Fact-checked by Grok 2 weeks ago

Electron paramagnetic resonance

Electron paramagnetic resonance (EPR), also known as electron spin resonance (ESR), is a spectroscopic technique that detects and characterizes chemical species with one or more unpaired electrons by measuring their absorption of microwave radiation in the presence of an external magnetic field. The method relies on the Zeeman effect, where the magnetic field splits the energy levels of the electron spins, and microwave photons at frequencies typically in the X-band range (around 9-10 GHz) induce transitions between the m_s = +1/2 and m_s = -1/2 states when the energy difference matches h\nu = g \mu_B B, with g as the electron g-factor, \mu_B the Bohr magneton, and B the magnetic field strength. This absorption is detected as a change in microwave power, producing a spectrum that reveals information about the electronic environment, spin states, and interactions of the paramagnetic centers. The discovery of EPR is credited to Soviet physicist Evgenii Konstantinovich Zavoisky in 1944, who observed the resonance in a sample of copper chloride dihydrate using a low-frequency setup at 133 MHz and a field of about 4.76 mT. Rapid advancements followed , as researchers adapted surplus radar technology—particularly klystrons operating at 9-10 GHz—to build practical spectrometers, establishing the X-band as the standard for high sensitivity. Early applications focused on fundamental physics, but by the , EPR was applied to biological materials, marking its expansion into interdisciplinary fields. In EPR spectra, the position of the resonance line is characterized by the g-factor, which is approximately 2.0023 for a but shifts due to spin-orbit coupling and the local chemical environment of the . Additional structure arises from hyperfine interactions between the and nearby nuclear spins, producing splitting patterns (e.g., N+1 lines for N equivalent nuclei with ), which provide details on the number and type of surrounding atoms. The technique's allows detection of paramagnetic at concentrations as low as parts per million, making it ideal for studying transient radicals, ions in proteins, and defects in solid-state materials. EPR finds broad applications across , , and , including the characterization of free radicals in , the oxidation states of metal centers in enzymes like and , and surface sites in catalysts for energy conversion processes such as electrocatalysis. variants, such as pulsed EPR and high-field/high-frequency methods (up to W-band at 95 GHz), enhance resolution for complex systems, while techniques enable spatial mapping of distributions in biological tissues or materials. Despite challenges like the need for paramagnetic samples and potential broadening from relaxation effects, EPR remains a powerful, non-destructive for probing electronic structure at the molecular level.

Overview

Definition and principles

Electron paramagnetic resonance (EPR), also known as electron spin resonance (ESR), is a spectroscopic technique that detects and characterizes materials containing unpaired electrons by measuring the resonant absorption of microwave radiation in the presence of a static magnetic field. This method probes paramagnetic species, where paramagnetism arises from the intrinsic spin angular momentum of unpaired electrons, typically with spin quantum number S = 1/2, which generates a magnetic dipole moment that interacts with external fields. In the presence of a magnetic field B, the Zeeman effect causes the splitting of the electron's spin energy levels, with the energy difference given by \Delta E = g \mu_B B, where g is the electron g-factor (approximately 2.0023 for free electrons), and \mu_B is the Bohr magneton. Resonance occurs when the energy of the applied microwave radiation matches this splitting, satisfying the condition h \nu = g \mu_B B, where h is Planck's constant and \nu is the microwave frequency (typically 9–35 GHz). At resonance, transitions between the split spin states are induced, producing a measurable absorption signal. EPR is analogous to nuclear magnetic resonance (NMR) but targets electron spins rather than nuclear spins. Common samples studied by EPR include organic free radicals, ions, and point defects in solids, which exhibit unpaired electrons and thus paramagnetic behavior. The technique is sensitive to species with lifetimes longer than approximately $10^{-7} s, as shorter-lived transients require specialized time-resolved methods for detection.

Historical development

The theoretical foundations of electron paramagnetic resonance (EPR) trace back to 1925, when and proposed the intrinsic spin of the electron to account for splittings in atomic spectra. This concept provided the quantum mechanical basis for the of unpaired electrons, essential for EPR phenomena. Early experimental efforts toward magnetic resonance, including unsuccessful attempts to detect in solids, were led by Cornelis Jacobus Gorter in the during the 1930s. Gorter's work highlighted the challenges of sensitivity and detection in paramagnetic systems, paving the way for subsequent refinements. Edward M. Purcell's 1946 discovery of in solids further advanced the broader field of magnetic resonance techniques, influencing EPR instrumentation through shared principles of spin transitions and signal detection. The first experimental observation of EPR occurred in 1944–1945, when Yevgeny Zavoisky at Kazan University in the Soviet Union detected resonance absorption in paramagnetic salts using a continuous-wave setup at a radio frequency of about 133 MHz (wavelength ~2.25 m). Zavoisky's breakthrough, published in 1945, marked the birth of EPR as a spectroscopic method, though initial applications were limited by equipment constraints. Post-World War II advancements accelerated in the United States and Europe, fueled by wartime radar technology that supplied reliable microwave sources operating at X-band frequencies (around 9–10 GHz). These developments enabled higher sensitivity and broader adoption; for instance, early U.S. efforts included observations at the University of Pittsburgh in 1946, contributing to the technique's refinement for chemical analysis. In the 1950s, commercialized EPR spectrometers, such as the V-4500 model, making the technique accessible to research laboratories worldwide and spurring applications in and . A pivotal milestone was the discovery of electron-nuclear double resonance (ENDOR) by George Feher, which combined EPR with radiofrequency irradiation to resolve hyperfine interactions in greater detail. The 1960s saw the introduction of pulsed EPR methods, also pioneered by Feher, allowing time-resolved studies of spin dynamics and transient species. EPR played a crucial role in early free radical detection, notably in research; and colleagues in 1954 used it to identify light-induced radicals in biological materials like chloroplasts, revealing paramagnetic signals from derivatives. The 1980s brought high-field EPR, with the first spectrometers operating above 94 GHz installed in 1983, enhancing resolution for g-anisotropy and in complex systems. Post-2000, EPR has found applications in , particularly in characterizing qubits in nitrogen-vacancy centers and semiconductor quantum dots, enabling precise measurements of times and entanglement. Contemporary advancements include compact benchtop EPR systems in the , which democratize access for routine analysis in chemistry and through improved portability and sensitivity at lower fields.

Fundamental Theory

Signal origin and quantum mechanics

Electron paramagnetic resonance (EPR) signals arise from the quantum mechanical behavior of unpaired electrons in a magnetic field, where the electron's intrinsic spin interacts with an applied microwave field to induce transitions between quantized spin states. The fundamental entity responsible is the electron's spin angular momentum, characterized by the quantum number S = \frac{1}{2}, which gives rise to two possible projections along a quantization axis, m_s = \pm \frac{1}{2}. This spin property endows the electron with a magnetic moment \boldsymbol{\mu} = -g \mu_B \mathbf{S}, where g is the electron g-factor (approximately 2.0023 for a free electron) and \mu_B is the Bohr magneton. In the presence of an external static \mathbf{B}, the is described by the Zeeman Hamiltonian for a : H = g \mu_B \mathbf{B} \cdot \mathbf{S}, which quantizes the states into two levels separated by the Zeeman splitting \Delta E = g \mu_B B, with the lower state corresponding to m_s = -\frac{1}{2} (aligned antiparallel to \mathbf{B}) and the higher to m_s = +\frac{1}{2}. When a perpendicular oscillating is applied at \nu, transitions between these states become possible if the matches the splitting, h\nu = g \mu_B B, leading to absorption or of photons. The selection rules governing these transitions are \Delta m_s = \pm 1 for the electron , with no change in any associated quantum numbers. The observable EPR signal originates from the net absorption of energy, which occurs because the two spin states have unequal populations at , following the ; the lower energy state is more populated, resulting in more upward (m_s = -\frac{1}{2} \to +\frac{1}{2}) transitions than downward ones. Quantum mechanically, the transition rates are calculated using , where the field acts as the perturbation H'. According to , the transition probability per unit time is w = \frac{\pi}{\hbar} \left| \langle f | H' | i \rangle \right|^2 \delta(E_f - E_i), with the matrix element \langle f | H' | i \rangle determining the strength of coupling between initial (i) and final (f) states, and the delta function ensuring . This framework quantifies the resonant absorption that produces the detectable EPR signal.

