Fact-checked by Grok 2 weeks ago

Spectral line

A spectral line is a narrow, distinct bright or dark feature in an , appearing against a continuous background of and resulting from the or of photons at precise wavelengths due to quantum transitions in atoms, ions, or molecules. These lines arise primarily from the quantized levels of electrons in atoms or molecules, as described by . In emission spectra, electrons excited to higher states release photons of specific energies when transitioning to lower states, producing bright lines at wavelengths given by the difference in levels, such as \Delta E = h\nu, where h is Planck's constant and \nu is the . Conversely, lines form when photons from a continuum source are absorbed by cooler gas, exciting electrons and creating dark gaps in the spectrum at those same characteristic wavelengths. The exact positions and intensities of spectral lines depend on factors like the involved, ionization state, , , and , leading to broadening effects such as Doppler, natural, or broadening. Spectral lines serve as unique signatures, or "fingerprints," for identifying chemical elements and molecules in various environments, from samples to distant astronomical objects. In astronomy, they enable precise measurements of , , , and radial velocities via Doppler shifts, which reveal motions such as galactic rotations or cosmic expansion through . Historically observed in as dark since the early 19th century, these features underpin modern , , and , facilitating applications in , plasma diagnostics, and fundamental tests of .

Basic Principles

Definition and Characteristics

A spectral line is a narrow or broadened feature in the that corresponds to the emission or absorption of light at a specific due to transitions between quantized levels in atoms, ions, or molecules./05%3A_Radiation_and_Spectra/5.05%3A_Formation_of_Spectral_Lines) These lines appear as discrete peaks or dips in an otherwise uniform spectrum, reflecting the quantized nature of atomic and molecular energy states. Key characteristics of spectral lines include their position, intensity, sharpness, and occasional polarization. The position is defined by the precise wavelength \lambda or frequency \nu of the line, which directly relates to the energy difference \Delta E between the involved quantum levels via the equation E = \frac{hc}{\lambda}, where h is Planck's constant and c is the speed of light. Intensity depends on the transition probability and the relative populations of the energy levels, determining the line's brightness or depth. In the ideal case, a spectral line is infinitely sharp, resembling a delta function \delta(\nu - \nu_0) at the central frequency \nu_0, though observed lines exhibit finite width. Polarization may occur in lines influenced by magnetic fields or anisotropic conditions./05%3A_Radiation_and_Spectra/5.05%3A_Formation_of_Spectral_Lines) The historical observation of spectral lines dates to Newton's 1666 experiments, where he used prisms to separate into a continuous color without resolving discrete features. In 1814, first identified dark lines in the solar , cataloging over 500 such features and establishing them as standards for wavelength measurement. These observations laid the groundwork for . Spectral lines differ fundamentally from continuous spectra, which display a smooth distribution of all wavelengths without gaps. Lines originate from quantum transitions in sparse gases or low-density media, whereas continuous spectra arise from thermal emission in hot, dense bodies () or free-free processes like in plasmas./05%3A_Radiation_and_Spectra/5.05%3A_Formation_of_Spectral_Lines)

Formation Mechanisms

Spectral lines form through and molecular processes involving the or of photons at discrete wavelengths corresponding to transitions. In the process, atoms or molecules in excited states return to lower energy states, releasing photons with energies equal to the difference between the levels. This occurs via , as seen in where light excites atoms, leading to visible re-emission, or in thermal incandescence of low-density gases where collisions populate excited states. In contrast, absorption lines arise when photons from a continuum source interact with atoms or molecules in their , exciting them to higher levels and removing specific wavelengths from the , producing dark lines against a bright background. This mechanism requires a cooler, intervening gas relative to the hotter source, such as in stellar atmospheres where photospheric imprints lines on the emerging radiation. Temperature plays a key role by determining the population of excited states through the Boltzmann distribution, with higher temperatures increasing the fraction of atoms in upper levels and thus enhancing emission line strengths. Density influences collision rates, which can excite atoms to higher states in hot, dense environments or de-excite them via inelastic collisions in higher-density regimes, affecting the overall line formation efficiency. These processes are codified in Kirchhoff's laws of spectroscopy, formulated in 1860, which relate spectral types to source conditions: a hot, dense body like a solid or liquid produces a continuous spectrum; a hot, low-density gas yields bright emission lines; and a cooler gas overlying a hot continuum source results in absorption lines. The relative populations of upper (n_u) and lower (n_l) energy levels in thermal equilibrium follow the Boltzmann equation: \frac{n_u}{n_l} = \frac{g_u}{g_l} \exp\left( -\frac{\Delta E}{kT} \right) where g_u and g_l are the degeneracies of the upper and lower levels, \Delta E = E_u - E_l is the energy difference, k is the Boltzmann constant, and T is the temperature. To derive this, consider a system of non-interacting atoms in local thermodynamic equilibrium under Maxwell-Boltzmann statistics. The probability of an atom occupying a specific quantum state with energy E is proportional to \exp(-E / kT), reflecting the entropic maximization of configurations at fixed energy. For discrete energy levels, each level i has degeneracy g_i, the number of accessible states at energy E_i, so the population n_i is n_i \propto g_i \exp(-E_i / kT). Normalizing to the total population and taking the ratio for two levels yields the equation, assuming the partition function cancels out. This distribution underpins line intensities, as emission or absorption rates scale with n_u or n_l.

Classification and Types

Emission Lines

Emission lines manifest as bright, discrete features superimposed on a predominantly dark spectral background, arising from the emission of photons during or ionic transitions in excited . These lines are primarily produced through , where electrons in higher energy states decay to lower states, releasing photons of specific wavelengths, or via in the presence of a field. Such processes commonly occur in low-density plasmas or gases, where collisional de-excitation is minimal, allowing radiative decay to dominate. Prominent observational examples include the of , which features emission lines in the resulting from transitions to the n=2 principal quantum level. The most intense of these is the H-alpha line at 656.3 nm, appearing as a vivid red feature in spectra from ionized regions. In astrophysical contexts, such as the , emission spectra reveal forbidden lines—like those from doubly ionized oxygen ([O III]) at 495.9 nm and 500.7 nm—which are characteristic of low-density environments where densities are below approximately 10^6 cm^{-3}, suppressing collisional of metastable states. The intensity of an emission line is directly proportional to the population of atoms or ions in the upper and the probability, quantified by the Einstein A (A_{ul}), which represents the rate in s^{-1}. The radiative rate, determining the line's emitted photon flux per unit volume, is expressed as \Gamma = A_{ul} \, N_{\rm upper}, where N_{\rm upper} is the of particles in the upper state; this relation underpins of spectral data. lines find critical applications in diagnostics, enabling inference of physical parameters like and from line ratios, as well as in laser spectroscopy for real-time characterization of excited in plasmas.

