L_\lambda(\lambda, \theta, \phi) for dependence or L_\nu(\nu, \theta, \phi) for dependence, where \theta and \phi specify the . A common for spectral radiance, when expressed per , is the watt per per square meter per nanometer (W sr⁻¹ m⁻² nm⁻¹). The corresponding SI is the watt per per cubic meter (W sr⁻¹ m⁻³). In the context of thermal radiation, spectral radiance plays a central role in describing blackbody emission through Planck's law, which provides the spectral distribution of radiation from an ideal blackbody at temperature T. The wavelength form of Planck's law is given by
L_\lambda(\lambda, T) = \frac{2hc^2}{\lambda^5} \frac{1}{e^{hc / \lambda kT} - 1},
where h is Planck's constant, c is the speed of light, k is Boltzmann's constant, and \lambda is the wavelength; this formula resolves classical inconsistencies like the ultraviolet catastrophe and underpins modern quantum physics. For blackbodies, spectral radiance is isotropic and depends solely on temperature, serving as a universal reference for calibrating light sources and detectors across the electromagnetic spectrum from ultraviolet to infrared.
Spectral radiance is essential for characterizing the full radiometric properties of electromagnetic fields, including , , and other forms, enabling precise modeling of propagation and interaction with . In and physics applications, it is critical for selecting and optimizing sources in , where it determines energy coupling into small apertures or fibers, and in optical system design for imaging and illumination. It also finds widespread use in for analyzing stellar spectra, in for atmospheric and surface monitoring, and in standards maintained by institutions like NIST to ensure in measurements from 225 nm to 2400 nm.

Fundamentals

Definition

Spectral radiance is a fundamental radiometric quantity that quantifies the amount of emitted, reflected, transmitted, or received per unit perpendicular to the of propagation, per unit , and per unit or interval. It describes the distribution of radiative energy across the in a specific from a point on a surface, making it essential for characterizing sources and scenes in . Denoted as L_\nu in the or L_\lambda in the wavelength domain, spectral radiance provides a detailed spectral breakdown of the directional intensity of radiation. The mathematical definition of spectral radiance in terms of frequency is given by the differential expression L_\nu(\nu, \theta, \phi) = \frac{d^3\Phi}{dA \cos\theta \, d\Omega \, d\nu}, where \Phi represents the radiant flux (in watts), A is the differential area of the surface, \theta is the angle between the surface normal and the direction of propagation (zenith angle), \Omega is the differential solid angle (in steradians), and \nu is the frequency (in hertz). This formulation accounts for the projected area A \cos\theta to ensure the measurement is independent of the observer's orientation relative to the surface. An analogous expression exists for the wavelength domain, replacing d\nu with d\lambda and adjusting for the relationship \nu = c / \lambda, where c is the speed of light. Spectral radiance is distinct from related broadband quantities such as total radiance, which integrates L_\nu or L_\lambda over the entire to yield power per unit projected area per unit solid angle (in W/m²·sr), and from , which integrates over all directions in the and the to give power per unit area (in W/m²). Unlike , which lacks directional information, spectral radiance preserves both spectral and angular details, enabling precise modeling of light propagation in free space. The term spectral radiance emerged within to address the need for spectrally resolved descriptions of in fields like and . Its conceptual roots trace to 19th-century , particularly Gustav Kirchhoff's 1860 introduction of principles, which emphasized wavelength-dependent and , and Max Planck's 1900 derivation of the blackbody spectral radiance law, resolving the through quantum hypothesis.

