Fact-checked by Grok 2 weeks ago

J -coupling

J-coupling, also known as scalar coupling or spin-spin coupling, is an indirect through-bond interaction between nuclear spins in a during () spectroscopy, mediated by the electrons in the intervening chemical bonds. This phenomenon arises from the magnetic influence of one on another via bonding electrons, resulting in the splitting of spectral lines into multiplets whose separation is defined by the J, typically expressed in hertz (Hz). The magnitude of J provides critical information about the number of bonds between coupled nuclei and the , making J-coupling a of for molecular structure determination. J-couplings are classified by the number of bonds separating the nuclei, with common types including geminal couplingJ, across two bonds, such as in H-C-H groups, often ranging from -15 to -10 Hz for aliphatic protons) and vicinal couplingJ, across three bonds, such as in H-C-C-H systems, typically 6-8 Hz for aliphatic chains). Longer-range couplings, like long-range coupling (⁴J or more), occur over four or more bonds and are smaller, usually less than 3 Hz, but can be significant in rigid or conjugated systems. Heteronuclear J-couplings, between different nuclear species (e.g., ¹H-¹³C), are also prevalent and often larger, aiding in assigning carbon-proton connectivities. The value of the coupling constant J depends on several molecular factors, including the between coupled nuclei for vicinal couplings, as described by the , which correlates ³J with torsional angles to infer and conformation. of adjacent atoms, bond hybridization, and further modulate J values, with electronegative substituents generally decreasing the magnitude of vicinal couplings. In cases of strong coupling, where the chemical shift difference approaches J, spectral patterns deviate from simple first-order multiplets, requiring advanced analysis techniques. In practice, J-coupling is indispensable for NMR applications in and biological , enabling the elucidation of molecular , stereochemical configurations, and dynamic behaviors through multiplet and 2D experiments like COSY or J-resolved . For instance, vicinal ³JHN-Hα values around 4-5 Hz indicate α-helical structures in proteins, while 8-9 Hz suggest β-sheets, facilitating refinement. These couplings also underpin quantitative NMR methods and optimizing synthetic designs in pharmaceuticals.

Fundamentals

Definition and Physical Origin

J-coupling, also known as scalar coupling, refers to the indirect through-bond interaction between nuclear magnetic moments in a , mediated by the electrons in the chemical bonds connecting the nuclei. This coupling is distinct from the direct dipolar interaction, which occurs through space without requiring intervening bonds. The effect manifests in (NMR) spectroscopy as a splitting of spectral lines, providing structural information about molecular connectivity. The physical origin of J-coupling lies in the hyperfine interactions between the spins and the surrounding , as first theoretically described by Norman Ramsey. These interactions arise from three primary contributions: the Fermi contact term, which dominates for couplings involving hydrogen nuclei like ^1H-^1H due to the s- density at the ; the magnetic dipole-dipole term, involving the orientation of and magnetic moments; and the orbital term, which accounts for the circulation of around the . The bonding transmit this interaction by polarizing their spin density in response to one , which then influences the local experienced by the other . The efficiency of this transmission depends on the hybridization of the intervening atoms (e.g., sp^3 in alkanes versus sp^2 in alkenes) and the bond angles, which modulate the overlap and delocalization of electron orbitals between the coupled . The magnitude of J-coupling is quantified by the coupling constant J (in Hz), which relates to the fundamental reduced coupling constant through the equation J = \frac{h \gamma_I \gamma_S}{2\pi} K, where h is Planck's constant and \gamma_I, \gamma_S are the magnetogyric ratios of the coupled nuclei. This relation isolates K as a measure of the electronic response independent of nuclear properties. K is derived from second-order , where the hyperfine perturbs the molecular wavefunction, yielding contributions from virtual excited states that mix electronic spin and orbital effects with the nuclear spins. In organic molecules, typical values for vicinal ^3J_{H-H} couplings across C-C single bonds are around 7 Hz, reflecting the average transmission through tetrahedral geometry./14%3A_NMR_Spectroscopy/14.12%3A_Coupling_Constants_Identify_Coupled_Protons)