Absorption Lines

Absorption lines manifest as dark features superimposed on a spectrum, arising when photons from a background source are by atoms or ions in cooler intervening material. This occurs at discrete wavelengths corresponding to quantum transitions from lower to upper levels, selectively removing those specific frequencies from the incident . The excited atoms subsequently re-emit the energy isotropically or through cascades to other levels, leading to a net reduction in intensity at the original wavelength rather than a directional contribution. In stellar spectra, absorption lines form primarily in the cooler outer layers of a star's atmosphere, where gas temperatures allow population of ground or low-lying states that intercept the continuum radiation emerging from the hotter interior. A classic example is the observed in sunlight, first cataloged in the early ; prominent among them are the calcium and lines at approximately 396.8 nm and 393.4 nm, respectively, produced by resonance transitions in singly ionized calcium (Ca II). These lines probe the solar photosphere and , with their depths reflecting local abundance and temperature conditions. Similarly, interstellar absorption lines appear in the spectra of distant stars, where foreground neutral or ionized gas along the absorbs continuum light; common features include the Ca II and lines and sodium D lines (Na I at 589.0 nm and 589.6 nm), which trace diffuse interstellar clouds and their column densities. The strength of an line is quantified by its depth and , which together indicate the amount of absorbing material. The line depth represents the fractional reduction in intensity at the line center, while the W_\lambda, defined as the integral of the normalized profile over , equals the width of a hypothetical rectangular dip with the same total absorbed flux. For optically thin lines, W_\lambda \approx \sigma_\lambda N f, where \sigma_\lambda is the wavelength-dependent cross-section, N is the column of absorbers (N = \int n \, dl, with n the and dl the path length), and f is the of the transition, a measure of its intrinsic probability. This relation allows astronomers to infer atomic abundances and physical conditions from observed spectra, though saturation effects in stronger lines require the full -of-growth analysis to accurately recover N. Absorption lines are observed across diverse contexts, including stellar atmospheres where they reveal elemental compositions and velocity fields; planetary atmospheres, such as those of or exoplanets, which imprint signatures on transmitted ; and laboratory , where controlled conditions enable precise measurements of parameters. The underlying process is described by the monochromatic \alpha_\nu = \frac{h\nu}{4\pi} B_{lu} n_l \phi(\nu), where h is Planck's constant, \nu the , B_{lu} the Einstein for , n_l the in the lower , and \phi(\nu) the normalized line profile function with \int \phi(\nu) \, d\nu = 1. Integrating \alpha_\nu along the yields the , which governs the observed line profile. These features contrast with emission lines formed in the same transitions but under conditions favoring net addition, such as in hotter, optically thin plasmas.

Band Spectra

Band spectra in molecular consist of series of closely spaced spectral lines arising from simultaneous changes in vibrational and rotational quantum numbers during electronic, vibrational, or pure rotational in molecules. Unlike the discrete, isolated lines observed in spectra, these lines form shaded or banded regions due to the dense packing of rotational levels within each vibrational , often appearing as continuous bands at lower resolution. The internal structure of a molecular band typically features P, Q, and R branches, corresponding to rotational changes of ΔJ = -1, 0, and +1, respectively. The P branch forms on the low-wavenumber side of the band origin, the R branch on the high-wavenumber side, and the Q branch, when allowed, clusters near the origin; in many diatomic cases, such as Σ–Σ transitions, the Q branch is absent due to selection rules. Vibrational progressions manifest as sequences of bands from Δv > 0 transitions, where higher vibrational levels produce successively weaker bands approaching the limit. The energy levels determining these bands for diatomic molecules are described by the anharmonic oscillator with rotation, with wavenumber given by \sigma(v, J) = \omega_e \left(v + \frac{1}{2}\right) - \omega_e x_e \left(v + \frac{1}{2}\right)^2 + B J(J+1) where \omega_e is the harmonic vibrational frequency, \omega_e x_e the anharmonicity constant, v the vibrational , B the rotational constant, and J the rotational ; centrifugal and other corrections may apply for precision. Prominent examples include the Swan bands of the C₂ molecule (d³Π_g – a³Π_u electronic system), observed in carbon-rich stars and flames with strong features around 5165 and 4737 , revealing molecular abundance and temperature. The violet system (B²Σ⁺ – X²Σ⁺) appears in cometary atmospheres and stellar spectra, exemplified by the (0,0) band near 3883 , aiding in diagnosing excitation conditions. Band widths depend on the rotational temperature, which governs the of J levels and thus the extent of populated branches, while high-vibrational bands truncate near molecular energies.

Theoretical Foundations

Quantum Mechanical Basis

The quantum mechanical basis of spectral lines originates from the quantization of atomic and molecular energy levels, which replaced classical models with discrete states leading to sharp emission or absorption features. In 1913, introduced a semi-classical model for the , postulating that electrons occupy stationary orbits with quantized L = n \hbar, where n is a positive integer and \hbar = h / 2\pi, preventing continuous energy loss via radiation. Transitions between these levels were assumed to emit or absorb photons with energy \Delta E = h \nu, explaining the discrete lines in spectra. Although successful for , the failed for multi-electron atoms and lacked a relativistic or wave description, paving the way for full . The foundational quantum mechanical treatment of atomic energy levels came from solving the time-independent for the in 1926, yielding exact discrete energy eigenvalues E_n = -\frac{13.6 \, \text{eV}}{n^2} for n = 1, 2, \dots, with bound states below the ionization threshold at E = 0 and a above it. These levels arise from the radial and angular solutions of \hat{H} \psi = E \psi, where \hat{H} = -\frac{\hbar^2}{2\mu} \nabla^2 - \frac{e^2}{4\pi \epsilon_0 r} for \mu and potential. For multi-electron atoms, the exact many-body becomes intractable due to electron-electron interactions, so the Hartree-Fock method approximates the wavefunction as a of single-particle orbitals, solving self-consistent field equations variationally to obtain approximate energy levels that capture exchange effects and correlate well with observed spectra for light atoms. In molecules, energy levels are more complex due to nuclear motion, addressed by the Born-Oppenheimer approximation in , which exploits the mass disparity between electrons and nuclei to separate the total wavefunction into electronic, vibrational, and rotational parts: the electronic is solved for fixed nuclear positions to yield surfaces, on which nuclei vibrate and rotate. This yields discrete electronic transitions split into vibrational (via anharmonic potentials) and rotational (via levels) substructure, forming band spectra rather than isolated lines. Above the dissociation or ionization limit, levels merge into continua, analogous to atomic cases. Spectral lines emerge from transitions between these discrete levels induced by electromagnetic perturbations, described by . The transition rate from initial state |i\rangle to a continuum of final states |f\rangle is given by : \Gamma = \frac{2\pi}{\hbar} |\langle f | \hat{H}' | i \rangle|^2 \rho(E_f), where \hat{H}' is the interaction (e.g., operator -\vec{\mu} \cdot \vec{E}) and \rho(E_f) is the of final states at energy E_f = E_i + \hbar \omega. This first-order formula, derived from expanding the time evolution operator, quantifies the probability of emission or , with line positions determined by \Delta E and widths influenced by level , though discrete levels inherently produce sharp lines in isolation.