Units and Notation

Spectral radiance is quantified in the (SI) using distinct forms depending on whether it is expressed per unit or per unit . In the , the spectral radiance L_\nu has units of watts per square meter per per hertz (W·m⁻²·⁻¹·Hz⁻¹). In the domain, the spectral radiance L_\lambda has units of watts per square meter per per meter (W·m⁻²·⁻¹·m⁻¹), though practical measurements often employ nanometers or micrometers, yielding W·m⁻²·⁻¹·nm⁻¹ or W·m⁻²·⁻¹·μm⁻¹, respectively. The relationship between these representations ensures conservation of radiant power across intervals, given by L_\nu = L_\lambda \cdot \frac{\lambda^2}{c}, where \lambda is the and c is the in (approximately 3 × 10⁸ m/s). This conversion arises from the differential relation d\nu = -\frac{c}{\lambda^2} d\lambda, maintaining L_\nu |d\nu| = L_\lambda |d\lambda|. Dimensionally, spectral radiance [L] is analyzed as power per unit projected area, per unit solid angle, per unit spectral interval, expressed as [L] = \frac{[\text{power}]}{[\text{area}] \cdot [\text{solid angle}] \cdot [\text{frequency}]} for the frequency form, or equivalently per wavelength. The projected area incorporates a \cos\theta factor, where \theta is the angle between the surface normal and the direction of propagation, accounting for the effective emitting area in the radiance definition L = \frac{d^3\Phi}{dA \cos\theta \cdot d\omega \cdot d\nu}, with \Phi as radiant power, A as area, \omega as solid angle, and \nu as frequency. Common notations for spectral radiance include L_{e,\nu} to denote emitted spectral radiance in the , emphasizing emitted power. Directional dependence is often indicated by subscripts or arguments, such as L(\omega) where \omega is the unit direction vector. In astronomy, particularly , is frequently expressed in janskys per (Jy/sr), where 1 Jy = 10⁻²⁶ ·m⁻²·Hz⁻¹, yielding units of 10⁻²⁶ ·m⁻²·Hz⁻¹·sr⁻¹. In contexts, units like watts per square centimeter per per micrometer (·cm⁻²·sr⁻¹·μm⁻¹) are prevalent for practical and photometry. In practical applications, such as , the notation aligns with for spectral radiance B_\nu(T) = \frac{2h\nu^3}{c^2} \frac{1}{e^{h\nu / kT} - 1}, where h is Planck's constant, k is Boltzmann's constant, and T is temperature in ; this form uses the frequency-domain units W·m⁻²·sr⁻¹·Hz⁻¹ to describe the ideal emitter's output.

Key Properties

Invariance Under Coordinate Transformations

Spectral radiance L_\nu, the radiant flux per unit area perpendicular to the direction of , per unit , and per unit , exhibits invariance under translations and rotations in . In relativistic settings, the quantity L_\nu / \nu^3 remains invariant under Lorentz transformations, accounting for Doppler shifts in \nu and the transformation of s and energies, ensuring consistency across inertial frames. This conservation arises from the underlying principles of number and energy preservation in free space, where no , , or occurs. In non-relativistic contexts, translations and rotations preserve L_\nu along ray paths due to the geometric nature of . The derivation of this invariance stems from in classical , applied to the six-dimensional of and for . states that the phase-space volume element d^3\mathbf{x} \, d^3\mathbf{p} is conserved along trajectories in the absence of collisions, as the flow is incompressible. For , the f(\mathbf{x}, \mathbf{p}), proportional to the spectral radiance, satisfies the collisionless , leading to constant f along rays. Integrating over the appropriate differentials, this implies that the quantity d^3\Phi / (dA \cos\theta \, d\Omega \, d\nu) = L_\nu remains constant, where d^3\Phi is the differential flux, dA the area element, \theta the angle to the normal, d\Omega the , and d\nu the interval. In relativistic extensions, the invariance of the phase-space volume under Lorentz transformations further ensures L_\nu / \nu^3 constancy. This invariance has profound physical implications for optical systems, enabling the preservation of brightness—the perceived intensity per unit area and —in devices such as and projectors, where travels through free space or lossless media. For instance, a can collect over a larger without altering the source's spectral radiance, limited only by and aberrations. In contrast, E, the per unit area, decreases as $1/r^2 with distance r due to geometric spreading, highlighting radiance's role as a conserved "" measure. A practical example is the observation of : the spectral radiance from a reaches the observer unchanged in , ignoring interstellar , allowing direct inference of the stellar .