Energy levels and transitions

In electron paramagnetic resonance (EPR), the application of an external B induces Zeeman splitting of the degenerate energy levels for an with S = 1/2. The energy of each level is given by E = [g](/page/G) \mu_B m_s [B](/page/List_of_punk_rap_artists), where m_s = \pm 1/2 is the , [g](/page/G) is the electron g-factor (approximately 2 for a ), and \mu_B is the . This results in two levels separated by \Delta E = [g](/page/G) \mu_B [B](/page/List_of_punk_rap_artists), with the lower energy state at m_s = -1/2 and the upper at m_s = +1/2. For [g](/page/G) \approx 2, the splitting exhibits linear dependence on the strength, enabling precise control over the separation through field variation. Microwave radiation induces transitions between these Zeeman levels when the photon energy matches the splitting, satisfying the resonance condition h \nu = g \mu_B B, or equivalently, the resonance frequency \nu = (g \mu_B B)/h, where h is Planck's constant. The allowed transitions follow the selection rule \Delta m_s = \pm 1, corresponding to absorption from the lower to the upper level. The transition probability is governed by the magnetic dipole moment operator, with intensity proportional to |\langle +1/2 | \hat{\mu} \cdot \mathbf{B_1} | -1/2 \rangle|^2 (N_- - N_+), where \mathbf{B_1} is the oscillating microwave magnetic field and N_- , N_+ are the populations of the lower and upper levels, respectively. In isotropic systems, this yields a sharp resonance line, but intensity factors such as the orientation of \mathbf{B_1} relative to the spin quantization axis modulate the signal strength. In non-spherical environments, the g-factor becomes anisotropic, described by a g-tensor, leading to angular dependence of the effective g-value: g(\theta, \phi) = \sqrt{g_x^2 \sin^2 \theta \cos^2 \phi + g_y^2 \sin^2 \theta \sin^2 \phi + g_z^2 \cos^2 \theta}. This anisotropy causes the resonance field to vary with molecular orientation, producing characteristic powder patterns in polycrystalline or frozen solution samples, where all orientations contribute to a broadened absorption envelope. The of the Zeeman levels follows the Maxwell-Boltzmann , with the ratio N_+ / N_- = \exp(-g \mu_B B / kT), where k is Boltzmann's constant and T is the temperature. At and typical fields (e.g., ~0.35 T for X-band EPR), g \mu_B B / kT \approx 0.001, resulting in a small population excess in the lower level (~0.1%) that drives net , though the signal remains detectable due to sensitive . This underscores the need for low temperatures or high fields to enhance signal intensity in EPR experiments.

Relaxation mechanisms

In electron paramagnetic resonance (EPR), relaxation mechanisms describe the processes by which excited spin states return to , influencing signal intensity, linewidth, and the feasibility of pulsed experiments. Spin-lattice relaxation, characterized by the T_1, involves the transfer of from the spin system to the surrounding vibrations, or phonons, restoring the longitudinal magnetization along the magnetic field direction. The primary mechanisms for T_1 relaxation are the , Raman, and Orbach processes. The process occurs through the absorption or emission of a single whose energy matches the Zeeman splitting, dominating at low temperatures where T_1 \propto T^{-1} and exhibits strong dependence, such as T_1 \propto B^{-4} for systems in non-Kramers ions. The Raman process involves of two phonons, with virtual intermediate states, and becomes prominent at higher temperatures with T_1 \propto T^{-5} to T^{-9} depending on the assumptions. The Orbach process is a resonant two-phonon mechanism via a real excited state at energy \Delta above the , yielding T_1 \propto \exp(\Delta / kT), often observed when low-lying excited levels exist. Overall, T_1 lengthens at low temperatures as phonon populations decrease, enabling longer-lived excitations. Spin-spin relaxation, governed by T_2, arises from of the transverse magnetization due to local fluctuations from dipole-dipole interactions between spins, hyperfine couplings, or molecular motions. This process does not involve energy transfer to the but leads to loss of phase coherence among spins. In the Bloch-Wangsness-Redfield theory for a system, the total transverse relaxation rate is given by \frac{1}{T_2} = \frac{1}{2T_1} + \frac{1}{T_2'}, where $1/(2T_1) accounts for the contribution from longitudinal relaxation, and T_2' represents pure dephasing from time-dependent local fields. Homogeneous broadening occurs when all spins experience the same average field, resulting in a Lorentzian lineshape with linewidth \Delta B \propto 1/T_2, specifically \Delta B = \frac{h}{g \mu_B \pi T_2} for the full width at half maximum in the unsaturated limit. In contrast, inhomogeneous broadening stems from a distribution of static local fields across spins, producing Gaussian-like envelopes without direct dependence on T_2. In solid samples, T_2 is often limited by hyperfine interactions with nearby nuclei, constraining dephasing times. At , typical T_1 values range from microseconds to milliseconds for many paramagnetic centers, such as organic radicals or ions in solution, which supports the application of pulsed techniques by allowing sufficient time for spin manipulation before decay. Both T_1 and T_2 exhibit field dependence through modulation of spin-phonon or spin-spin couplings, with T_1 generally shortening at higher fields due to enhanced direct process efficiency.

Instrumentation

Magnets and field control

In electron paramagnetic resonance (EPR) spectroscopy, the static magnetic field is generated by electromagnets or superconducting magnets, selected based on the desired field strength and experimental requirements. Electromagnets, typically consisting of water-cooled iron-core solenoids, are commonly employed for standard EPR experiments at moderate fields ranging from 0.3 to 2 T, such as the approximately 0.35 T required for X-band operation at 9-10 GHz. These systems allow precise and rapid adjustment of the field, essential for sweeping through resonance conditions dictated by the Zeeman effect. For high-field EPR, superconducting magnets, often using niobium-titanium or high-temperature superconductors, enable fields exceeding 2 T and up to 10 T or more, facilitating enhanced resolution of g-anisotropy and hyperfine interactions in complex systems. Field homogeneity is critical in EPR to minimize broadening of spectral lines, typically achieved to better than 1 μT over the sample volume of a few cubic millimeters. This is accomplished through passive shimming with iron pieces and active shimming using dedicated shim coils that generate compensating gradient fields, such as Z1, , and higher-order terms, to correct inhomogeneities arising from magnet imperfections or environmental influences. Modern EPR systems incorporate programmable shim sets to maintain this homogeneity across a wide field range, ensuring sharp lines for narrow resonances down to 0.01 mT linewidths. To acquire spectra, sweep coils mounted on the magnet poles provide a linear of the main field, typically at rates corresponding to sweeps of several minutes for full spectral coverage, with low-frequency components around 100 Hz in some configurations for precise control. Superimposed on this is a small oscillatory field, often at 100 kHz, to enable phase-sensitive detection via lock-in , enhancing without distorting the lineshape. Field stability is paramount for resolving , with modern systems achieving drifts below 1 nT/s through feedback control, preventing baseline distortions in long acquisitions. of the field is performed using Hall probes positioned near the sample for monitoring or proton NMR probes for absolute accuracy, ensuring precise alignment with for resonance.