Types of J-Coupling

J-couplings are classified primarily by the number of bonds separating the interacting nuclei, denoted as ^nJ, where n indicates the bond count. couplings (^2J) occur between nuclei attached to the same atom, such as two protons on a (H-C-H). Vicinal couplings (^3J) involve nuclei separated by three bonds, typically H-C-C-H in aliphatic chains. Long-range couplings (^4J and higher) span four or more bonds and are observed when aligns the nuclei favorably, such as in rigid or conjugated systems. Homonuclear J-couplings involve nuclei of the same isotope, like ^1H-^1H in organic molecules, while heteronuclear couplings connect different isotopes, such as ^1H-^{13}C or ^1H-^{19}F. Homonuclear ^1H-^1H couplings are common in proton NMR and provide structural insights through splitting patterns. Heteronuclear examples include the one-bond ^1J(^{1}H-^{13}C) in C-H groups, which exhibits large values due to the direct bond, and ^2J(^{1}H-^{19}F) in geminal H-C-F groups of fluorinated compounds, where the high gyromagnetic ratio of ^{19}F amplifies the coupling compared to other heteronuclei. The characteristics of J-couplings are influenced by , which can make nuclei magnetically equivalent and suppress observable splitting; restricted rotation, as in double bonds or rings, fixes dihedral angles and standardizes coupling magnitudes; and , where increased often enhances J values by 4-7% for vicinal and longer-range interactions in polar molecules like fluorobenzenes. Typical ranges for homonuclear ^1H-^1H couplings include geminal ^2J values of -15 to -10 Hz in aliphatic H-C-H (negative sign predominant) and 0-3 Hz in alkenes; vicinal ^3J around 6-8 Hz in flexible alkanes (positive), rising to 12-18 Hz for trans alkenes and 6-12 Hz for cis; and long-range ^4J of 1-3 Hz in aromatics. These vicinal magnitudes preview dependence on dihedral angles via the Karplus relation, with larger values for antiperiplanar orientations. Heteronuclear ^1J(^{1}H-^{13}C) spans 125-250 Hz, increasing with s-character, while ^2J(^{1}H-^{19}F) is 40-60 Hz.
TypeDescriptionTypical Range (Hz)Example MoleculeValue (Hz)
^2J (geminal, ^1H-^1H)H-C-H in -15 to -10CH_3CH_2- ()~ -12
^3J (vicinal, ^1H-^1H)H-C-C-H in derivative6-8CH_3-CH_2- (-like)~7
^3J (vicinal, ^1H-^1H)H-C=C-H trans in 12-18CH_2=CH_2 (trans analog)15-17
^3J (vicinal, ^1H-^1H)H-C=C-H cis in 6-12CH_2=CH_2 (cis analog)8-10
^1J (one-bond, ^1H-^{13}C)H-C in 125-135CH_3-CH_3 ()~125
^2J (geminal, ^1H-^{19}F)H-C-F40-60CH_3F~47
^4J (long-range, ^1H-^1H)H-CCC-H in aromatic1-3 (meta)~2