Selection Rules and Transitions

Selection rules govern which quantum transitions between levels result in observable spectral lines, arising from the laws and properties in the interaction between matter and under the electric approximation. These rules determine the allowed changes in quantum numbers for transitions, with the electric term dominating for permitted lines due to its relatively strong coupling. In atomic systems, the primary selection rules for electric (E1) transitions are Δl = ±1 for the orbital angular momentum quantum number of the , ΔS = 0 to conserve total , and ΔJ = 0, ±1 for the total angular momentum quantum number, excluding the case of J = 0 to J = 0. For molecular spectra, selection rules differ based on the involved. In vibrational transitions within the approximation, the change in vibrational is restricted to Δv = ±1, enabling transitions while are weaker. Electronic transitions in diatomic molecules follow ΔΛ = 0 for Σ-Σ transitions and more generally ΔΛ = 0, ±1, with additional constraints from , such as g ↔ u for heteronuclear diatomics or + ↔ - for certain symmetries in homonuclear cases. Transitions violating electric dipole selection rules are termed forbidden and proceed via weaker mechanisms like (M1) or electric (E2) interactions, resulting in much lower transition probabilities and intensities observable primarily in low-density environments such as nebulae. A prominent example is the [O II] forbidden lines at 372.6 nm and 372.9 nm, arising from M1 and E2 transitions in singly ionized oxygen, which are key diagnostics for ionized regions in planetary nebulae./07%3A_Atomic_Spectroscopy/7.25%3A_Some_forbidden_lines_worth_knowing) The strength of an allowed transition is quantified by the dimensionless oscillator strength f, which measures the transition probability relative to a classical electron oscillator and is given by f = \frac{8\pi^2 m_e \nu}{3 h e^2} |\mu|^2, where m_e is the electron mass, ν is the transition frequency, h is Planck's constant, e is the elementary charge, and μ is the electric dipole transition moment. Oscillator strengths typically range from 10^{-3} to 1 for strong lines, providing a direct link between theoretical models and observed line intensities. In cases where electronic transitions are symmetry-forbidden, can be borrowed from nearby allowed through vibronic coupling, as described by the Herzberg-Teller mechanism, where vibrational modes mix states and induce a non-zero transition dipole. This effect is particularly relevant for vibronically allowed bands in polyatomic molecules, enhancing otherwise weak spectral features via second-order perturbation.

Nomenclature and Measurement

Wavelength Designation

Spectral lines are designated by their wavelengths, which specify the position of the line in the . For , the most prominent series are named based on the principal of the lower involved in the transition. The corresponds to transitions from higher levels (n ≥ 2) to n = 1, producing lines in the region. The involves transitions to n = 2 from higher levels (n ≥ 3), resulting in lines primarily in the . The Paschen series features transitions to n = 3 from higher levels (n ≥ 4), with lines in the infrared region. For other elements, spectral lines follow analogous series notations where applicable, or are labeled by specific multiplet structures. In , lines from neutral helium (He I) are often grouped into series similar to hydrogen's , while the Pickering series designates prominent lines from singly ionized helium (He II), such as those observed in hot stellar atmospheres. Iron lines, for instance, are cataloged using multiplet tables that group transitions by term symbols, with individual lines identified by their approximate wavelength, such as the neutral iron line Fe I at 4045 . Wavelengths are conventionally reported in air or , with a correction applied to convert between them due to the of air. The approximate shift for standard conditions (P = 1 atm) is \Delta \lambda / \lambda \approx 2.7 \times 10^{-4}, which scales linearly with P (in atm) at constant temperature. More precise conversions account for using formulas like that of Ciddor (1996). wavelengths are preferred in space-based observations to avoid atmospheric effects. In , spectral lines are denoted using a notation that specifies the and state (e.g., He I for neutral , Fe II for singly ionized iron), followed by the wavelength in angstroms (). This system facilitates identification in astronomical spectra. For enhanced precision in spectroscopic analysis, particularly in the and , wavelengths are often expressed as s \sigma = 1/\lambda in units of cm^{-1}, which directly relate to energy and allow easier interpolation in line lists. This unit is in databases like those from the National Institute of Standards and Technology (NIST).

Intensity and Profile Notation

The of a spectral line is often characterized by measures such as peak height, which indicates the maximum deviation from the , or integrated , representing the total energy absorbed or emitted across the line. A widely used metric is the W, which quantifies the line strength independently of the level and broadening effects; it is defined as the width of a rectangular (or ) feature with the same area as the actual line relative to the . Mathematically, W = \int_{-\infty}^{\infty} \left(1 - \frac{I(\lambda)}{I_c}\right) d\lambda, where I(\lambda) is the observed intensity at wavelength \lambda and I_c is the continuum intensity; W is typically expressed in angstroms (Å) or nanometers (nm)./11%3A_Curve_of_Growth/11.01%3A_Introduction_to_Curve_of_Growth) Spectral line profiles describe the shape and distribution of intensity across the line, influenced by various broadening mechanisms. The Gaussian profile arises primarily from Doppler broadening due to thermal motions, given by G(x) = \frac{1}{\sqrt{\pi} \Delta \nu_D} \exp\left( -\left(\frac{x}{\Delta \nu_D}\right)^2 \right), where x = \nu - \nu_0 is the frequency offset from the line center \nu_0 and \Delta \nu_D is the Doppler width. The Lorentzian profile, associated with natural and pressure broadening, has the form L(x) = \frac{\Gamma / 2\pi}{(x)^2 + (\Gamma / 2)^2}, where \Gamma is the full width at half maximum (FWHM). In many cases, the observed profile is a Voigt function, the convolution of Gaussian and Lorentzian components, V(x, a) = \frac{a}{\pi} \int_{-\infty}^{\infty} \frac{\exp(-y^2)}{(x - y)^2 + a^2} \, dy, with the damping parameter a = \frac{\Gamma}{4\pi \Delta \nu_D}; this hybrid shape features Gaussian-like cores and Lorentzian wings. The intrinsic strength of a transition is denoted by the oscillator strength f, a dimensionless parameter related to the transition probability; in atomic databases, lines are commonly listed with \log gf, where g = 2J + 1 is the statistical weight of the lower level and J its total angular momentum quantum number. The NIST Atomic Spectra Database provides critically evaluated \log gf values alongside wavelengths and intensities for thousands of transitions, enabling comparisons of line strengths across elements. The relationship between equivalent width and the column density N of the absorbing or emitting species is described by the curve of growth, which illustrates how W increases with N. For optically thin lines (low N), W grows linearly with N as W \propto N f \lambda, where \lambda is the , since is unsaturated. As N increases, the line core saturates, causing W to grow more slowly in a square-root regime dominated by Doppler effects, before flattening in the damping regime where Lorentzian wings contribute disproportionately, yielding W \propto \sqrt{N} or logarithmic behavior at high N. This curve is essential for inferring abundances from observed spectra without resolving individual profiles./11%3A_Curve_of_Growth/11.01%3A_Introduction_to_Curve_of_Growth)