Microwave sources and bridges

In electron paramagnetic resonance (EPR) spectroscopy, microwave sources provide the coherent radiation necessary to induce transitions between spin states, typically operating in the X-band frequency range around 9.5 GHz with output powers of approximately 100-200 mW to avoid saturation of the sample. Traditional sources include klystrons, which are vacuum-tube oscillators capable of generating stable, tunable microwave signals but require high-voltage operation and mechanical tuning. These have largely been supplanted in modern continuous-wave (CW) EPR systems by solid-state devices such as Gunn diodes, which offer compact design, lower power requirements, and reliable output without the maintenance issues of klystrons, delivering fixed-frequency radiation that can be adjusted via attenuators. For higher frequencies, such as Q-band (around 35 GHz), extended interaction klystrons or advanced solid-state sources like IMPATT diodes may be employed to maintain sufficient power levels. The microwave bridge serves as the core circuitry for coupling the source to the sample , isolating the weak EPR signal from the incident , and enabling phase-sensitive detection. Common designs include reflection cavity bridges, where the EPR absorption is detected as a change in the reflected from the , often using a to route the incident wave to the and direct the reflected signal to a detector while isolating the source from back-s. Alternatively, magic-T hybrid junctions provide balanced operation by splitting the into sample and reference paths, with the magic-T ensuring effective (typically >20 ) between the transmitter and receiver arms to prevent overload. Superheterodyne bridges, incorporating a for frequency down-conversion, are favored in high-frequency or low-signal applications to improve and reduce noise, though they add complexity with amplification stages. A reference arm in the bridge delivers an attenuated portion of the source signal (often 1-10% of incident ) to the detector, establishing a stable bias current (around 200 µA for Schottky diodes) and allowing phase adjustment via shifters to optimize the for dispersive or absorptive signal components. Field , typically sinusoidal at 100 kHz with amplitudes of 0.1-1 G (less than the linewidth to avoid distortion), is superimposed on the static , converting the DC EPR absorption into an AC signal that is demodulated using lock-in to suppress broadband noise and achieve phase-sensitive output. This configuration enables detection sensitivities down to approximately 10^{11} unpaired for standard samples, with modern bridges introduced after 2010 offering programmable , automated , and software-defined for enhanced flexibility and reproducibility in diverse experimental setups.

Resonators and detection systems

In electron paramagnetic resonance (EPR) spectroscopy, resonators, often referred to as microwave cavities, are essential components that confine the magnetic field to interact efficiently with the sample, thereby enhancing signal . The most commonly employed resonators include rectangular cavities operating in the TE102 mode and cylindrical cavities in the TE011 mode, particularly at X-band frequencies around 9-10 GHz. These designs provide a high concentration of the oscillating (B1) at the sample position while minimizing losses, with typical unloaded quality factors (Q) exceeding 5000, which is crucial for achieving detectable signal intensities in low-concentration samples. The sensitivity of an EPR experiment is strongly influenced by the filling factor η, which quantifies the fraction of the resonator's magnetic field that overlaps with the paramagnetic sample volume. The EPR signal amplitude is proportional to η √(P ), where P is the incident microwave power; this relationship underscores the need to optimize sample size and to maximize η without excessively degrading due to losses. For instance, larger samples can increase η but may introduce losses that reduce , so empirical —often by adjusting sample dimensions to the cavity's sensitive —is standard practice to balance these parameters. Detection systems in EPR convert the weak microwave power changes induced by paramagnetic absorption into measurable electrical signals. Traditional setups employ or crystal detectors, which rectify the reflected microwave power from the , providing a direct measure of the . To improve , phase-sensitive detection is commonly used, which mixes the detector output with a reference signal phase-locked to the field , yielding the first of the rather than the direct ; this enhances for narrow lines but requires careful phase adjustment to avoid distortions. Temperature control is integral to resonators, enabling studies of thermally activated dynamics and relaxation. Variable temperature inserts, typically helium-flow cryostats, allow operation from 4 to 300 , with precise regulation (±0.1 ) to investigate cryogenic effects like reduced linewidths or enhanced polarization in biomaterials. For aqueous samples, which suffer from high dielectric losses that degrade Q, specialized flat cells—thin quartz capillaries oriented perpendicular to the electric field—are used to minimize perturbation of the resonator while accommodating larger sample volumes. Modern dielectric resonators, often incorporating low-loss materials like rutile (TiO2), further improve sensitivity for biomaterials by supporting multi-channel sample holders that reduce losses and enable high-throughput analysis of biological fluids or tissues.

Spectral Analysis

g-factor determination

The g-factor serves as a key spectroscopic parameter in electron paramagnetic resonance (EPR) spectroscopy, quantifying the deviation of an unpaired electron's magnetic moment from that of a free electron due to its local electronic environment. It is defined by the resonance condition g = \frac{h \nu}{\mu_B B_{\text{res}}}, where h is Planck's constant, \nu is the applied microwave frequency, \mu_B is the Bohr magneton, and B_{\text{res}} is the magnetic field strength at which resonance occurs. For a free electron, the g-factor is precisely 2.0023, but interactions such as spin-orbit coupling cause shifts in this value for bound electrons in molecules or solids. In isotropic systems, such as solution-phase samples with tumbling radicals, the g-factor is a scalar quantity representing an over orientations. However, in anisotropic environments like single crystals or oriented samples, the g-factor becomes a tensor with principal components g_x, g_y, and g_z, reflecting the directional dependence of the electron's magnetic . For polycrystalline or samples, which are common in EPR studies, the resulting spectrum is a superposition of resonances from all orientations, often appearing as broadened features; these spectra are analyzed through to extract the principal g-values and their . Determination of the g-factor involves calibrating the EPR spectrometer against standard markers to ensure accurate field and frequency measurements. A widely used standard is 2,2-diphenyl-1-picrylhydrazyl (), which has a well-characterized g-factor of 2.0036, enabling determinations with errors typically below 0.0001 under controlled conditions such as low microwave power and stable temperature. In spectra exhibiting , an effective g-factor is briefly assessed from the center of the multiplet to account for the overall resonance position. Interpretation of the g-factor provides insights into the electronic structure of the paramagnetic . Organic radicals featuring an in a π orbital, such as those in conjugated systems, exhibit g-factors slightly greater than 2.0023 (often around 2.003), due to partial orbital contribution from spin-orbit mixing. In contrast, d¹ ions, like V⁴⁺ in vanadyl complexes, typically show g-factors less than 2 (e.g., ≈1.9), arising from the quenching of orbital angular momentum in ligand fields combined with negative spin-orbit coupling effects.

Hyperfine structure and multiplicity

In electron paramagnetic resonance (EPR) , arises from the magnetic interaction between the spin and the magnetic moments of nearby atomic nuclei possessing non-zero spin. This interaction splits the EPR signal into a multiplet of lines, providing detailed information about the local environment of the paramagnetic center, such as the number and type of coupled nuclei. The hyperfine interaction is described by the Hamiltonian term H_{hf} = \mathbf{A} \cdot \mathbf{I} \cdot \mathbf{S}, where \mathbf{I} and \mathbf{S} are the nuclear and electron spin operators, respectively, and \mathbf{A} is the hyperfine coupling tensor with units typically in MHz or mT. The tensor \mathbf{A} decomposes into isotropic and anisotropic components: the isotropic part, known as the Fermi contact interaction, originates from the s-electron spin density at the nucleus and is given by a_{iso} = \frac{2}{3} \mu_0 g_e g_n \mu_B \mu_n |\psi(0)|^2, while the anisotropic part arises from the classical dipolar interaction between the electron and nuclear magnetic moments. In powder or frozen solution spectra, the anisotropic contributions lead to broadened lineshapes, whereas in single crystals, they manifest as orientation-dependent splittings. For a paramagnetic species coupled to n equivalent nuclei of spin quantum number I, the hyperfine interaction produces a multiplicity of $2nI + 1 equally spaced lines, with relative intensities following the ./04%3A_Chemical_Speciation/4.08%3A_EPR_Spectroscopy) For example, coupling to a single proton (I = 1/2) results in two lines of equal intensity, while interaction with a single ^{14}\mathrm{N} nucleus (I = 1) yields three lines in a 1:1:1 ratio./04%3A_Chemical_Speciation/4.08%3A_EPR_Spectroscopy) Spectral simulation is essential for interpreting complex hyperfine patterns, particularly for systems with multiple nuclei. For simple cases like S = 1/2 coupled to I = 1/2, exact diagonalization of the Hamiltonian matrix yields the energy levels and transition probabilities. Software packages such as EasySpin facilitate these simulations by incorporating the full hyperfine tensor and allowing fitting to experimental data. Representative examples illustrate the scale of hyperfine couplings. In semiquinone radicals, proton hyperfine constants are typically around 0.5 mT, reflecting partial spin density on the adjacent carbon atoms. For the nitrogen dioxide (NO_2) radical, the ^{14}\mathrm{N} hyperfine coupling is approximately 5 mT, dominated by the isotropic Fermi contact term due to significant p-orbital spin density on the nitrogen. For weakly coupled nuclei where hyperfine constants are smaller than 0.1 mT, standard EPR resolution is often insufficient, but electron-nuclear double resonance (ENDOR) spectroscopy enhances sensitivity by directly measuring the nuclear frequencies, allowing precise determination of small couplings./Spectroscopy/Magnetic_Resonance_Spectroscopies/Electron_Paramagnetic_Resonance/ENDOR_-_Theory)