Reciprocity Principle

The reciprocity principle in states that, for passive linear media, the spectral radiance propagating from a source at position \mathbf{r}_s in direction \hat{\mathbf{n}}_s to a detector at \mathbf{r}_d in direction \hat{\mathbf{n}}_d equals the spectral radiance propagating in the reverse direction: L_\nu(\mathbf{r}_s, \hat{\mathbf{n}}_s \to \mathbf{r}_d, \hat{\mathbf{n}}_d) = L_\nu(\mathbf{r}_d, \hat{\mathbf{n}}_d \to \mathbf{r}_s, \hat{\mathbf{n}}_s). This theorem, an extension of Helmholtz reciprocity from wave to radiometric quantities, holds for and applies to the across frequencies. The mathematical basis derives from the reciprocity property of the electromagnetic scattering operator, rooted in field theory principles such as and time-reversal , which ensure symmetric propagation in lossless or absorbing passive media. This extends to bidirectional reflectance distribution functions (BRDFs), where the exchanged quantity is the specific intensity, equivalent to spectral radiance in radiometry. In imaging systems, the principle ensures symmetry in point spread functions (PSFs), constraining the radiation field such that the beam spread function exhibits reciprocal equality between forward and reverse directions, which aids in modeling light propagation through turbid media. It is also fundamental to radiometric , allowing source and detector positions to be interchanged without altering the measured response, thereby validating thermodynamic consistency in enclosure-based setups. The principle breaks down in active media, where time-varying elements like amplifiers introduce nonreciprocity by violating time-reversal invariance, leading to asymmetric field ratios upon source-detector exchange. For polarized light, extensions incorporate or to maintain reciprocity, though violations can occur in chiral or magneto-optic materials.

Optical Theorems and Applications

Étendue and Brightness Conservation

The étendue, denoted as G, quantifies the throughput of light in an optical system and is defined as G = n^2 A \Omega, where n is the of the medium, A is the cross-sectional area perpendicular to the beam, and \Omega is the subtended by the beam. In lossless optical systems, étendue is conserved, meaning it remains constant along the propagation path despite transformations in area or angle by lenses, mirrors, or other elements. This conservation arises directly from the invariance of spectral radiance under coordinate transformations, as the product n^2 L_\nu (where L_\nu is the spectral radiance) remains unchanged, ensuring that the total \Phi_\nu = L_\nu A \Omega is preserved for uniform sources. The theorem, a key consequence of étendue conservation, states that no passive optical system can increase the of beyond that of , with spectral radiance satisfying L_\nu \leq L_{\nu, \text{source}} at all points. This leads to fundamental limits on concentration: the maximum concentration factor C_{\max} for transferring from an input medium with refractive index n_{\text{in}} and acceptance half-angle \theta_{\text{in}} to an output medium with n_{\text{out}} and maximum emission half-angle \theta_{\text{out}} (typically approaching 90°) is given by C_{\max} = \left( \frac{n_{\text{out}}}{n_{\text{in}}} \right)^2 \left( \frac{\sin \theta_{\text{out}}}{\sin \theta_{\text{in}}} \right)^2. This limit derives from the conserved throughput integral \int L_\nu \cos \theta \, dA \, d\Omega = \text{constant}, which, under the assumption of invariant L_\nu for a uniform source, reduces to invariance and imposes constraints on concentrator design. Specifically, matching at input and output ensures maximum flux transfer without exceeding source brightness, preventing thermodynamic violations in energy concentration. In practical applications, étendue conservation governs the design of solar concentrators, where it sets the theoretical upper bound on sunlight intensity at photovoltaic cells—for instance, achieving up to 46,000 times concentration for the sun's angular of 0.267° in air, though real systems fall short due to imperfections. Similarly, in fiber optics, étendue determines the maximum light acceptance via the \text{NA} = n \sin \theta, limiting coupling efficiency from extended sources to the fiber core area times product.

Collimated Beam Analysis

In a , composed of nearly rays, the spectral radiance L_\nu is uniform across the beam's cross-section for an ideal case without diffraction effects. This uniformity arises because all rays propagate in the same direction, concentrating the radiant flux within a minimal . In practice, however, the effective \Omega subtended by the beam is not exactly zero but finite, typically on the order of microradians, due to inherent and any residual beam imperfections. During propagation through free space, the spectral radiance of a remains conserved in the absence of , , or other losses. This conservation principle ensures that L_\nu at any point along the equals its initial value, as the optical throughput is preserved in lossless media. , however, gradually increases the 's étendue, limiting the extent to which the can be focused without loss of . For a diffraction-limited , the far-field half-angle \theta is approximated by \theta \approx \frac{\lambda}{\pi w_0}, where \lambda is the wavelength and w_0 is the beam waist radius; a common rough estimate uses the beam diameter D \approx 2w_0, yielding \theta \approx \lambda / D. Measurement of spectral radiance in collimated beams from lasers or LEDs typically involves integrating spheres to capture the total radiant flux, coupled with a spectrometer for wavelength resolution, allowing L_\nu to be computed from the flux, beam area, and estimated solid angle. Alternatively, goniophotometers or goniospectroradiometers provide detailed angular profiles of L_\nu by scanning the beam's directionality, essential for characterizing near-collimated sources where the emission is highly directional. These methods account for the beam's small divergence to ensure accurate radiance values. A prominent application is in spectroscopy, where the exceptionally high spectral radiance of collimated beams—often exceeding $10^6 W m^{-2} sr^{-1} Hz^{-1}—enables precise and detection of atomic or molecular transitions with minimal . This contrasts sharply with divergent sources like lamps, which have orders-of-magnitude lower spectral radiance due to their extended source areas and wide emission solid angles, making them unsuitable for high-resolution spectroscopic tasks.