Spectral Manifestations

Multiplicity in NMR Spectra

In nuclear magnetic resonance (NMR) spectroscopy, J-coupling between magnetically nonequivalent nuclei leads to the splitting of signals into multiplets, providing key information on molecular connectivity. For first-order spectra, where the chemical shift difference between coupled nuclei (Δν) is much larger than the coupling constant (J, typically Δν/J > 10), the multiplicity follows the n+1 rule: a proton (or nucleus) coupled to n equivalent neighboring protons splits into n+1 equally spaced lines. This rule arises from the spin states of the neighboring protons, each of which can align with or against the external field, creating distinct energy levels for the observed nucleus. Common first-order patterns include the singlet for an isolated proton with no equivalent neighbors (n=0), the doublet for coupling to one neighbor (n=1, as in -CH-CH3 where the methine proton splits the methyl into a doublet), the triplet for two equivalent neighbors (n=2), and the quartet for three equivalent neighbors (n=3). A classic example is the ethyl group (-CH2-CH3) in ethanol, where the methyl protons (coupled to two methylene protons) appear as a triplet and the methylene protons (coupled to three methyl protons) as a quartet, separated by the vicinal ^3J coupling. The relative intensities of lines within these multiplets follow binomial coefficients, visualized by , which accounts for the statistical probabilities of spin alignments in homonuclear coupling to equivalent protons.
n (neighbors)MultiplicityRelative Intensities (Pascal's Triangle)
01
11 : 1
2Triplet1 : 2 : 1
3Quartet1 : 3 : 3 : 1
4Quintet1 : 4 : 6 : 4 : 1
For instance, a triplet's 1:2:1 ratio reflects the two central combinations (one up, one down) being twice as likely as the outer ones (both up or both down). This pattern holds for homonuclear cases like ^1H-^1H coupling in aliphatic chains. When Δν/J is smaller (typically <10), second-order effects distort the first-order patterns, leading to uneven spacing, extra lines, or "deceptively simple" multiplets that deviate from the n+1 rule. In such cases, the spectrum requires quantum mechanical analysis, as the simple vector model fails. A prominent example is the AA'BB' system in para-disubstituted benzenes (e.g., p-xylene), where the four aromatic protons form two pairs of chemically equivalent but magnetically inequivalent nuclei, resulting in two symmetrical doublets of doublets (often appearing as two doublets) due to ^4J meta and ^3J ortho couplings, rather than a simple first-order quartet. In symmetric molecules, virtual coupling can further complicate spectra, where an observed nucleus appears coupled to more protons than physically connected, due to near-equivalent pathways through magnetically inequivalent but chemically equivalent nuclei. This manifests as broadened or distorted multiplets, such as pseudo-triplets in -CH2-CH2- fragments of symmetric systems like 1,3-dichloropropane, misleading first-order interpretations.

Magnitude of J-Coupling

The magnitude of J-coupling constants in NMR spectroscopy is influenced by several molecular factors, primarily the dihedral angle between the coupled nuclei, but also substituent electronegativity, bond lengths, and hybridization states of the intervening atoms. The dihedral angle exerts the strongest effect on vicinal (³J) couplings, with the coupling constant reaching maxima when the nuclei are antiperiplanar (dihedral angle ≈180°) or synperiplanar (≈0°) due to optimal orbital overlap, and minima near 90° where overlap is poor. Electronegative substituents, such as oxygen or nitrogen, generally increase the magnitude of vicinal couplings when oriented gauche to the coupled protons but decrease them in trans orientations, while bond shortening enhances coupling through improved electron transmission. Hybridization affects the magnitude indirectly via changes in bond angles and lengths; for instance, sp²-hybridized carbons in alkenes yield larger ³J values (up to 18 Hz) compared to sp³ in alkanes (typically 4–12 Hz) owing to greater s-character in the bonds. The relationship between vicinal proton-proton couplings (³J_HH) and dihedral angle θ in H-C-C-H systems is empirically described by the : {}^3J_{\ce{HH}} = A \cos^2 \theta + B \cos \theta + C where A, B, and C are system-specific parameters accounting for substituent effects and hybridization. For aliphatic H-C-C-H fragments, typical parameters from the original formulation are A ≈ 9.0 Hz, B ≈ -0.5 Hz, and C ≈ -0.3 Hz, though refined versions incorporate electronegativity corrections, such as the , which adds terms like -2.32 cos θ Σ Δχ_i for substituent electronegativities (χ_i). These equations predict coupling values that oscillate with θ, enabling estimation of torsion angles from measured J. Computational and empirical data illustrate this dependence, as shown in the following table of approximate ³J_HH values for an unsubstituted ethane-like system using the basic parameters:
Dihedral Angle (θ)Approximate ³J_HH (Hz)
8.2
60°1.7
90°0.0
120°2.2
180°9.2
Isotope substitution also modulates J magnitudes through the dependence on gyromagnetic ratios (γ), as the coupling constant scales with the product γ_I γ_S in the expression J = (h / 2π) K γ_I γ_S, where K is the reduced coupling constant. Nuclei with larger γ, such as ¹H (γ = 42.58 MHz/T), yield larger J compared to heavier isotopes like ²H (γ ≈ 6.54 MHz/T) or ¹³C (γ ≈ 10.71 MHz/T), resulting in smaller observed J for deuterated or carbon-coupled systems despite similar K; for example, ¹J_HD is about 1/6.5 of ¹J_HH due to the γ ratio. Experimentally, J magnitudes vary systematically with bond order: one-bond couplings like ¹J_CH range from 120–200 Hz, reflecting strong direct transmission in C-H bonds of varying hybridization (e.g., 125–140 Hz for sp³, 160–170 Hz for sp²). Geminal (²J) proton-proton couplings span -20 to +20 Hz, often negative in hydrocarbons but positive when electronegative atoms intervene. Long-range couplings (⁴J and beyond) are typically small, less than 2 Hz, and diminish rapidly with increasing bond separation unless rigid conformations or π-systems enhance transmission.