Linewidth and lineshape parameters

In electron paramagnetic resonance (EPR) spectroscopy, the linewidth, often denoted as \Delta B_{1/2}, represents the width of the absorption signal at half-maximum height (FWHM) or, in the first-derivative presentation common to continuous-wave EPR, the peak-to-peak separation \Delta B_{pp}. This parameter quantifies the broadening of the resonance line, which deviates from the ideal infinitely narrow transition due to various physical mechanisms. The lineshape, describing the overall form of the signal, is typically Lorentzian for homogeneous broadening, where all spins experience equivalent local fields, or Gaussian for inhomogeneous broadening arising from a distribution of resonance frequencies across the sample. Homogeneous broadening primarily stems from the transverse relaxation time T_2, which limits the lifetime of the excited spin state and introduces uncertainty in the energy levels via the time-energy relation. For a Lorentzian lineshape, the linewidth is given by \Delta B = \frac{\hbar}{g \mu_B T_2}, where g is the electron g-factor, \mu_B is the Bohr magneton, and \hbar is the reduced Planck's constant; shorter T_2 values yield broader lines. Inhomogeneous contributions include g-factor anisotropy, where orientational dependence of the g-tensor in powdered or crystalline samples disperses resonance fields; unresolved hyperfine interactions from nearby nuclei that overlap without clear splitting; and lattice strain, which induces local field variations in solids. Analysis of lineshapes involves techniques such as to separate overlapping or Gaussian components from experimental spectra, enabling isolation of broadening mechanisms. The second moment method calculates the mean-square deviation of the field from the resonance center, providing a quantitative measure of inhomogeneous broadening independent of lineshape assumptions. In solution-phase , rapid molecular motion or spin exchange can lead to exchange narrowing, where fast averaging of local fields reduces the effective linewidth, as described in the Anderson-Weiss theory; this effect is prominent in concentrated solutions, narrowing lines from potential Gaussian forms to near-. Typical EPR linewidths range from 0.01 mT in well-resolved solution spectra to 1 mT in solids, reflecting the balance of intrinsic relaxation and environmental broadening. Experimental artifacts, such as from excessive exceeding the natural linewidth, can distort lineshapes by introducing asymmetric broadening or derivative-like features, necessitating careful to preserve accurate \Delta B_{1/2} measurements.

Advanced Techniques

Continuous-wave vs. pulsed EPR

Continuous-wave electron paramagnetic resonance (CW-EPR) operates by applying a continuous field at a fixed while slowly sweeping the , recording the steady-state or of microwave energy by paramagnetic . This approach uses field modulation and phase-sensitive detection, typically at 100 kHz, to produce lineshapes with enhanced for detecting stable, steady-state in various samples. Pulsed EPR, in contrast, employs short, high-power pulses to selectively excite and manipulate the spin system, capturing time-resolved signals that reveal transient behaviors inaccessible to methods. A core technique is the Hahn echo, or two-pulse sequence (90°-τ-180°), where the initial 90° tips spins into the , they dephase over time τ, and the 180° refocuses them to form an echo, allowing direct measurement of the spin-spin relaxation time T₂ from the echo decay envelope. Pulsed methods excel in studying spin dynamics, such as times, by resolving time-domain evolution without the saturation or passage effects that distort spectra in samples with long relaxation times. Among advanced pulsed sequences, double electron-electron resonance (DEER) measures long-range distances between unpaired electrons, typically in the 2-8 nm range, by pumping one spin population at a different while observing dipolar in the echo of another, providing structural insights into biomolecular conformations. Additional examples include (FID), which detects the initial transverse magnetization decay after a single 90° pulse to probe rapid dephasing; electron spin echo envelope (ESEEM), a variation of the two- or three-pulse echo that reveals weak hyperfine interactions with surrounding nuclei through periodic modulations in the echo intensity; and inversion recovery (180°-t-90°-τ-180°), which quantifies the spin-lattice relaxation time T₁ by monitoring recovery of longitudinal magnetization over delay t. Pulsed EPR requires samples with sufficiently short spin-spin relaxation times, generally T₂ < 10 μs, to outrun instrumental dead time and capture signals before excessive decay, though refocusing echoes extend applicability to longer T₂. Commercial pulsed EPR instrumentation, such as Bruker's Elexsys series, became widely available in the post-1990s period, building on earlier prototypes from the 1980s and enabling routine time-resolved experiments. These techniques complement CW-EPR by directly accessing relaxation parameters like T₁ and T₂, which determine the viability of pulsed measurements.

High-field and high-frequency EPR

High-field and high-frequency electron paramagnetic resonance (EPR), operating at magnetic fields greater than 3 T and microwave frequencies above 95 GHz, significantly enhances spectral resolution compared to conventional lower-field techniques. This improvement arises from the increased Zeeman interaction, which better separates anisotropic components of the g-tensor and reduces overlap in hyperfine splitting patterns, enabling detailed characterization of paramagnetic centers in complex systems. Additionally, for transition metal ions and clusters with spin S > 1/2, high fields allow resolution of zero-field splittings that are otherwise obscured at lower fields, facilitating the study of and spin Hamiltonians. Key frequency regimes include the W-band at 94–95 GHz (corresponding to ~3.3 T), the D-band at ~263 GHz (~9.4 T), and terahertz EPR extending to 1 THz or higher (up to 36 T). These higher frequencies provide proportionally greater resolution, with the g-factor becoming particularly discernible, as the spread in resonance fields scales with frequency. Despite these benefits, high-field EPR encounters substantial challenges, primarily from the reduced skin depth of microwaves at elevated frequencies, which restricts penetration into conductive or metallic samples and increases resistive losses. Dielectric losses also intensify in aqueous or polar samples, such as those in biological contexts, degrading signal quality and necessitating careful sample preparation. To mitigate transmission inefficiencies in the millimeter- and sub-millimeter-wave regime, where standard waveguides are impractical, non-resonant bridges based on quasi-optical designs are employed, often using free-space propagation to bypass resonant cavity limitations. Hardware for these experiments relies on superconducting magnets to achieve the required fields, typically up to 15 T, with high homogeneity for precise features. delivery incorporates quasi-optical components, such as mirrors, lenses, and Fabry-Perot resonators, to focus Gaussian beams onto small samples and optimize coupling efficiency. In protein , these advancements have enabled techniques like relaxation-induced dipolar modulation enhancement (RIDME) in the , providing and for spin-labeled biomolecules at W- and D-bands. While sensitivity generally decreases at higher frequencies due to smaller sample volumes and losses, this is counterbalanced by enhanced selectivity, which isolates specific molecular and improves signal-to-noise for disordered systems like frozen protein solutions.