Ray-Based Descriptions

In geometric ray optics, spectral radiance L_\nu is conceptualized as a quantity propagated along individual rays of light, where each ray in a lossless, homogeneous medium carries an invariant value of L_\nu directed along its path. This invariance arises from the conservation of energy in the ray's propagation, ensuring that the power per unit area perpendicular to the ray, per unit solid angle, and per unit frequency remains constant unless altered by absorption, emission, or scattering. For a bundle of closely parallel rays subtending a differential solid angle d\Omega, the effective spectral radiance is the average value over the bundle, computed as the total power divided by the product of the projected area and d\Omega, providing a local measure of directional intensity. When tracing rays through interfaces between media with different refractive indices, dictates the change in ray direction, while the spectral radiance adjusts to maintain the invariance of L_\nu / n^2, where n is the . For a ray refracting from medium 1 (index n_1) to medium 2 (index n_2), the transmitted spectral radiance is given by L_{\nu,2} = L_{\nu,1} \left( \frac{n_2}{n_1} \right)^2, reflecting the compression or expansion of the subtended by the ray bundle due to the bending at the interface; this holds for lossless without . Along the ray path s within a medium, the describes potential variations as dL_\nu / ds = -\kappa_\nu L_\nu + j_\nu, where \kappa_\nu is the coefficient and j_\nu the emission term, though in ideal geometric tracing through non-absorbing media, dL_\nu / ds = 0, preserving constancy. This ray-based framework finds extensive application in , where ray tracing algorithms compute spectral radiance maps by backward-tracing rays from the observer through scene geometries to light sources, enabling realistic rendering of illumination and color spectra. The RADIANCE synthetic imaging system exemplifies this, using stochastic ray tracing to simulate and generate high-fidelity radiance distributions for architectural visualization. In illumination engineering, similar techniques trace ray bundles to optimize light distribution in optical systems, ensuring conservation of L_\nu for efficient design of luminaires and displays. The geometric ray approximation underlying these descriptions holds only for scales much larger than the radiation wavelength, where diffraction and interference effects are negligible; at smaller scales comparable to the wavelength, wave optics must be invoked to account for phenomena like spreading and phase coherence.