Line Broadening and Shifts

Natural and Lifetime Broadening

Natural broadening, also known as lifetime broadening, arises intrinsically from the quantum mechanical uncertainty in the energy of excited atomic states due to their finite lifetimes. According to the Heisenberg uncertainty principle, the product of the uncertainty in energy \Delta E and the uncertainty in time \Delta t satisfies \Delta E \Delta t \geq \frac{\hbar}{2}, where \hbar is the reduced Planck's constant. For an excited state with lifetime \tau, the time uncertainty is on the order of \tau, leading to an energy uncertainty \Delta E \approx \frac{\hbar}{\tau}. This manifests as a broadening of the spectral line, with the frequency distribution following a Lorentzian profile. The full width at half maximum (FWHM) of this profile, denoted \Gamma, is given by \Gamma = \frac{1}{\tau} in angular frequency units, reflecting the exponential decay of the excited state amplitude. The lifetime \tau of an excited state is determined by the total decay rate, primarily dominated by in dilute gases or isolated atoms, with rate A_{ul} for transition from upper level u to lower level l. Additional contributions include non-radiative decay processes, such as collisional quenching or , and if the lower level is significantly populated. The total decay rate is thus $1/\tau = A_{ul} + \sum A_{\text{non-rad}} + B_{ul} \rho(\nu), where B_{ul} is the stimulated emission coefficient and \rho(\nu) is the radiation density at \nu; however, in typical laboratory or astrophysical conditions for natural broadening, prevails. The Einstein relations connect these rates: for transitions between levels of equal degeneracy, A_{ul} = \frac{8\pi h \nu^3}{c^3} B_{ul}, linking the spontaneous emission probability to the stimulated processes derived from equilibrium. This relation, originally formulated by Einstein, underscores the fundamental role of in line formation. In , the natural linewidth originates from the interaction of the atomic dipole with vacuum fluctuations of the , which provide the that triggers . The Weisskopf-Wigner theory formalizes this by treating the atom as coupled to a of modes, resulting in the lineshape as the of the decaying wavefunction. For the Lyman-alpha transition (1s-2p at 121.6 nm), the natural linewidth is approximately 100 MHz, corresponding to the 2p state's lifetime of about 1.6 dominated by with A_{ul} \approx 6.3 \times 10^8 s^{-1}. This intrinsic width is typically much smaller than Doppler or pressure broadenings in practical spectra, but it sets the fundamental limit observable in ultra-high-resolution experiments, such as of .

Doppler and Thermal Broadening

Doppler broadening of spectral lines occurs due to the relative velocities between the source and the observer or within the emitting medium, causing a spread in the observed wavelengths via the . When atoms or molecules in a gas have random thermal motions following a Maxwellian , the resulting line profile is Gaussian, reflecting the statistical of line-of-sight velocities. This thermal component dominates in low-density environments like stellar atmospheres or laboratory vapors, where it provides a measure of the kinetic . Bulk motions, such as or , can also contribute to kinematic broadening, but systematic shifts from uniform velocities are distinguished from dispersive broadening. The thermal Doppler effect arises from the projection of the three-dimensional Maxwell-Boltzmann velocity distribution onto the line of sight, yielding a Gaussian intensity profile I(\lambda) \propto \exp\left[ - \left( \frac{c \Delta \lambda}{\lambda_0 \sqrt{2 k T / m}} \right)^2 \ln 2 \right], where \lambda_0 is the central wavelength. The full width at half maximum (FWHM) is \Delta \lambda / \lambda = (2 \sqrt{\ln 2} / c) \sqrt{2 k T / m}, with c the speed of light, k Boltzmann's constant, T the temperature, and m the mass of the atom or molecule; this expression scales with \sqrt{T/m} and is independent of density. In practice, only the longitudinal (radial) velocity component contributes to the shift for each emitter, but for unresolved sources like distant stars or gas clouds, the isotropic thermal motions produce the full Gaussian width. For spatially resolved sources, transverse motions (perpendicular to the line of sight) may partially contribute if projected, as in stellar rotation where differential rotation across the disk creates an additional broadening kernel limited by the rotational velocity v \sin i, typically convolved with the thermal profile. Representative examples illustrate the scale of this broadening. In the solar photosphere, thermal and turbulent motions broaden absorption lines by approximately 0.1 at 5000 , allowing inference of atmospheric dynamics from high-resolution spectra. The cosmic microwave background dipole, arising from the Solar System's bulk velocity of about 370 km s^{-1} relative to the rest frame, appears as a hemispheric temperature shift \Delta T / T \approx v / c \approx 1.23 \times 10^{-3}, equivalent to a uniform Doppler shift rather than dispersive broadening. Turbulent broadening, from coherent velocity fields like or magnetohydrodynamic , adds a further Gaussian component with dispersion \sigma_v = \xi, where \xi is the root-mean-square turbulent speed (often 1–5 km s^{-1} in stellar atmospheres). This is equivalent to an increased T_\mathrm{eff} = T + m \xi^2 / k, enhancing the total Doppler width without altering the . Cosmological redshift, parameterized by z = \Delta \lambda / \lambda, stretches all lines uniformly due to the expansion of , acting as a global shift rather than broadening. Similarly, gravitational redshift from causes a wavelength increase \Delta \lambda / \lambda = GM / (c^2 r) for emitters in a , again a systematic shift without . The overall profile in many cases is the convolution of this Gaussian with the narrower natural , though Doppler effects typically dominate the observed width.

Pressure and Collision Broadening

Pressure and collision broadening, also known as collisional broadening, occurs in dense gases where interatomic collisions interrupt the coherent of the electromagnetic wave emitted or absorbed by the radiator. These interruptions arise from the temporary perturbation of the radiator's energy levels by the electric field of the approaching perturber, leading to a phase shift in the oscillating . For atoms lacking permanent electric , such as metals, the dominant mechanism is the quadratic Stark effect, though linear Stark effects can contribute in cases involving ions or specific symmetries. In the impact approximation, valid when collision durations are much shorter than intervals between collisions (typically at pressures below ~10 atm), the resulting line profile exhibits Lorentzian wings with a full width at half maximum (FWHM) given by \Gamma = N \sigma \bar{v}, where N is the perturber , \sigma is the effective collisional cross-section for phase-changing interactions, and \bar{v} is the mean relative speed of the colliding particles. This linear dependence on density distinguishes pressure broadening from other mechanisms and results in symmetric Lorentzian profiles centered near the unperturbed . Collisions also induce a frequency shift linear in density, expressed as \Delta \omega = -k N, where k is a positive constant reflecting the average second-order phase shift from the quadratic Stark interaction during close encounters. This shift moves the line center toward lower frequencies for most neutral systems and is generally smaller in magnitude than \Gamma. Representative examples include the air broadening of sodium D lines, measured at approximately 0.01 cm^{-1} per Torr due to collisions with N_2 and O_2 molecules. In astrophysical contexts, pressure broadening manifests in the quasar Lyman-alpha forest, where dense intergalactic gas clouds produce Lorentzian wings on absorption lines, aiding density diagnostics in the intergalactic medium. The cross-section \sigma often scales as T^{-1/2} for neutral-neutral collisions under hard-sphere-like assumptions, while van der Waals interactions between neutrals introduce a weaker temperature dependence, typically \sigma \propto T^{-0.2} to T^{-0.4}. Combined with \bar{v} \propto T^{1/2}, the FWHM \Gamma at constant pressure generally decreases with increasing temperature, often as T^{-0.5} or milder.