Applications

Chemical and materials characterization

Electron paramagnetic resonance (EPR) spectroscopy plays a crucial role in characterizing transient radical species during chemical reactions, enabling the identification of short-lived intermediates that are otherwise difficult to detect. In organic and inorganic chemistry, EPR is widely used to study radical mechanisms, such as those involving hydroxyl radicals (•OH), where spin trapping with 5,5-dimethyl-1-pyrroline N-oxide (DMPO) forms stable adducts detectable by their characteristic EPR spectra. For instance, the DMPO/•OH adduct exhibits a four-line spectrum with nitrogen hyperfine splitting constant aN ≈ 14.9 G and proton splittings aH ≈ 14.9 G (β-H) and 1.49 G (γ-H), allowing unambiguous identification of the radical type. This technique is particularly valuable for probing reaction kinetics, as the EPR signal intensity is directly proportional to the radical concentration, facilitating time-resolved studies of radical formation and decay rates. In , EPR excels at detecting and quantifying paramagnetic defects that influence electronic properties, such as dangling bonds in semiconductors. In silicon-based materials, silicon dangling bonds produce a narrow EPR signal at g ≈ 2.005, attributed to unpaired electrons localized on threefold-coordinated silicon atoms at interfaces or in amorphous regions. These defects act as deep traps, impacting carrier mobility and recombination, and their concentration can be correlated with variations through combined EPR and electrical measurements. For example, in hydrogenated (a-Si:H), light-induced EPR signals at g = 2.005 reveal metastable defects that degrade photovoltaic performance over time. Specific applications highlight EPR's versatility in chemical and materials analysis. In , EPR detects peroxy (ROO•) and alkyl (R•) radicals formed during oxidative or thermal breakdown, providing insights into chain scission and stabilization mechanisms; for , irradiation-induced radicals show dose-dependent EPR intensities that follow linear response up to high doses. In , EPR characterizes active sites, such as isolated Cu²⁺ ions in zeolites like Cu-CHA, where g-values (g∥ ≈ 2.22, g⊥ ≈ 2.05) and hyperfine couplings (A∥ ≈ 200 G) distinguish coordination environments influencing selectivity in reactions like reduction. For , EPR measures radiation-induced radicals in materials like , yielding linear dose-response curves (e.g., signal intensity proportional to dose up to 10⁴ ) for retrospective dose assessment in accidental exposures. Additionally, EPR using flow cells enables real-time monitoring of electrogenerated radicals during electrochemical processes, such as •OH formation at electrodes, by integrating the sample cell directly into the EPR cavity. Hyperfine interactions, like the nitrogen splitting aN ≈ G in aminoxyl radicals, further aid in distinguishing across these contexts.

Biological and medical studies

Electron paramagnetic resonance (EPR) spectroscopy has emerged as a vital tool in biological studies for characterizing paramagnetic centers in metalloproteins, such as iron in , where the high-spin ferric state exhibits a characteristic g-factor of approximately 6. This signal arises from the S=5/2 spin state of the iron, allowing researchers to probe the electronic environment and in oxygen-transport proteins without disrupting their native structure. Similar EPR signatures are used to investigate other metalloproteins, including copper-containing enzymes like , revealing insights into dynamics and metal-ligand interactions essential for enzymatic function. Site-directed spin labeling with nitroxide radicals attached to proteins or DNA enables the assessment of molecular mobility through correlation times derived from EPR spectral linewidths and rotational dynamics. In proteins, these labels report on local flexibility, with slower correlation times (on the order of nanoseconds) indicating restricted motion in folded regions, while faster times reflect unstructured loops. For DNA, nitroxide attachment via phosphorothioate linkages monitors base-pair dynamics and helix stability, providing data on conformational changes during replication or repair processes. Pulsed EPR techniques, such as double electron-electron (DEER), measure long-range distances between labels in biomolecules, typically resolving distributions up to about 10 with high precision. This method complements continuous-wave EPR by quantifying inter- distances in large assemblies, aiding in the structural modeling of protein complexes or folding. In studies of , DEER applied to spin-labeled amyloid-β peptides has mapped oligomer formation, showing parallel β-sheet structures in aggregates that contribute to . In medical applications, EPR oximetry non-invasively assesses oxygenation (pO₂) by monitoring the linewidth broadening of probes like nitroxides or , which is linearly sensitive to oxygen concentration. This technique, often employing exogenous contrast agents, enables real-time mapping of hypoxic regions in tumors or wounds, guiding therapeutic interventions. EPR quantifies radiation doses in radiotherapy for , using the stable free radicals induced in or pellets to verify delivered doses with sub-Gy accuracy. Recent advancements in the 2020s include compact, portable EPR systems for during clinical trials, facilitating superficial monitoring of dose distributions in patients undergoing .

Industrial and emerging uses

Electron paramagnetic resonance (EPR) spectroscopy plays a vital role in industrial , particularly for detecting through the identification of stable formed in materials like spices, fruits, and nuts. This method allows regulators to verify compliance with irradiation standards by analyzing the characteristic EPR signals of these radicals, which persist post-treatment and enable non-destructive testing of packaged goods. For instance, EPR spectra of irradiated samples reveal distinct cellulose radical features, facilitating accurate detection even after alcoholic pretreatments. Similarly, in , EPR quantifies nitrogen-related defects, such as NV centers, to classify diamond types (e.g., Type Ia, Ib, IIa) and distinguish natural from synthetic or treated stones based on paramagnetic center concentrations. This technique correlates EPR signal intensities with coloration and structural impurities, aiding in assessment. Benchtop EPR spectrometers, first developed in the late , have seen ongoing advancements enhancing portability for industrial applications, enabling on-site detection without large superconducting magnets. These compact systems, operating at X-band frequencies, support routine and in facilities, reducing costs and improving for non-specialist users. In emerging fields, NV centers in diamonds leverage EPR for quantum sensing, particularly magnetometry, where their spin properties enable nanoscale detection with sensitivities down to nanotesla levels. This approach has been optimized for ensemble measurements, promising applications in geophysical and biomedical diagnostics. EPR also monitors radical formation in lithium-ion batteries via in situ spectroscopy, tracking paramagnetic species during charge-discharge cycles to study degradation mechanisms and improve stability. Real-time EPR imaging visualizes distribution and radical concentrations, providing insights into growth and safety risks. In archaeology, EPR dating of burnt flint artifacts measures accumulated radiation-induced signals, such as E' centers, to estimate ages up to hundreds of thousands of years by assessing dose rates in heated samples. This technique has dated tools from sites like , offering chronological context for prehistoric human activity. Environmentally, EPR detects persistent free radicals in pollutants, such as those on particulate matter from combustion sources, linking them to and air quality issues. Quantification of these environmentally persistent free radicals (EPFRs) via EPR spectra reveals their stability and role in generating , informing mitigation strategies. Additionally, hybrid EPR-MRI systems integrate radical mapping with anatomical imaging, using dual-frequency resonators to overlay functional EPR data on MRI structures for enhanced visualization of paramagnetic probes .