Mathematical Formulations

Spectral Radiance in Frequency and Wavelength Domains

Spectral radiance can be expressed in either the frequency domain, denoted as L_\nu(\nu), or the wavelength domain, denoted as L_\lambda(\lambda), where \nu is the frequency and \lambda is the wavelength related by \nu = c / \lambda with c the speed of light in vacuum. The two representations describe the same physical quantity but differ in how the radiance is distributed over the spectral variable, ensuring that the energy in a differential interval is invariant: L_\lambda(\lambda) \, d\lambda = L_\nu(\nu) \, d\nu. Since d\nu = -(c / \lambda^2) \, d\lambda, the magnitude relation yields L_\lambda(\lambda) = L_\nu(\nu) \cdot (c / \lambda^2), highlighting that radiance values in the wavelength domain scale inversely with the square of the wavelength for a given energy interval. This transformation is crucial for accurate spectral analysis, as neglecting the Jacobian factor c / \lambda^2 leads to errors in peak positions and integrated intensities when switching domains. For , the spectral radiance follows in both domains. In the , the Planck function is B_\nu(\nu, T) = \frac{2 h \nu^3}{c^2} \frac{1}{e^{h \nu / k T} - 1}, where h is Planck's constant, k is Boltzmann's constant, and T is the . In the domain, it becomes B_\lambda(\lambda, T) = \frac{2 h c^2}{\lambda^5} \frac{1}{e^{h c / \lambda k T} - 1}. These expressions are related by the transformation above, and integrating either over all frequencies or wavelengths gives the total radiance B(T) = \sigma T^4 / \pi, where \sigma = 5.670374419 \times 10^{-8} \, \mathrm{W \cdot m^{-2} \cdot K^{-4}} is the Stefan-Boltzmann constant. The peak of the spectral radiance shifts with according to : in , \lambda_\mathrm{max} T \approx 2898 \, \mu\mathrm{m \cdot K}; in frequency, \nu_\mathrm{max} / T \approx 5.879 \times 10^{10} \, \mathrm{Hz/K}. This shift arises because the \lambda^{-5} term in B_\lambda biases the peak toward longer wavelengths compared to a direct conversion from the frequency form. In practice, the choice of domain depends on the application. The is preferred for phenomena involving quantum effects, such as physics, where energy levels are proportional to \nu via E = h \nu, allowing direct assessment of linewidths in hertz (e.g., lasers with linewidths of 1–10 MHz). Conversely, the domain is standard in visible and , where detector responses and material absorption features are calibrated in nanometers or micrometers. A numerical example of conversion errors occurs for a 300 K blackbody: the wavelength is at \lambda_\mathrm{max} \approx 9.66 \, \mu\mathrm{m} with B_\lambda \approx 1.29 \times 10^7 \, \mathrm{W \cdot m^{-2} \cdot sr^{-1} \cdot \mu\mathrm{m}^{-1}}, corresponding to \nu \approx 3.10 \times 10^{13} \, \mathrm{Hz}; however, the true is at \nu_\mathrm{max} \approx 1.76 \times 10^{13} \, \mathrm{Hz}, and omitting the \lambda^{-2} factor in conversion can lead to significant errors in the radiance value at this point, distorting flux calculations in approximations. In , particularly , the domain predominates due to its alignment with atmospheric transmission windows and / signatures tabulated in units. The IEEE 4001 standard, finalized in 2025, emphasizes in terms of distinguishable s for hyperspectral cameras, facilitating of spectral contrast in without the nonlinear scaling issues of conversion. This domain choice minimizes errors in retrieving surface properties from data, where frequency-domain processing is reserved for advanced signal analysis like Fourier transforms rather than primary radiance measurements.

Alternative Derivations and Approximations

In wave optics, spectral radiance can be derived from the and the degree of for electromagnetic fields. The time-averaged \langle \mathbf{S} \rangle = \frac{1}{\mu_0} \langle \mathbf{E} \times \mathbf{B} \rangle provides the spectral irradiance (energy flux per unit and area), approximated for a monochromatic in a medium as \frac{|E|^2}{2 \eta Z_0} (with |E| the amplitude, \eta the , and Z_0 the free-space impedance). For partially coherent , radiance is obtained by integrating field correlations over the and dividing by the , linking theory to radiometric quantities. For , approximations simplify the full Planck spectrum in specific regimes. In the low-frequency Rayleigh-Jeans limit, where h\nu \ll kT, the spectral radiance B_\nu(T) \approx \frac{2\nu^2 kT}{c^2}, treating as classical with equipartition of among modes. At high frequencies, where h\nu \gg kT, Wien's approximation yields B_\nu(T) \approx \frac{2h\nu^3}{c^2} e^{-h\nu / kT}, emphasizing the exponential cutoff due to thermal occupation probabilities. Pre-quantum approaches provided foundational alternatives for spectral radiance. Josef Stefan's 1879 empirical law described total radiance as \sigma T^4, later theoretically derived by in 1884 using and , assuming blackbody equilibrium without spectral detail. Wilhelm Wien's 1893 displacement law and 1896 approximation posited a spectral form B_\lambda(T) \propto \frac{1}{\lambda^5} f(\lambda T), fitting short-wavelength data but diverging at long wavelengths, predating quantum corrections. In modern computational radiometry, ray tracing simulates spectral radiance in complex scenes by stochastically tracing photon paths, accounting for , absorption, and emission to compute integrated radiance fields. This method excels for non-uniform geometries, as in atmospheric or material simulations, where analytic solutions fail. Recent advancements include neural radiance fields (), introduced in 2020, which parameterize scenes as continuous functions outputting view-dependent spectral radiance and density from sparse images, enabling high-fidelity reconstruction for rendering and analysis.