Other Broadening Mechanisms

Instrumental broadening in spectral lines arises from the limitations of the observing equipment, primarily the finite of the spectrograph. This effect convolves the intrinsic line profile with the instrument's line spread function, typically resulting in a Gaussian broadening characterized by the (FWHM) Δλ ≈ λ / R, where λ is the central and R is the of the instrument. For diffraction gratings, the R is given by R = N m, with N the total number of grooves illuminated and m the diffraction order, while the dλ/dx determines the separation per unit length on the detector. In practice, slit width or sampling can further limit , especially in high-precision where R > 10^5 is required to resolve fine details. Opacity broadening occurs in optically thick media where self-absorption redistributes photons within the line, leading to extended wings that dominate the far profiles beyond the core. This mechanism is prominent in dense plasmas or stellar atmospheres, where the line opacity τ(ν) follows a form, causing the wings to decay as 1/Δν², slower than the exponential tails of Doppler profiles. In spectra during cosmic , for instance, neutral damping wings absorb flux redward of Lyα, suppressing emission and broadening the apparent line width. Macroscopic Doppler broadening stems from large-scale bulk motions in the source, such as radial , , or , which superimpose additional velocity shifts on the motions of individual atoms. In galaxies, rotational velocities up to several hundred km/s can broaden lines like Hα across tens of angstroms, while in active galactic nuclei outflows adds asymmetric redshifts. This effect is distinct from microscopic , as it reflects coherent kinematics of the emitting region rather than random particle motions. Inhomogeneous broadening arises from variations in physical conditions along the , integrating multiple shifted components into a composite profile. In stellar winds, velocity gradients—accelerating from to supersonic speeds—produce P Cygni profiles with broadened absorption troughs spanning the wind's velocity range, often 1000–3000 km/s for hot stars. Such gradients, driven by , result in non-uniform Doppler shifts that mimic turbulent broadening without local collisions. Magnetic fields induce broadening through the , splitting and producing multiple closely spaced components that appear as a net widening when unresolved. The longitudinal splitting is given by \Delta \lambda_B = \frac{e B \lambda^2}{4\pi m_e c^2}, where e is the electron charge, B the strength, m_e the , and c the ; for typical stellar fields of 1–100 G and optical λ ≈ 5000 Å, Δλ_B ranges from ~0.01 to 1 milliangstrom. Primarily a shift mechanism, unresolved Zeeman triplets contribute to effective broadening in magnetized plasmas like sunspots.

Applications and Examples

Spectral Lines of Elements

The of consists of emission lines resulting from transitions to the principal n=2, with prominent wavelengths including H-α at 6562.8 , H-β at 4861.3 , H-γ at 4340.5 , and H-δ at 4101.7 . The H-α line, in particular, is a key diagnostic for , as its emission arises from recombination in ionized regions surrounding young, massive stars. Helium's spectral lines were first identified in 1868 during observations of the solar spectrum by , who noted a novel yellow line at 587.6 nm (5876 ) not matching any terrestrial elements. In the solar corona, metastable lines of neutral , such as those originating from the 2³S (e.g., at 1083.0 nm or 10830 ), are significant due to their long lifetimes and role in collisional excitation processes. Among metals, the sodium D —comprising lines at 5890.0 Å (D₂) and 5895.9 Å (D₁)—arises from transitions in neutral sodium atoms and is exploited in low-pressure sodium-vapor lamps for high-efficiency monochromatic yellow illumination near the peak of human visual sensitivity. Similarly, the calcium H and K lines, at 3968.5 Å (H) and 3933.7 Å (K) in singly ionized calcium, serve as proxies for chromospheric activity in stars, with enhanced core emission indicating magnetic heating and plages. Spectral lines from different ionization stages of elements are denoted using Roman numerals, where the neutral atom is I, singly ionized is II, doubly ionized is III, and so on; for example, O III designates lines from doubly ionized oxygen (O²⁺). Comprehensive catalogs of elemental spectral lines, including wavelengths, intensities, and transition probabilities, are available through resources like the NIST Atomic Spectra Database, which provides critically evaluated data for over 100 elements and their ions. In astrophysics, the atomic linelists compiled by R. L. Kurucz offer extensive transition data tailored for modeling stellar and interstellar spectra. These databases account for effects like line broadening in observed elemental spectra, as detailed in studies of pressure and Doppler mechanisms.

Role in Astrophysics and Chemistry

In astrophysics, spectral lines play a crucial role in measuring the redshift of distant galaxies, which provides evidence for the expansion of the universe as described by Hubble's law, where the redshift z approximates H_0 d / c for nearby objects, with H_0 as the Hubble constant, d the distance, and c the speed of light. This shift in spectral line wavelengths allows astronomers to determine recession velocities and map cosmic structure on large scales. Additionally, spectral lines enable the determination of elemental abundances in stellar atmospheres through the curve of growth technique, which relates the equivalent width W of absorption lines to the column density N of the absorbing species, accounting for saturation effects at higher densities. For the Sun, this method, combined with three-dimensional atmospheric modeling, has refined metallicity estimates, maintaining the solar iron abundance reference [Fe/H] = 0 by definition, with absolute log ε_Fe ≈ 7.50 ± 0.04 as of analyses up to 2009. The logarithmic abundance is derived as [X/H] = log(N_X / N_H) - log(N_X⊙ / N_H⊙), incorporating corrections for damping wings, oscillator strengths, and micro-turbulence via the curve of growth, with the Saha-Boltzmann equation providing the ionization balance: \frac{N_{i+1} n_e}{N_i} = \frac{2}{\Lambda^3} \left( \frac{2 \pi m_e k T}{h^2} \right)^{3/2} e^{-\chi_i / k T}, where i denotes the ionization stage, n_e the electron density, \chi_i the ionization potential, and \Lambda the thermal de Broglie wavelength. In , spectral lines facilitate qualitative analysis through tests, where metal ions emit characteristic colors upon excitation in a , such as the yellow of sodium or of , allowing rapid identification of elements in compounds without advanced instrumentation. For quantitative measurements, (LIF) exploits spectral lines to detect species concentrations; a excites atoms or molecules to higher states, and the emitted intensity is proportional to the ground-state population, enabling sensitive probing of trace gases in flames or solutions, often down to parts-per-billion levels. Recent advancements highlight spectral lines' role in exoplanet studies, where transmission during transits reveals atmospheric compositions; for instance, sodium D-line in hot Jupiter HD 209458b indicated a sodium abundance consistent with values but with hazy upper layers reducing the signal. In quantum computing, Rydberg atoms—excited to high principal quantum numbers—leverage their giant dipole moments and transitions for strong, tunable interactions, enabling the implementation of entangling gates in neutral-atom arrays for scalable operations. James Webb Space Telescope (JWST) observations since 2022 have resolved fine structures in emission lines like [O III] and C IV from early-universe galaxies at z > 10, revealing outflows and rates that challenge models of rapid cosmic evolution. As of 2025, JWST data suggest the presence of the universe's first-generation stars in high-redshift galaxies, identified through distinctive features in "little red dots."