References

  1. [1]
    [PDF] Short-short introduction to Electron Paramagnetic Resonance
    EPR spectroscopy measures energy differences between atomic/molecular states, mainly by the interaction of unpaired electrons with a magnetic field, using ...
  2. [2]
    [PDF] Basic EPR Theory
    EPR measures energy differences between atomic/molecular states. Spin states diverge with magnetic field, and hyperfine interactions provide information about ...Missing: fundamentals | Show results with:fundamentals
  3. [3]
    None
    ### Historical Background and Basic Principles of EPR
  4. [4]
    An overview of solid-state electron paramagnetic resonance ...
    EPR spectroscopy is used to study the nature of unpaired (paramagnetic) electrons in a given system, based on transitions between their spin states. Since ...
  5. [5]
    EPR in the USSR: the thorny path from birth to biological ... - PubMed
    In 1944, electron paramagnetic resonance (EPR) was discovered by Evgenii Konstantinovich Zavoisky in the USSR (Union of the Soviet Socialist Republics).
  6. [6]
    [PDF] Electron Paramagnetic Resonance in Enzymology - Aimin Liu
    EPR was first applied to biological materials in 1954 (1), 10 years after the discovery of an experimental approach to detect- ing electron spin resonance by ...
  7. [7]
    Electron paramagnetic resonance spectroscopy in radiation research
    Principle of EPR. Electron Paramagnetic Resonance is the resonant absorption of microwave radiation by paramagnetic systems in the presence of an applied ...
  8. [8]
    George Uhlenbeck and the Discovery of Electron Spin - Physics Today
    How two young Dutchmen, one with only o master's degree, the other a graduate student, made a most important finding in theoretical atomic physics.Missing: Zavoisky EPR
  9. [9]
    Celebration of 80 years of EPR - PMC - PubMed Central
    Jul 24, 2024 · EPR was developed as a means of measuring electron spin relaxation times. Theories of spin relaxation developed prior to the discovery of EPR ...
  10. [10]
    Discovery of Nuclear Magnetic Resonance: Rabi, Purcell, and Bloch
    Apr 10, 2020 · Gorter was Dutch physicist Cornelius Jacobus Gorter (1907-1980), who had tried to induce and detect NMR in bulk materials in 1936 at Groningen ...Nmr In Molecular Beams · Nmr In Bulk Materials · East Coast Nmr: Mit And...
  11. [11]
    The Physiologic Use of Lower Frequency Electron Paramagnetic ...
    As expected the EPR was born at low field. The development of radar during the preparation for WWII stimulated the development of higher frequency and ...
  12. [12]
    History of EPR Studies from the H.M. Swartz Laboratories: Part 1 ...
    Jan 5, 2022 · Our major developments were in the development of a very sensitive and reliable microwave bridge and in the development of suitable resonators.
  13. [13]
    Electron-nuclear double resonance - PMC - PubMed Central - NIH
    ENDOR spectroscopy is primarily directed to study the magnetic interactions of the unpaired electron spin with the spins of magnetic nuclei.Missing: post WWII Yoder Varian
  14. [14]
    Free Radicals in Biological Materials - Nature
    COMMONER, B., TOWNSEND, J. & PAKE, G. Free Radicals in Biological Materials. Nature 174, 689–691 (1954). https://doi.org/10.1038/174689a0. Download citation.
  15. [15]
    High field electron paramagnetic resonance spectroscopy under ...
    Aug 26, 2014 · Since the first high field EPR spectrometer installation in 198321 the use of frequencies ≥94 GHz has steadily increased and the first ...
  16. [16]
    Electron paramagnetic resonance spectroscopy using a single ...
    Mar 29, 2019 · In this paper, we demonstrate sensitive local EPR spectroscopy using a superconducting flux qubit. The target spin system is directly attached ...Missing: review | Show results with:review
  17. [17]
    A Low-Cost, Tabletop LOD-EPR System for Nondestructive ... - NIH
    Jun 7, 2025 · This work describes the performance of a low-cost, tabletop, longitudinally detected electron paramagnetic resonance (LOD-EPR) system to address this issue in ...
  18. [18]
    [PDF] Electron Paramagnetic Resonance Theory E. Duin - Auburn University
    This course manual will provide the reader with a basic understanding needed to be able to get useful information using the technique of electron paramagnetic ...Missing: history paper
  19. [19]
    [PDF] Simulating Frequency-Domain Electron Paramagnetic Resonance
    Jul 8, 2015 · ABSTRACT: We present a comparison of experimental and simulated frequency- and field-domain electron paramagnetic resonance (EPR) spectra of ...
  20. [20]
    [PDF] Fundamentals and Applications of Electron Paramagnetic ...
    Detection is limited by noise components. • Phase-sensitive amplitude detection. • Small amplitude sinusoidal field is modulated during a scan.
  21. [21]
    None
    Summary of each segment:
  22. [22]
    EPR - Interpretation - Chemistry LibreTexts
    Jan 29, 2023 · Electron paramagnetic resonance spectroscopy (EPR), also called electron spin resonance (ESR), is a technique used to study chemical species with unpaired ...
  23. [23]
    Electron Paramagnetic Resonance (EPR) - Zurich Instruments
    The cwEPR spectrometer (see image) applies a magnetic field of about 3500 G (0.35 T) and measures the absorption of microwaves in the 9-10 GHz regime (X-Band).Missing: strength | Show results with:strength
  24. [24]
    High Frequency EPR - Bruker
    Advantages of High Frequency EPR​​ MW-Nutation of Coal Sample at 94 GHz. The nutation period is 62 ns with the 400 mW High Power Upgrade unit.
  25. [25]
    [PDF] Terahertz EPR spectroscopy using a 36-tesla high-homogeneity series
    This study uses a 36-tesla magnet for high-field EPR, enabling high-resolution study of g-anisotropies and resolving signals from unpaired spins.
  26. [26]
    Development of a Hybrid EPR/NMR Coimaging System - PMC - NIH
    Typically, proton MRI and EPRI systems utilize very different magnetic field strengths. For proton MRI, it is generally desirable to work at high field ...
  27. [27]
    Active Shims for EPR Magnets - Bridge12
    Active shims are commonly used in high-field, superconducting NMR magnets to improve the magnetic field homogeneity. However, shimming EPR magnets is less ...
  28. [28]
    Rapid-Scan EPR Imaging - PMC - PubMed Central - NIH
    Superimposed on that slowly-varying field is magnetic-field modulation, often at a frequency of 100 kHz.
  29. [29]
    Magnettech ESR5000 |Benchtop - EPR Instruments - Bruker
    Field homogeneity, ± 5 μT (50 mG) within sample region. Field stability, 1.0 μT/h (10 mG/h). Magnetic field range, -10 to 650 mT (-100 to 6500 G). Modulation ...
  30. [30]
    [PDF] The operation of an electron paramagnetic resonance spectrometer
    A number of papers have discussed electron paramagnetic resonance (EPR) microwave bridge spectrometers and the signals obtained from them for substances ...
  31. [31]
    None
    ### Summary of Microwave Bridge, Sources, and Detection in CW-EPR
  32. [32]
    [PDF] 2. Practical Aspects - 2.1 The EPR Spectrometer
    11: Boltzmann distribution for the electrons in the S = - ½ level (n0) and the S = ½ level (n1). Fig. 12: Change in the electron distributions between the S = - ...
  33. [33]
    A Review of the Dawn of Benchtop EPR Spectrometers—Innovation ...
    To solve the issue, the time-tested klystron (which bore all the disadvantages of electro-vacuum devices) was replaced by a Gunn diode, which was a solid-state ...
  34. [34]
    High-Frequency EPR Instrumentation | Applied Magnetic Resonance
    Nov 15, 2009 · More convenient are solid-state sources, such as Gunn oscillators and IMPATT sources. The reference arm allows full control of absorption ...<|separator|>
  35. [35]
    Instrumentation for high-field dynamic nuclear polarization and ...
    Jan 29, 2025 · For high-frequency DNP applications either solid-state sources or gyrotrons are used to generate the required microwave radiation [10, 27, 28].
  36. [36]
    Uniform Field Resonators for EPR Spectroscopy: A Review - PMC
    Cavity resonators are often used for electron paramagnetic resonance (EPR). Rectangular TE102 and cylindrical TE011 are common modes at X-band even though ...
  37. [37]
    [PDF] EPR Resonators 5