References

  1. [1]
    [PDF] Spectral radiance calibrations - NIST Technical Series Publications
    Spectral radiance, denoted L^, is defined as the radiant flux at a given point, direction, and wavelength per unit of projected area, solid angle, and ...
  2. [2]
    Radiometry - RP Photonics
    Radiometry is the science and technology of quantifying and measuring properties of electromagnetic radiation, including visible, infrared and ultraviolet ...
  3. [3]
    Understanding Radiance (Brightness), Irradiance and Radiant Flux
    Definitions of Radiometry and Photometry ... Radiometry is the science of measuring radiation energy in any portion of the electromagnetic spectrum. In practice, ...
  4. [4]
    Blackbody Radiation - The Physics Hypertextbook
    E = hf · Planck's law is a formula for the spectral radiance of an object at a given temperature as a function of frequency (Lf) or wavelength (L · When these ...
  5. [5]
    [PDF] Radiometry Terms and Definitions - Purdue Engineering
    Radiance (from Extended Source). (θ: angle from surface normal). L φ/(ωAcos(θ)) dφ/(dωdAcos(θ)). Watt / (sr m2). Spectral Radiance. Lλ. dL/dλ. Watt / (sr m3)
  6. [6]
    [PDF] Photometry and Radiometry - UBC Computer Science
    Radiance is measured in watts per square meter per steradian. 3.4.1 Spectral Radiance. Spectral radiance is radiance per unit wavelength interval at wavelength ...
  7. [7]
    [PDF] TOPICAL REVIEW Quantum radiometry
    May 20, 2009 · Gustav Kirchhoff introduced the term 'blackbody' and described this phenomenon in 1860. The derivation and physical understanding of blackbody ...<|control11|><|separator|>
  8. [8]
    [PDF] APPENDIX I THE SI SYSTEM AND SI UNITS FOR RADIOMETRY ...
    ... spectral irradiance (W/m2-nm), or luminance (cd/m2) and spectral radiance (W/m2-sr-nm). This equation represents a weighting, wavelength by wavelength, of ...
  9. [9]
    Measurement of Spectral Radiance - Bentham
    Spectral radiance in a given direction, at a given point of a real or imaginary surface is defined by the equation: Symbol: Ll. Unit: W∙m-2∙sr-1∙nm-1.Missing: L_nu | Show results with:L_nu
  10. [10]
    cosmolopy.magnitudes
    Units are: erg s^-1 cm^-2 Hz^-1 for f_nu and erg s^-1 cm^-2 Ang^-1 for f_lambda . Works because f_nu = f_lambda * lambda**2/c and f_lambda = f_nu * nu**2/c , ...Missing: spectral radiance
  11. [11]
    Definitions and some radiation theory - Pyspectral's documentation!
    With the spectral radiance given as a function of wavelength the equation looks like this: T_B = T(B_{\lambda}) = \frac{hc}. In python it may look like this ...Missing: L_nu | Show results with:L_nu
  12. [12]
    2 Radiation Fundamentals‣ Essential Radio Astronomy
    Spectral luminosity is an intrinsic property of the source because it does not depend on the distance d between the source and the observer—the d2 in Equation.
  13. [13]
    [PDF] radiation.pdf
    Planck's Law describes the amplitude of radiation emitted (i.e., spectral radiance) from a black body. It is generally provided in one of two forms; ...
  14. [14]
    [PDF] The Geometry of Radiative Transfer - arXiv
    Jul 29, 2013 · This provides a modern rationale for the classical law of “conservation of radiance along a ray” [36]. It also reveals a surprising similarity ...
  15. [15]
    [PDF] Radiance
    The spectral radiant emissivity (A)=Wx/W,BB0/0λ. Hence, the subscript notation e, which could be confused with e/?, is not recommended, although it is often ...
  16. [16]
    [PDF] HIGH ENERGY ASTROPHYSICS - Lecture 4
    Remember: Iν/ν3 was invariant under Lorentz transformation (Iν ≡ I(ν) the specific intensity). • Thus, observed intensity of a moving blob: I(νobs) ν3 obs.
  17. [17]
  18. [18]
    Hidden symmetry of the beam spread function resulting from the reciprocity theorem
    ### Summary: Reciprocity Theorem and Symmetry in Beam Spread Functions (BSF)
  19. [19]
    Thermodynamic constraints on reflectance reciprocity and Kirchhoff’s law
    ### Summary of Reciprocity in Radiometry from https://opg.optica.org/ao/abstract.cfm?uri=ao-37-16-3464
  20. [20]
    Electromagnetic Nonreciprocity
    ### Key Points on Limitations of Reciprocity in Active Media
  21. [21]
    [PDF] Étendue
    Étendue is one of the most basic yet important concepts in the design of nonimaging and illumination optics. First, it explains the flux transfer ...
  22. [22]
    Etendue and Optical Invariants - Eclipse Optics