References

  1. [1]
    7 Spectral Lines‣ Essential Radio Astronomy
    Spectral lines are narrow emission or absorption features in spectra of gaseous and ionized sources, used to identify atoms/molecules and measure velocities.
  2. [2]
    5.5: Formation of Spectral Lines - Physics LibreTexts
    Apr 9, 2022 · When electrons move from a higher energy level to a lower one, photons are emitted, and an emission line can be seen in the spectrum.Learning Objectives · The Hydrogen Spectrum · Energy Levels and Excitation
  3. [3]
    Spectral Lines
    A spectral line is a dark or bright line in a spectrum, resulting from photon interaction with a quantum system, causing absorption and re-emission.
  4. [4]
    Atomic Spectroscopy - Spectral Line Shapes, etc. | NIST
    Oct 3, 2016 · Observed spectral lines are always broadened, partly due to the finite resolution of the spectrometer and partly due to intrinsic physical causes.
  5. [5]
    What Do Spectra Tell Us? - Imagine the Universe!
    Sep 24, 2020 · Two very important things we can learn from spectral lines is the chemical composition of objects in space and their motions.
  6. [6]
    Spectroscopy 101 – Beyond Temperature and Composition
    The spectral lines are important because they make it possible to measure redshift precisely (or even notice that it happened at all).
  7. [7]
    Spectroscopy - Center for Astrophysics | Harvard & Smithsonian
    Astronomers also use known spectra to measure the distance to galaxies. The universe is expanding, carrying galaxies along with it, so distant galaxies appear ...
  8. [8]
  9. [9]
    Basic Atomic Spectroscopic Data Handbook | NIST
    Sep 3, 2009 · The relative intensities of the spectral lines observed for any element depend upon the light source and excitation conditions. Thus, even ...Missing: definition | Show results with:definition
  10. [10]
    [PDF] Introduction to Spectral Line Shape Theory - OSTI.GOV
    The shapes of spectral lines are sensitive to the density and temperature of the plasma. A dense plasma will cause spectral lines to shift and broaden. To ...Missing: sharpness | Show results with:sharpness
  11. [11]
    The Era of Classical Spectroscopy - MIT
    These were the first spectral lines ever observed, and Fraunhofer employed the most prominent of them as the first standards for comparing spectral lines ...
  12. [12]
    [PDF] 5. Atomic radiation processes
    Atomic radiation processes involve Einstein coefficients for absorption and emission, line transitions, and line broadening due to damping, collisional, and ...
  13. [13]
    8.4: Boltzmann's Equation - Physics LibreTexts
    Sep 14, 2022 · Boltzmann's Equation shows just what the distribution of the atoms will be among the various energy levels as a function of energy and temperature.
  14. [14]
    [PDF] Chemische Analyse durch Spectralbeobachtungen
    Chemische Analyse durch. Spectralbeobachtungen von. G. Kirchhoff und R. Bunsen. Annalen der Physik und Chemie, Bd. 110 No. 6, 1860, S. 161–189. Neu ...
  15. [15]
    Types of Spectra and Spectroscopy - NASA Science
    Jul 11, 2018 · An emission spectrum is mostly dark with bright lines of color known as emission lines. Emission lines also correspond to specific atoms. Each ...
  16. [16]
    Emission and Absorption Lines
    Emission-line spectra. Low-density clouds of gas floating in space will emit emission lines if they are excited by energy from nearby stars. Planetary ...Emission-Line Spectra · Atomic Energy Levels And... · Example: The Spectrum Of...
  17. [17]
    [PDF] Emission mechanisms. II
    Emission mechanisms include spontaneous emission, where an excited atom emits a photon, and stimulated emission, where a radiation field triggers the emission.
  18. [18]
    Bohr's Model of the Hydrogen Atom – University Physics Volume 3
    ... lines are: red (656 nm), called the H- \text{α} line ... The series of emission lines given by the Balmer formula is called the Balmer series for hydrogen.
  19. [19]
    Astronomy 122 - The Interstellar Medium
    some very rare lines are seen (referred to as "forbidden lines"). example of "forbidden lines" is green light from the Orion Nebula (M42). doubly ionized ...<|control11|><|separator|>
  20. [20]
    [PDF] Rate Equations and Detailed Balance Initial question: Last time we ...
    Let A21, the “Einstein A coefficient”, be the rate of spontaneous emission. This is the transition probability per unit time that an atom currently in level 2 ...
  21. [21]
    Optical Emission Spectroscopy
    Optical emission spectroscopy allows non-perturbing characterization of plasma parameters survey of atomic and molecular species via the emission from excited ...
  22. [22]
    [PDF] Optical Emission Diagnostics Of Laser Produced Plasma From ...
    Optical emission spectroscopy is a technique which analyzes the light emitted from the plasma. The optical spectroscopic studies of laser produced plasma are ...
  23. [23]
    Absorption and Emission Lines
    An absorption line occurs when an electron jumps from a lower energy state to a higher energy state, extracting the required photon from an outside source of ...
  24. [24]
    Spectra and What They Can Tell Us
    A spectrum is simply a chart or a graph that shows the intensity of light being emitted over a range of energies. Have you ever seen a spectrum before?Missing: sharpness position
  25. [25]
    [PDF] Chapter 2
    Among the deep and conspicuous absorption lines, which were labeled by Fraunhofer, are the calcium H and K lines at 3968.5 angstrom and 3933.7 angstrom, ...Missing: definition | Show results with:definition
  26. [26]
    [PDF] Development of a New Ca II H and K Spectrophotometric ...
    The Ca II H and K lines were observed as absorption lines in the spectrum of the Sun in 1814 by Joseph von Fraunhofer. Fraunhofer catalogued and measured the ...<|control11|><|separator|>
  27. [27]
    A survey of Ultraviolet interstellar absorption lines. - ADS
    The present survey provides data on ultraviolet absorption lines in the spectra of 88 early-type stars.
  28. [28]
    [PDF] STELLAR SPECTRA A. Basic Line Formation
    Feb 6, 2007 · It illustrates that detailed physics, in this case atomic physics, is usually needed to explain cosmic ... spectral line. When atoms absorb or ...<|control11|><|separator|>
  29. [29]
    The curve of growth
    The oscillator strength for this transition is f = 0.645. Exercise: Measure the equivalent width of this sodium absorption line; Find its location on the ...
  30. [30]
    [PDF] Absorption Features - NMSU Astronomy
    That is, the equivalent width is degenerate for various combinations of column density, N, and absorption coefficient, α(λ). For a given atomic transition ...
  31. [31]
    [PDF] II Neutral Atomic Hydrogen (HI) Regions
    . Line Emission Coefficient (jul). The line emission coefficient describes radiative transitions from an upper excited level into lower levels: ul line j. j d ...