    A microwave cavity is simply a metal box with a rectangular or cylindrical shape which resonates with microwaves much as an organ pipe resonates with sound ...Missing: T circulator
  38. [38]
    ER 4012ST | Standard EPR Resonator - Bruker
    The cavity operates in the TE102 mode with a center frequency of 9.8 GHz. It features a medium unloaded Q value of 6000. Optical access is provided by a ...Missing: factor | Show results with:factor
  39. [39]
    Dielectric resonator in rectangular ТЕ102 cavity for electron ...
    Mar 1, 2023 · Each part of the manufactured DR is characterized by low dielectric losses (Q = 5400) and do not have own EPR signals at room temperature. The ...
  40. [40]
    Multibore sample cell increases EPR sensitivity for aqueous samples
    EPR signal intensity S is proportional to the filling factor η and the unloaded quality factor of the resonant cavity S ∝ χ ″ η Q U ′ P 1 / 2 [1].
  41. [41]
    Resonators for Clinical Electron Paramagnetic Resonance (EPR)
    Aug 9, 2020 · To achieve a uniform sensitivity of EPR detection within the resonator, a homogeneous B1 field needs to be established. Three factors alter the ...
  42. [42]
    [PDF] Appendix E: Instrumentation and Technical Performance
    The microwave cw source, say, at 10 GHz, can be a klystron or a solid-state. (Gunn) diode. Its output in part goes to an on/off switch, computer-controlled ...
  43. [43]
    Sample Temperature System - Bruker
    3.8 K to 300 K with the Helium Temperature Control ... Bruker now introduces three cryogen-free EPR variable temperature systems for both CW and Pulse EPR ...Missing: inserts 4-300
  44. [44]
    [PDF] High-Field/High-Frequency EPR Spectrometer Operating in Pulsed ...
    The temperature can be set with a temperature con- trol unit (ITC503, Oxford Instruments) in the range of 4-300 K. For cooling the probe head, liquid helium ...<|separator|>
  45. [45]
    Aqueous flat-cells perpendicular to the electric field for use in ...
    Additional improvement in EPR signal of up to 75% can be achieved by using sample holder materials with lower dielectric constants. Introduction. A conventional ...Missing: biomaterials | Show results with:biomaterials
  46. [46]
    The evolution of biomedical EPR (ESR) - Lawrence J. Berliner, 2016
    Feb 1, 2016 · The aqueous flat cell was key to these and all future biochemical studies, which allowed the reproducible observation of aqueous samples.
  47. [47]
    Improved EPR sensitivity for aqueous biological samples using low ...
    Jun 7, 2025 · In this work, the design, fabrication, and testing of dielectric resonators and multi-channel aqueous sample cells for applications in X-band (nominally 9.5 ...Missing: flat | Show results with:flat
  48. [48]
    EPR Resonator for High Dielectric EPR Samples - Bruker
    The ER 4103TM has been developed particularly for studies with aqueous samples, tissue or samples exhibiting a high dielectric constant.Missing: biomaterials | Show results with:biomaterials
  49. [49]
    Quantitative Interpretation of Multifrequency EPR Spectra
    g-FACTOR ANISOTROPY. From the EPR spectrum and the resonance condition, the g-value of the spin system can be determined. The g-value is derived from ...
  50. [50]
    Use of EPR to Solve Biochemical Problems - PMC - PubMed Central
    A free electron in a vacuum has a g value (ge) of 2.00231930436153 (uncertainty 0.00000000000053), which is one of the most precisely known values in physics.
  51. [51]
    Detection of L-band electron paramagnetic resonance in the DPPH ...
    The resonance field increased linearly with frequency and the obtained Landé g factor of 1.999 ± 0.0197 is close to the accepted value of 2.0036, measured in ...
  52. [52]
    Exploring the Radical Nature of a Carbon Surface by Electron ... - NIH
    Apr 24, 2014 · For free electrons the g-value is 2.00232. Variations in the g-value from the 2.00232 are related to magnetic interactions involving the orbital ...
  53. [53]
    [PDF] EPR Spectroscopy - The University of Manchester
    (ii) transition metal ions: small ∆E (coloured; crystal field splittings) ... e.g. {VIVO}2+, d1 typical g ≈ 1.9. e.g. Cu2+, d9 typical g ≈ 2.1. ( )g ge.
  54. [54]
    Hyperfine Interaction - Electron Paramagnetic Resonance
    Hyperfine interaction couples electron and nuclear spins via dipole-dipole and Fermi contact interactions, generally resulting in an anisotropic interaction.
  55. [55]
    [PDF] Electron Paramagnetic Resonance: Hyperfine Interactions
    What is EPR (ESR)?. • Electron Paramagnetic (Spin) Resonance. • Applies to atoms and molecules with one or more unpaired electrons. • ...<|control11|><|separator|>
  56. [56]
    [PDF] EPR-Spectroscopy.pdf
    EPR spectrometers typically vary the magnetic field and hold the microwave frequency. EPR spectrometers are available in several frequency ranges, and X band is ...<|control11|><|separator|>
  57. [57]
    [PDF] EasySpin, a comprehensive software package for spectral ...
    Sep 26, 2005 · EasySpin provides extensive EPR-related functionality, ranging from elementary spin physics to data analysis. In addition, it provides routines ...
  58. [58]
    EPR, ENDOR, and TRIPLE resonance spectroscopy on the neutral ...
    Both EPR spectra show an emerging hyperfine splitting of 0.85 mT that could be assigned to the isotropic hyperfine coupling constant (hfc) of the proton at N(5) ...
  59. [59]
    Automated On-the-Fly Detection and Correction Procedure for EPR ...
    ... lineshape function that is later used for deconvolution. In order to improve the signal-to-noise ratio (SNR), it is a common practice to observe the first ...
  60. [60]
    Forty Years of the d1/d Parameter - IntechOpen
    A simple-measured d 1 /d parameter was used rather often during these years. By using a variety of biochemical and biophysical approaches, a helix packing model
  61. [61]
    Exchange Narrowing in Paramagnetic Resonance
    In this paper the problem of the line shape in paramagnetic resonance when large exchange interaction is present is discussed from the standpoint of a ...Missing: EPR | Show results with:EPR
  62. [62]
    Linewidth - an overview | ScienceDirect Topics
    Linewidths. The theoretical width of an ESR line is ∼0.01 mT and this may be approached in solutions of free radicals as shown in Figure 2.
  63. [63]
    Application of magnetic field over-modulation for improved EPR ...
    Magnetic field modulation in CW electron paramagnetic resonance (EPR) is used for signal detection. However, it can also distort signal lineshape.Missing: relaxation | Show results with:relaxation
  64. [64]
    Comparison of Continuous Wave, Spin Echo, and Rapid Scan EPR ...
    To acquire spectra with comparable signal-to-noise, both pulsed and rapid scan EPR require less time than conventional CW EPR. Rapid scan spectroscopy does not ...
  65. [65]
    None
    ### Pulsed EPR Techniques Summary
  66. [66]
    Pulsed Electron Double Resonance in Structural Studies of Spin ...
    This review deals with the application of the pulsed electron double resonance (PELDOR) method to studies of spin-labeled DNA and RNA with complicated spatial ...
  67. [67]
    elexsys e580 - Bruker
    Oct 29, 2019 · From the very beginning Bruker set the milestones in commercial Pulse-EPR instrumentation with the ESP 380 introduced in 1987. The next landmark ...Missing: date | Show results with:date
  68. [68]
    Order-of-magnitude SNR improvement for high-field EPR ... - Science
    Sep 20, 2023 · High-frequency EPR (frequencies above 100 GHz, fields above 3.5 T, “HFEPR”), however, is advantageous for increased resolution and is ...
  69. [69]
    High-field/High-frequency EPR Spectroscopy in Protein Research
    Dec 13, 2022 · Application of ever higher magnetic fields and microwave frequencies results in spectacular improvements of both spectral and time resolution.
  70. [70]
  71. [71]
    Terahertz EPR spectroscopy using a 36-tesla high-homogeneity ...
    The magnet's intrinsic homogeneity (25 ppm, that is 0.9 mT at 36 T over 1 cm diameter, 1 cm length cylinder) was previously established by NMR.Missing: μT | Show results with:μT
  72. [72]
    Spin trapping of superoxide and hydroxyl radical: Practical aspects
    Spin trapping uses nitrones to detect superoxide and hydroxyl radicals, but is inefficient for superoxide. DMPO is a versatile spin trap. Nitrones can form ...