    Jun 4, 2024 · This is why the definition says that Etendue can not decrease in a lossless system. Let's consider a simple example, the system in figure 4.
  23. [23]
    [PDF] Section 11 Radiative Transfer
    The spectral characteristics of the optical system (source spectrum, transmission and detector responsivity) must be considered in radiometric calculations.
  24. [24]
    [PDF] Intensity, Brightness and´Etendue of an Aperture Lamp 1 Problem
    Of course, Liouville's theorem and the related conservation of étendue give no insight as to how brightness can be enhanced in situations to which they don't ...
  25. [25]
    Technical Note: Etendue and Optical Throughput Calculations
    Etendue (G) describes the ability of a source to emit light or the ability of an optical system to accept light. For a monochromator, its étendue of accepting ...
  26. [26]
    Radiance - RP Photonics
    λ is the spectral radiance with respect to wavelength; the same can be done with optical frequencies. The units of ; λ are W cm−2 sr−1 nm−1, and those of ; ν are ...Missing: L_nu | Show results with:L_nu
  27. [27]
    Beam Divergence – angle - RP Photonics
    For a diffraction-limited Gaussian beam, the 1 / e 2 beam divergence half-angle is λ / ( π w 0 ) , where λ is the wavelength (in the medium) and w 0 the beam ...Quantitative Definitions of... · Divergence of Gaussian... · Spatial Fourier Transforms
  28. [28]
    [PDF] Basic Optics: Radiance Outline - eng . lbl . gov
    In general, the basic radiance, defined by: L/N2 of a narrow beam of radiation is conserved as the beam propagates through any loss-less optical system.
  29. [29]
    Integrating Sphere Fundamentals and Applications - Newport
    An integrating sphere and calibrated detector setup is suitable for accurate, absolute value light power measurement of divergent beams from laser diodes, ...
  30. [30]
    Goniophotometry versus Goniospectroradiometry - Pro-Lite ...
    Abstract. In recent years, the goniometer based upon a spectrometer (the “goniospectroradiometer”) has become a popular choice for measuring the output.
  31. [31]
    Basic Radiance and Radiance Invariance - SPIE Digital Library
    If rays are traced along a lossless boundary between two media having different indices of refraction, the solid angle changes according to Snell's law.
  32. [32]
    The RADIANCE 6.0 Synthetic Imaging System
    RADIANCE was developed as a research tool for predicting the distribution of visible radiation in illuminated spaces.
  33. [33]
    [PDF] Conservative Radiance Interpolants for Ray Tracing
    Abstract. Classical ray-tracing algorithms compute radiance returning to the eye along one or more sample rays through each pixel of an image.
  34. [34]
    Different forms of Planck's law | tec-science
    Feb 19, 2020 · Conversion of the wavelength form into the frequency form of the spectral ... wavelength to a frequency by the formula f=c/λ. This is ...
  35. [35]
    and Frequency-Dependent Formulations of Wien's Displacement Law
    Interrelations between the wavelength- and frequency-dependent formulations of Wien's displacement laws have been derived from the corresponding energy ...
  36. [36]
    The IEEE 4001 Standard for Hyperspectral Imaging
    This extensive work has resulted in the IEEE 4001 standard, which as of April 2025 is in the final stages of approval by IEEE. This standard defines a ...Missing: frequency wavelength
  37. [37]
    About Hyperspectral Remote Sensing Data | NSF NEON
    Sep 27, 2023 · Hyperspectral data consists of many bands from the electromagnetic spectrum, collected by imaging spectrometers, with NEON data collected ...Missing: standards choice<|control11|><|separator|>
  38. [38]
    [PDF] A Concise History of the Black-body Radiation Problem - arXiv
    Aug 12, 2022 · Along with Robert Bunsen, Kirchhoff studied the coincidence of the emission lines of sodium with particular absorption lines in the solar ...
  39. [39]
    (PDF) Monte Carlo method in optical radiometry - ResearchGate
    Aug 6, 2025 · In the ratio method, spectral radiance ratios to the pilot and geometric means were used to calculate the difference from the RV and its ...
  40. [40]
    A Path‐Tracing Monte Carlo Library for 3‐D Radiative Transfer in ...
    Jul 15, 2019 · This library is devoted to the acceleration of ray tracing in complex data, typically high-resolution large-domain grounds and clouds.
  41. [41]