  32. [32]
    [PDF] 12. Megamasers
    → line absorption coefficient negative. < 0 → line emission exponentially amplified by stimulated emission κ⌫ = hν. 4π φ(ν)Blu (nl - gl gu nu). I. 0. ⌫. I ...
  33. [33]
    Vibrational Emission Study of the CN and C2 in Nylon and ZnO ...
    Sep 5, 2022 · In the present work, molecular emissions, CN (B2Σ+ to X2Σ+) violet and the C2 (d3 Пg–a3 Пu) swan band system of nylon polymer plasma, at various ...
  34. [34]
    [PDF] 1913 On the Constitution of Atoms and Molecules
    On the theory of this paper the only neutral atom which contains a single electron is the hydrogen atom. The permanent state of this atom should correspond ...
  35. [35]
    [PDF] An undulatory theory of the mechanics of atoms and molecules - ISY
    At first sight this equation seems to offer ill means of solving atomic problems, e.g. of defining discrete energy-levels in the hydrogen atom. Being a ...
  36. [36]
    [PDF] On the Quantum Theory of Molecules - UFPR
    In order to determine the eigenfunctions and thereby the transition probabilities only to the zeroth approximation, the energy calculation must be carried out ...
  37. [37]
    [PDF] The Quantum Theory of the Emission and Absorption of Radiation.
    The Quantum Theory of the Emission and. Absorption of Radiation. P. A. M. Dirac, St. John's College,. Cambridge, and Institute for Theoretical Physics, ...
  38. [38]
    Atomic Spectros. - Spectral Lines - NIST
    Selection rules for discrete transitions. Electric dipole (E1) ("allowed"), Magnetic dipole (M1) ("forbidden"), Electric quadrupole (E2) ("forbidden") ...
  39. [39]
    [PDF] Notes on Atomic Structure 1. Introduction 2. Hydrogen Atoms and ...
    Jan 7, 2011 · The selection rules for our most common types of transitions are: Electric dipole: Exact: ΔJ=−1, 0, or +1 (no 0→0); opposite parity.
  40. [40]
    [PDF] Molecular Spectroscopy. Born-Oppenheimer Approximation
    Selection rules for electronic transitions in diatomic molecules. Allowed transitions. Examples. ΔΛ = 0, ±1. Σ ↔ Σ, Π ↔ Π,. Σ ↔ Π, Δ ↔ Π. ΔS = 0. 1Σ ↔ 1Σ, 3Π ...
  41. [41]
    [PDF] Molecular Structure & Spectroscopy 1. Introduction - Caltech (Tapir)
    Feb 4, 2010 · The electronic states do however obey the usual selection rules from atomic physics: |ΔΛ|≤1; no Σ+ Σ−; and for homonuclear molecules, the ...
  42. [42]
    Atomic Spectroscopy - Spectral Lines | NIST
    Oct 3, 2016 · 17. Spectral Lines: Selection Rules, Intensities, Transition Probabilities, Values, and Line Strengths. Selection rules for discrete transitions ...Missing: definition | Show results with:definition
  43. [43]
    Interference between Franck–Condon and Herzberg–Teller Terms ...
    Aug 5, 2022 · (33,34) suggested that non-Condon vibronic coupling plays a central role in facilitating electronic–vibrational energy transfer and enhancing ...Supporting Information · Author Information · References
  44. [44]
    Search for trans-iron elements in hot, helium-rich white dwarfs with ...
    ... multiplets of C IV at 1108, 1169, 1198,. 1230Ε (Figs. 2, A.3, and A.4). For the determination of g we used lines of the He II Pickering series (Fig. 3) ...
  45. [45]
    NIST: Atomic Spectra Database Lines Form
    Selecting this option would generate spectral lines distribution for one or several ions. All lines within a specific ion stage are given the same intensity in ...
  46. [46]
    [PDF] Conversion from vacuum to standard air wavelengths
    Apr 8, 2011 · We explore possible expressions for the conversion between vacuum wave- lengths and wavelengths in standard air. We find that the equation ...
  47. [47]
    [PDF] The classification of stellar spectra
    Notation: - neutral element: H I, He I, Fe I, etc. - single-ionized element: H II, O II, etc. - double-ionized element: O III, etc.
  48. [48]
    Wavenumber - RP Photonics
    Spectroscopic wavenumbers are usually used with units of cm−1. If these units are found, it is usually clear that a spectroscopic wavenumber is meant. Note ...
  49. [49]
    The shape of spectral lines
    Equivalent width · stretches vertically from zero intensity to the continuum intensity · has the same area as the line.Equivalent width · Weak lines and strong lines · Natural broadening
  50. [50]
    Line Profile Functions - AstroBaki - CASPER
    Aug 18, 2021 · The Voigt Profile is a convolution of the Lorentzian Profile from Natural Broadening and the Gaussian Profile from Doppler Broadening. It is ...Line Profile Function · 2 Doppler (Gaussian... · 3 The Voigt Profile
  51. [51]
    NIST: Atomic Spectra Database - Spectral Lines Help File
    The Line Identification Plot is a stick plot showing the line positions for all chosen ions in the given wavelength range. It can be used to diagnose which ions ...
  52. [52]
    [PDF] 1 CHAPTER 11 CURVE OF GROWTH 11.1 Introduction The curve of ...
    The curve of growth (die Wachstumskurve) is a graph showing how the equivalent width of an absorption line, or the radiance of an emission line, increases with ...Missing: explanation | Show results with:explanation
  53. [53]
    Particle lifetimes from the uncertainty principle - HyperPhysics
    Γ is often referred to as the "natural line width". It is of great importance in high energy accelerator physics where it provides the means for determining the ...
  54. [54]
    [PDF] Chapter 7. Spectral Linewidth and Line shape
    This is called the natural linewidth of the transition. Ik → li>. The frequency distribution was first obtained quantum mechanically by Weisskopf and Wigner in ...
  55. [55]
    Spontaneous Emission – quantum effect, vacuum fluctuations, laser ...
    Spontaneous emission is a quantum effect, which in a semiclassical picture can be described as an emission which is stimulated by vacuum noise.
  56. [56]
    Observation of the 1S–2P Lyman-α transition in antihydrogen | Nature
    Aug 22, 2018 · With its natural line width of about 100 MHz (comparable to our laser width), the Lyman-α transition can in principle be employed to cool ...
  57. [57]
    The Doppler Widths of Solar Absorption Lines - NASA ADS
    In the solar atmosphere, Doppler broadening and collision broadening are far more important than natural width. This paper deals chiefly with Doppler broadening ...Missing: Å | Show results with:Å
  58. [58]
    Spectral Lines
    These motions are often accounted for by an effective Doppler width, which includes both thermal and turbulent broadening, assuming a gaussian distribution for ...
  59. [59]
    Ab initio computation of the broadening of water rotational lines by ...
    Oct 27, 2010 · The very general impact approximation states that collision times are much shorter than the time between collisions. Within that approximation, ...
  60. [60]
    The phase shifts leading to the broadening and shift of spectral lines
    The calculated values are based on the use of the Lindholm–Foley theory [4] for the broadening and shift of spectral lines. It was shown in [10] that the values ...