  73. [73]
    A mini review: Fundamental aspects of spin trapping with DMPO
    This review describes fundamental aspects of spin trapping with DMPO, including basic knowledge for analyzing ESR spectra in biological samples.
  74. [74]
    The effects of vacuum-ultraviolet radiation on defects in low-k ...
    Oct 1, 2016 · The defect with the g factor of 2.0054–2.0050 is likely to be from the silicon-dangling bonds. The defect with the g factor of 2.0018–2.0020 is ...
  75. [75]
    Role of the paramagnetic donor-like defects in the high n-type ...
    Oct 15, 2020 · Traditionally, the conductivity of n-type in semiconductor materials was assigned for the existence of the intrinsic defects or impurities. The ...
  76. [76]
    Cluster-model calculations of hyperfine parameters and g values for ...
    We confirm from these calculations that the ESR signal with g=2.0055 and the g=2.0043- and 2.012-components of the light-induced ESR signal in a-Si:H originate ...Missing: 2.005 | Show results with:2.005
  77. [77]
    A review of experimental studies of the role of free-radicals in ...
    This paper reviews current understanding of the role of free radicals in the oxidation of polyethylene induced by high-energy irradiation (gamma or e-beam).
  78. [78]
    Coordination Environment of Copper Sites in Cu-CHA Zeolite ...
    Sep 10, 2014 · The well-defined copper sites in the six-membered ring of the CHA structure are found to be EPR active, to give two distinct sets of signals in ...
  79. [79]
    17O-EPR determination of the structure and dynamics of copper ...
    Jul 30, 2021 · The bonding of copper ions to lattice oxygens dictates the activity and selectivity of copper exchanged zeolites. By 17O isotopic labelling ...
  80. [80]
    A unique calibration curve for alanine EPR dosimetry systems
    The alanine EPR dose-response curve was proposed to follow a saturating exponential function. In this study the characteristic parameter D373of the ...
  81. [81]
    A continuous in-situ EPR electrochemical reactor as a rapid in-depth ...
    Numerous electrodes that can fit inside a commercial flat-cell or EPR-tube for in-situ radical production have been developed. These designs include electrodes ...
  82. [82]
    Correlation of radical structure with EPR spin adduct parameters
    Correlation of radical structure with EPR spin adduct parameters: utility of the proton, carbon-13, and nitrogen-14 hyperfine splitting constants of aminoxyl ...<|control11|><|separator|>
  83. [83]
    EPR spectroscopy of whole blood and blood components - NIH
    The other two EPR signals are from ferric haem in Hb. The signal at g≈6 is caused by the ferric haem in the high spin (S=5/2) state. Another ...
  84. [84]
    [PDF] Electron paramagnetic resonance spectroscopic studies of iron and ...
    This review highlights the unique application of EPR spectroscopy to the study of iron and copper in biological systems. Mention is made of a select number of ...Missing: hemoglobin | Show results with:hemoglobin
  85. [85]
    Nitroxide spin labels and EPR spectroscopy - ScienceDirect.com
    SDSL-EPR uses nitroxide spin labels attached to proteins, then studied by EPR to detect local conformational dynamics and measure distances between labels.
  86. [86]
    NITROXIDE SPIN-SPIN INTERACTIONS: Applications to Protein ...
    For commercially available EPR spectrometers, which op- erate in the 1- to 100-GHz frequency range, this corresponds to correlation times less than 1 ns.
  87. [87]
    Experimental mapping of DNA duplex shape enabled by global ...
    Aug 4, 2014 · Nitroxide spin labels, designated as R5a (Figure 1B), were site-specifically attached to DNAs using the phosphorothioate scheme as previously ...
  88. [88]
    Benchmark Test and Guidelines for DEER/PELDOR Experiments on ...
    Oct 19, 2021 · The study and the guidelines focus on proteins labeled with nitroxides and on double electron–electron resonance (DEER aka PELDOR) measurements.Introduction · PELDOR/DEER... · Guidelines · Deriving and Interpreting...
  89. [89]
    Structural characterization of biomolecules through atomistic ... - NIH
    Double Electron-Electron Resonance (DEER) is a popular technique that exploits attached spin-labels to probe the collective dynamics of biomolecules in a ...
  90. [90]
    Monitoring Alzheimer Amyloid Peptide Aggregation by EPR - PMC
    The aggregation of the β-Amyloid (Aβ) peptide into fibrils and ultimately plaques is the chief indicator of Alzheimer's disease. Information on the structure of ...
  91. [91]
    Application of In Vivo EPR for Tissue pO2 and Redox Measurements
    In order to measure pO2 using EPR oximetry, a detector (also referred to as a resonator) is placed on the surface of the skin above the tissue injected with the ...
  92. [92]
    Estimation of pO2 distribution in EPR oximetry - PMC - NIH
    A desirable property of the EPR oximetry in vivo is the ability to measure spatial distribution of oxygen. ... EPR signal linewidth to p O 2 [15]. These ...
  93. [93]
    Radiation dosimetry in human bone using electron paramagnetic ...
    Accurate measurements of dose in bone are required in order to improve the dosimetry of systemic radiotherapy for osseous metastases.
  94. [94]
    Prospective superficial EPR in-vivo dosimetry study during ...
    Oct 30, 2021 · In the present prospective IVD study, EPR dosimeters were applied superficially in order to monitor and verify total dose delivery at target, ...
  95. [95]
    Compact Electron Paramagnetic Resonance on a Chip ...
    Sep 26, 2024 · In this way, the region of best homogeneity of the magnetic field can be accessed to perform EPR measurements. The EPRoC and the permanent ...
  96. [96]
    Alcoholic extraction enables EPR analysis to characterize radiation ...
    Nov 19, 2014 · EPR analysis on various sample pretreatments in the irradiated spice samples demonstrated that the spectral features of the cellulose radical ...
  97. [97]
    Application of EPR spectroscopy to identify irradiated Indian ...
    Improved approaches based on the EPR technique explained in this study may be useful to identify irradiated products and become beneficial to food regulators ...
  98. [98]
    Diamonds | Bruker
    EMX series Electronic Paramagnetic Resonance (EPR) analyzers provide an unambiguous technique for quantifying nitrogen centers in diamonds, providing a tool ...
  99. [99]
    Correlation between EPR spectra and coloration of natural diamonds
    In this study, we developed a new EPR based technique for a quantitative measurement of N3 center and C-center in diamond through quantitative EPR spectroscopy.
  100. [100]
    A Review of the Dawn of Benchtop EPR Spectrometers—Innovation ...
    Sep 14, 2022 · In this paper, we have given a brief history of development of the benchtop EPR spectrometers in the USSR. For the Soviet science, this project ...
  101. [101]
    Sensitivity optimization for NV-diamond magnetometry
    Mar 31, 2020 · This review analyzes present and proposed approaches to enhance the sensitivity of broadband ensemble-NV-diamond magnetometers.
  102. [102]
    Nanoscale quantum sensing with Nitrogen-Vacancy centers in ...
    In this review, we introduce the field of ODMR spectroscopy of NV-nanodiamonds and how it can be used to sense different quantities.
  103. [103]
    Electron paramagnetic resonance imaging for real-time monitoring ...
    Feb 9, 2015 · Here we present the first report on Electron Paramagnetic Resonance operando spectroscopy and in situ imaging of a Li-ion battery.
  104. [104]
    Electron spin resonance dating of burned flint from Kebara Cave ...
    Electron spin resonance (ESR) measurements of heated flint from Kebara cave show the presence of two radiation sensitive signals: E' and Al.Missing: burnt | Show results with:burnt
  105. [105]
    ESR dating of ancient flints - Nature
    Mar 1, 1981 · We describe here ESR measurements on a damage signal which we have found to be present without exception in a wide variety of natural flints.Missing: burnt | Show results with:burnt
  106. [106]
    Quantification of environmentally persistent free radicals and ... - ACP
    In this study, we have detected and quantified a wide range of particle-associated radicals using electron paramagnetic resonance (EPR) spectroscopy.