Sign of J-Coupling

The sign of a J-coupling constant in NMR spectroscopy is defined by the convention that J is positive when the energy of a nucleus (e.g., spin A) is lower in states where its coupled partner (e.g., spin X) has the opposite spin orientation (αβ or βα), and negative when the spins are parallel (αα or ββ). This convention aligns with the physical origin from the Fermi contact mechanism, where one-bond couplings like are positive, two-bond geminal couplings in sp^{3} CH_{2} groups are typically negative (around -10 to -20 Hz), and three-bond vicinal couplings are positive, particularly for antiperiplanar arrangements in hydrocarbons (up to 12-15 Hz). For instance, in ethane derivatives, the staggered conformation yields a positive due to favorable orbital overlap in the antiperiplanar geometry. In first-order NMR spectra, where the chemical shift difference Δν greatly exceeds the coupling magnitude (|Δν/J| >> 10), the sign of J is unobservable because the splitting pattern depends solely on the |J|, rendering multiplets symmetric regardless of . This limitation necessitates advanced techniques for sign determination, which become essential in complex . The sign in the J-coupling term 2π J \mathbf{I}_1 \cdot \mathbf{I}_2 reflects this polarity, influencing energy level splittings in non-isolated systems. Determination of the J sign often relies on spectral simulation of second-order systems, where |Δν/J| < 10 leads to asymmetric multiplets (e.g., distorted doublets or "roofing" effects) whose patterns vary with the relative signs of couplings; iterative fitting to simulated spectra reveals the polarity by matching observed line intensities and positions. Heteronuclear comparisons provide another approach, such as analyzing ^{13}C satellites in proton spectra or using E.COSY-type 2D experiments to extract relative signs between homonuclear J_{HH} and known heteronuclear J_{CH}, assuming the latter's positive sign from one-bond conventions. Isotope shifts offer a complementary method; for example, secondary deuterium isotope effects on coupling constants (e.g., reduced ^{3}J_{HH} by ~0.1-0.5 Hz upon D substitution) can confirm signs by comparing isotopomer spectra, as the effect's direction depends on the original J polarity. For absolute signs, oriented media like liquid crystals allow order matrix calculations to disentangle J from dipolar contributions, directly yielding polarity. In rigid molecules, such as cyclic or constrained systems, the sign of J distinguishes syn (often small or negative ^{3}J) from anti (positive ^{3}J) couplings, aiding stereochemical assignment beyond magnitude alone; for example, negative ^{2}J_{HH} in CH_{2} groups confirms geminal interactions in sp^{3} environments, while positive vicinal J supports trans or antiperiplanar geometries. These techniques are crucial for conformational analysis, though they require high-resolution data and computational support to avoid ambiguities in overlapping signals.