  61. [61]
    A theoretical and experimental study of pressure broadening of the ...
    May 28, 2014 · Spectrum in the impact approximation. The rotationally resolved A-band magnetic dipole absorption spectrum concerns transitions of the type ...
  62. [62]
    Power-law temperature dependence of collision broadening and ...
    The usual approach is to assume the temperature dependence of the collision broadening coefficient in the form of a power-law: γ ( T ) = γ ( T 0 ) T T 0 n , ...<|control11|><|separator|>
  63. [63]
    [PDF] Quantifying resolving power in astronomical spectra - arXiv
    Aug 5, 2013 · The spectral resolving power R = λ/δλ is perhaps the most important single property of a spectrograph. The wavelength increment δλ is the ...
  64. [64]
    [PDF] 1 Spectroscopy 2 Physical Origins of Lines - Loren Anderson
    Spectral lines arise from transitions between energy states. ... 4.2 Line broadening. A number of ... Last thing on diffraction gratings, resolving power.
  65. [65]
    Spectral Resolution and Line Spread Functions - STScI
    May 21, 2025 · The dominant effect in the observed spectrum is a broadening of the spectral features due to the finite resolution of the instrument.Missing: instrumental | Show results with:instrumental
  66. [66]
    [PDF] 14 Shape of Spectral Lines - Astrophysics Data System
    Imagine a line with an arbitrarily sharp atomic absorption coefficient [that is, Sν = δ(ν-ν0)]. If there were no motion in the atmosphere, the lowest-lying ...Missing: sharpness | Show results with:sharpness
  67. [67]
    [PDF] JWST observations of galaxy damping wings during reionization ...
    Aug 10, 2023 · Many of these observed galaxies exhibit strong damping wing absorption redward of their Lyman-𝛼 emission. These observations have been used to ...
  68. [68]
    [PDF] Science with the X-ray Imaging and Spectroscopy Mission - HEASARC
    With XRISM, we will be able to resolve individual spectral lines and trace gas motions through the Doppler broadening and line energy shifts in many galaxies.
  69. [69]
    [PDF] Spectroscopic Study of Stellar Winds of Massive Stars
    Dec 6, 2022 · For the optically thick case, the line acceleration is independent of the opacity, and instead varies in proportion to the velocity gradient.
  70. [70]
    [PDF] ALMA spectral line and imaging survey of a low and a high mass ...
    Adding an additional force in the subsonic part of the wind (in this case due to radiation pressure on the grains) results in a smaller wind velocity gradient ...
  71. [71]
    Atomic Spectroscopy - Zeeman Effect | NIST
    Oct 3, 2016 · The magnetic flux density is B, and µB is the Bohr magneton (µB = eℏ/2me). The wavenumber shift Δσ corresponding to this energy shift is. Δσ = ...Missing: broadening | Show results with:broadening
  72. [72]
    Strong Lines of Hydrogen ( H ) - Physical Measurement Laboratory
    Intensity, Air Wavelength (Å), Spectrum, Reference. 5, 3835.384, H I, RCWM80. 6, 3889.049, H I, RCWM80. 8, 3970.072, H I, RCWM80. 15, 4101.74, H I, RCWM80.
  73. [73]
    H α emission in local galaxies: star formation, time variability, and ...
    The nebular recombination line H α is widely used as a star formation rate (SFR) indicator in the local and high-redshift Universe. We present a detailed H α ...
  74. [74]
    This Month in Physics History | American Physical Society
    Some 5,000 miles away, on October 20, 1868, the English astronomer Joseph Norman Lockyer also succeeded in observing the solar prominences in broad daylight.
  75. [75]
    Helium Line Emissivities in the Solar Corona - IOPscience
    Jul 24, 2020 · We present new collisional-radiative models (CRMs) for helium in the quiescent solar corona and predict the emissivities of the He and He + lines.
  76. [76]
    Hydrogen-Like Atoms:Sodium - HyperPhysics
    The sodium spectrum is dominated by the bright doublet known as the Sodium D-lines at 588.9950 and 589.5924 nanometers.
  77. [77]
    Hα and the Ca II H and K lines as activity proxies for late-type stars
    The main chromospheric activity indicator is the S index, which is the ratio of the flux in the core of the Ca II H and K lines to the continuum nearby, and is ...
  78. [78]
    Spectroscopic Notation
    Spectroscopists often refer to ionization stages of elements by what is called the "spectroscopic notation": The neutral atom (of, e.g., Neon) is called "Ne I", ...
  79. [79]
    Atomic Spectra Database | NIST
    Spectral lines and associated energy levels displayed in wavelength order with all selected spectra intermixed or in multiplet order.
  80. [80]
    Kurucz/Linelists
    New atomic results combined with old gfall remnants are given in /GFNEW GF10 wavelength-sorted atomic lines in 10 nm intervals GF100 wavelength-sorted atomic ...Missing: astrophysics | Show results with:astrophysics
  81. [81]
    Hubble Cosmological Redshift - NASA Science
    Sep 17, 2024 · The dark lines of an absorption spectrum are shifted toward the right or red end of the spectrum. This shift is the result of the expanding ...
  82. [82]
    [PDF] The chemical composition of the Sun - arXiv
    Sep 4, 2009 · The reader is referred to Nordlund, Stein & Asplund (2009) for a recent review of the physics of solar surface convection and the numerical ...Missing: growth | Show results with:growth
  83. [83]
    [PDF] excitation and ionization, Saha's equation stellar spectral ...
    Saha-Boltzmann predictions of the fractional concentration. Nr,s /N of the lower level of the lines indicated in the upper panel against temperature T (given ...
  84. [84]
    Misconceptions and Insights about Flame Tests - ACS Publications
    Sep 9, 2025 · Flame tests have been used for elemental analysis since the earliest days of chemistry, when the elements were first isolated and identified.Missing: authoritative | Show results with:authoritative
  85. [85]
    Laser induced fluorescence measurements and modeling of nitric ...
    Laser-induced fluorescence (LIF) has been applied to the quantitative measurement of nitric oxide (NO) in premixed, laminar, high-pressure flames.
  86. [86]
    DETECTION OF AN EXTRASOLAR PLANET ATMOSPHERE1
    ABSTRACT. We report high-precision spectrophotometric observations of four planetary transits of HD 209458, in the region of the sodium resonance doublet at ...Missing: hot | Show results with:hot
  87. [87]
    Quantum technologies with Rydberg atoms - Frontiers
    Jul 11, 2024 · Rydberg atoms have highly controllable exotic properties such as strong inter-atomic interaction, high polarizability, and long lifetimes ...Abstract · Quantum information... · Harnessing Rydberg atoms for... · Discussion
  88. [88]
    Super-early JWST galaxies, outflows, and Lyα visibility in the Epoch ...
    The overabundance of super-early (redshift z > 10) luminous (MUV < −20) blue galaxies detected by JWST has been explained as being due to negligible dust ...