Theoretical Description

Vector Model

The classical vector model provides an intuitive semi-classical description of J-coupling effects in nuclear magnetic resonance (NMR) spectroscopy, visualizing how the interaction between nuclear spins influences precession frequencies and leads to spectral splitting. In this framework, the magnetic moment of a neighboring nucleus creates a local magnetic field at the observed nucleus through electron-mediated polarization, effectively shifting the precession frequency of the observed spin vector. For a spin-1/2 nucleus I coupled to a neighboring spin-1/2 nucleus S, the two possible alignments of S—denoted as α (m_S = +1/2, parallel to the external field B_0) and β (m_S = -1/2, antiparallel)—produce opposing local fields at I. This results in two distinct precession frequencies for the I-spin magnetization: ν_I^α = ν_I + (1/2)J_{IS} when S is in the α state, and ν_I^β = ν_I - (1/2)J_{IS} when S is in the β state, where ν_I is the uncoupled Larmor frequency and J_{IS} is the scalar coupling constant in Hz. The J-coupling arises from a through-bond interaction that is slow compared to the Larmor precession (typically J << ν_0), allowing the nuclear spin states to remain effectively static on the NMR timescale while the intervening electrons average rapidly, yielding a time-independent effective local field. This local field can be understood as originating from the hyperfine interaction term in the Hamiltonian, approximated semi-classically as a torque on the I-spin vector due to the J \mathbf{I} \cdot \mathbf{S} coupling. In vector terms, the torque manifests as a differential precession: the I magnetization vector experiences an additional rotation rate proportional to the projection of S along the field direction, causing the ensemble of I vectors to fan out or diverge at relative rates of ±πJ_{IS} during free evolution. For equal populations of α and β states (as in thermal equilibrium for spin-1/2 systems), this divergence produces two equally intense components of the I magnetization precessing at the shifted frequencies, observable as a doublet in the spectrum. (from Ernst et al., Principles of Nuclear Magnetic Resonance in One and Two Dimensions, 1987) To illustrate in a simple heteronuclear AX system (where A and X have large chemical shift separation Δν_{AX} >> ), consider the A-spin after a 90° pulse, initially aligned transverse to B_0. The A vectors split into two subpopulations: half precess at ν_A + (1/2) (coupled to X=α), and half at ν_A - (1/2) (coupled to X=β). of these components yields a pattern centered at ν_A, with line separation , as the transverse components from each subpopulation interfere constructively at the offset frequencies. This derivation follows from resolving the into in-phase components and tracking their phase evolution under the weak-coupling approximation, where the chemical shift dominates over coupling, preventing significant mixing of states. The model thus explains the first-order multiplicity without invoking full quantum mechanics, though it relies on the Hamiltonian's scalar term for the underlying torque. This vector description is limited to the weak-coupling regime (Δν/J > 10), where the overwhelms the , ensuring the states evolve independently; in strong-coupling cases (Δν/J ≈ 1), quantum effects like virtual transitions distort the simple splitting, requiring treatments.

J-Coupling Hamiltonian

The J- term in the () describes the indirect interaction between mediated by bonding electrons. For isotropic , prevalent in solution-state NMR due to rapid molecular tumbling, the takes the form \hat{H}_J = 2\pi \hbar \sum_{I < S} J_{IS} \mathbf{I}_I \cdot \mathbf{I}_S, where \mathbf{I}_I and \mathbf{I}_S are the dimensionless spin operators for nuclei I and S, J_{IS} is the scalar coupling constant in hertz (Hz), \hbar is the reduced Planck's constant, and the sum runs over unique pairs of interacting spins. This bilinear form arises because the interaction energy is proportional to the dot product of the nuclear magnetic moments, \vec{\mu}_I \cdot \vec{\mu}_S = \gamma_I \gamma_S \hbar^2 \mathbf{I}_I \cdot \mathbf{I}_S, where \gamma_I and \gamma_S are the magnetogyric ratios; the observed J_{IS} thus incorporates these heteronuclear factors as J_{IS} = \frac{\hbar \gamma_I \gamma_S}{2\pi} K_{IS}, with K_{IS} the reduced coupling constant independent of nuclear properties. In anisotropic environments, such as solids or oriented media, the isotropic form generalizes to a tensorial expression \hat{H}_J = 2\pi \hbar \sum_{I < S} \mathbf{I}_I^T \cdot \overleftrightarrow{J}_{IS} \cdot \mathbf{I}_S, where \overleftrightarrow{J}_{IS} is the second-rank coupling tensor, whose isotropic part is J_{IS} = \frac{1}{3} \mathrm{Tr}(\overleftrightarrow{J}_{IS}) and anisotropic components reflect directional dependencies. The full tensor originates from four mechanisms in : (dominant for covalent bonds), spin-dipolar, paramagnetic spin-orbit, and diamagnetic spin-orbit contributions, each computed as tensor elements. This operator derives from second-order perturbation theory applied to the molecular electronic Hamiltonian perturbed by nuclear magnetic moments. The unperturbed Hamiltonian \hat{H}^{(0)} yields excited states |n\rangle with energies E_n > E_0, and the perturbation \hat{H}^{(1)} includes terms linear in the nuclear moments \vec{m}_I and \vec{m}_S. The second-order energy correction for the ground state |0\rangle is E^{(2)} = -\sum_{n \neq 0} \frac{|\langle 0 | \hat{H}^{(1)}(\vec{m}_I) | n \rangle|^2}{E_n - E_0} + \text{cross terms}, leading to the \frac{1}{2} \sum_{kl} K_{kl,IS} m_{I,k} m_{S,l}, where indices k,l denote Cartesian components and K encapsulates orbital responses via sums over molecular orbitals. This links directly to the reduced constant K_{IS}, which is basis-set independent in the non-relativistic limit and quantifies the electronic response without \gamma factors, typically on the order of $10^{19} T^{-2} A^{-2} for one-bond couplings. In high-field NMR, where the Larmor |\omega_I - \omega_S| \gg J_{IS}, the secular approximation simplifies the operator by truncating non-resonant terms that cause forbidden transitions. For heteronuclear pairs, the effective retains only the longitudinal and transverse components aligned with the external : \hat{H}_J^\mathrm{sec} = 2\pi \hbar J_{IS} \left( I_{I,z} I_{S,z} + \frac{1}{2} (I_{I,x} I_{S,x} + I_{I,y} I_{S,y}) \right) = 2\pi \hbar J_{IS} \left( I_{I,z} I_{S,z} + \frac{1}{2} (I_{I,+} I_{S,-} + I_{I,-} I_{S,+}) \right), preserving the dominant spectral effects while neglecting I_{I,x} I_{S,y} - I_{I,y} I_{S,x} terms, which average to zero under the rotating frame. The units of J_{IS} remain in Hz, reflecting the frequency splitting in spectra, and the \gamma factors ensure comparability across isotopes (e.g., larger J for high-\gamma pairs like ^1\mathrm{H}-^1\mathrm{H}). Modern predictions of J_{IS} and \overleftrightarrow{J}_{IS} rely on (DFT), often via coupled-perturbed Kohn-Sham equations to solve for response functions. For instance, BLYP functionals with augmented basis sets like aug-cc-pVTZ achieve mean absolute errors of 0.5–2 Hz for ^1\mathrm{H}-^1\mathrm{H} couplings in small organics, while for heteronuclear ^1J(\mathrm{H,F}) in HF, DFT yields 553 Hz (Fermi contact: 355 Hz, paramagnetic spin-orbit: 204 Hz) versus experimental 529 Hz (gas phase, plus vibrational corrections of 26–37 Hz). like B3LYP improve accuracy for transition-metal complexes, with errors under 1 Hz for one-bond ^1J(\mathrm{C,H}), enabling reliable structural predictions when benchmarked against coupled-cluster references.

Experimental Techniques

Measurement and Analysis

In one-dimensional (1D) NMR spectroscopy, J-coupling constants for protons are commonly extracted through spectral fitting of multiplicity patterns, where splittings adhere to the n+1 rule in systems, allowing direct measurement of vicinal or couplings in isolated resonances. However, this approach falters in crowded spectra due to overlapping signals that distort multiplet intensities and complicate . To address overlaps, two-dimensional (2D) correlation (COSY) identifies coupled partners via cross peaks, with multiplet displacements in phase-sensitive COSY spectra yielding homonuclear J values through pattern analysis. The 2D J-resolved experiment further simplifies analysis by projecting s onto one axis and J-couplings onto the perpendicular axis, enabling precise readout of multiple couplings from a single tilted multiplet without chemical shift interference. For heteronuclear J-couplings, particularly in isotopically labeled samples, the exclusive (E.COSY) correlates active and passive couplings via a third , producing displaced doublets that isolate the desired J value with high accuracy, even for small heteronuclear interactions below 1 Hz. Quantitative J-heteronuclear multiple bond (J-HMBC) adapts the standard HMBC pulse sequence to quantify long-range heteronuclear couplings (typically 2–10 Hz) by monitoring antiphase signal modulation across a fixed evolution period, offering sensitivity for sparse correlations in natural abundance samples. Specialized software enhances measurement reliability through and iterative fitting; for instance, MestReNova's module generates theoretical 1D/ spectra from user-defined chemical shifts and J matrices, allowing least-squares optimization against experimental to refine couplings. SpinWorks similarly supports system definition via a J-coupling editor, simulating complex ABX or AA'BB' patterns for iterative matching to observed spectra, particularly useful for second-order effects. Key challenges include severe overlap in spectra of large or symmetric molecules, which broadens effective linewidths and reduces resolution of , often requiring higher fields or selective excitation. Additionally, J-couplings exhibit temperature dependence due to conformational equilibria altering angles, with vicinal ³J_HH values potentially varying by 1–2 Hz over 50 K in flexible systems like amides. Error analysis reveals precision limits of approximately 0.1 Hz for proton J-couplings under optimal conditions (linewidth <0.5 Hz, high signal-to-noise), constrained by digital resolution, phase errors, and baseline distortions, though sub-0.05 Hz accuracy is achievable in resolved 2D data. These techniques reference basic multiplicity rules from 1D spectra for initial estimates.

Decoupling Methods

Decoupling methods in nuclear magnetic resonance (NMR) spectroscopy are employed to suppress interactions between nuclei, resulting in simplified spectra where multiplet structures collapse into singlets. This simplification enhances spectral resolution and sensitivity, particularly for low-abundance nuclei like ^{13}C, by removing the effects of heteronuclear scalar couplings. The primary approach involves applying radiofrequency (RF) irradiation to the coupled nucleus, which rapidly flips its spin states and averages the J-coupling term in the Hamiltonian, effectively setting the expectation value \langle \mathbf{I} \cdot \mathbf{S} \rangle = 0 over the irradiation period. Modern NMR spectrometers require dedicated decoupler channels—separate RF amplifiers and probe circuitry—to deliver this irradiation concurrently with observation of the target nucleus, enabling efficient heteronuclear decoupling without interfering with the acquisition signal. Broadband decoupling techniques provide uniform suppression across a wide chemical shift range, making them essential for routine ^{13}C NMR of organic compounds where proton decoupling (^{13}C{^1H}) is standard. Seminal pulse sequences like , developed through computer optimization of phase-cycled rectangular pulses, achieve effective decoupling bandwidths exceeding 50 kHz at moderate RF powers, outperforming earlier methods like by reducing heating and improving uniformity. Similarly, (globally optimized alternating-phase rectangular pulses) extends this capability with adaptive phase modulation, offering robust performance over bandwidths up to 100 kHz and minimal artifacts in multidimensional experiments. These sequences are widely implemented in inverse-gated or gated decoupling modes to control nuclear Overhauser effect () buildup for quantitative analysis. Selective decoupling targets specific resonances to isolate J-couplings for precise measurement, contrasting with broadband methods by using narrowband RF irradiation on a chosen spin. In one-dimensional (1D) selective decoupling experiments, irradiation of a particular proton collapses only the associated multiplets in the observed spectrum, allowing direct extraction of coupling constants like ^nJ_{CH} without interference from other interactions. This approach is particularly useful for resolving ambiguities in crowded spectra, such as distinguishing geminal from vicinal couplings in complex molecules. Despite their utility, decoupling methods introduce artifacts that must be mitigated. NOE buildup during prolonged irradiation enhances signal intensity through through-space dipolar relaxation but can distort quantification if not gated off during the relaxation delay, as seen in where proton saturation boosts carbon signals by up to 200%. Decoupling sidebands, arising from incomplete averaging or asynchronous modulation, appear as spurious peaks offset from the centerband and can mimic true signals in multidimensional spectra; optimization of pulse phases or acquisition timing suppresses these by factors exceeding 1000. Power dependence further complicates application, as higher RF powers (typically several hundred watts for broadband sequences in solution NMR) are needed for effective averaging but risk sample heating and hardware limitations, especially in biological systems. Modern variants like adiabatic decoupling address challenges at high magnetic fields (>14 T), where chemical shift dispersions demand broader bandwidths with lower power to avoid RF inhomogeneity. These methods use chirped RF pulses that follow adiabatic trajectories in the spin space, achieving decoupling efficiencies over 1 MHz bandwidth at RF fields as low as 1 kHz, thus enabling applications in biomolecular NMR without excessive heating.

Applications

Chemical Structure Elucidation

In (NMR) spectroscopy, the multiplicity arising from J-coupling provides critical information for identifying CHₙ groups and the number of adjacent atoms in a . A proton signal split into n+1 lines indicates coupling to n equivalent neighboring protons, allowing distinction between methyl (CH₃, typically a doublet if coupled to one proton), methylene (CH₂, triplet if coupled to two), and methine (CH, doublet if coupled to one) groups. The magnitude of the vicinal ³J coupling constants further refines this identification; for example, ³J values around 6-8 Hz are common for aliphatic CH-CH couplings, while smaller ²J values (≈12-15 Hz) confirm methylene protons on the same carbon. Homonuclear correlation spectroscopy, such as COSY, leverages J-coupling to through-bond connectivities between protons up to three or four bonds apart, revealing the skeletal of the . Cross-peaks in a COSY spectrum appear between protons that share a J-coupling, enabling the tracing of spin systems like -CH₃-CH₂- chains where the methyl correlates with the methylene . This technique is particularly valuable for resolving overlapping signals in complex mixtures, confirming adjacency without relying solely on chemical shifts. The combination of J-coupling patterns and chemical shifts facilitates complete proton assignments in small molecules, distinguishing structural isomers by their unique coupling networks. For instance, n-butane (CH₃-CH₂-CH₂-CH₃) exhibits a classic AA'BB'XX'XX' pattern with ³J vicinal s of ≈7 Hz between the methyl and methylene protons, resulting in a triplet for the CH₃ and multiplet for the CH₂ groups, whereas ((CH₃)₃CH) shows a decuplet (or multiplet) for the central CH proton (coupled to nine equivalent hydrogens from three methyl groups with ³J ≈7 Hz) and a for the nine methyl protons, highlighting the branched connectivity. Integrating these with characteristic chemical shifts (e.g., ≈0.9 for CH₃) allows unambiguous structure determination. In biomolecular applications, J-coupling aids peptide sequencing by assigning amide NH to alpha CH protons within residues. The vicinal ³J_{NHα} coupling constants, measured via quantitative J experiments, correlate with the phi (φ) through the , helping identify types and sequential order in short ; for example, values >8 Hz indicate β-sheet-like extensions, while <5 Hz suggest α-helical turns, enabling residue-by-residue mapping when combined with COSY correlations. This approach was pivotal in early NMR studies of polypeptides, such as , where ³J_{NHα} data confirmed the primary sequence connectivity.

Conformational Analysis

The Karplus relationship provides a foundational tool for inferring angles from measured vicinal ^3J coupling constants in NMR , enabling the determination of molecular conformations in biomolecules such as proteins and carbohydrates. In proteins, ^3J_{Hα-HN} and ^3J_{Hα-Hβ} couplings are particularly useful for estimating backbone φ and χ_1 angles, respectively, with parameterized equations refined for specific residue types to improve accuracy in secondary structure assignment. For carbohydrates, analogous Karplus equations relate ^3J_{H-H} values across glycosidic linkages to torsion angles like those in the ^1C_4 chair conformation, aiding in the elucidation of ring puckering and linkage . These applications rely on empirical calibrations derived from quantum mechanical calculations and experimental data, ensuring reliable correlations within defined and substituent contexts. Representative examples illustrate how J-coupling differences distinguish conformational preferences. In alkenes, the vicinal ^3J_{H-H} coupling across the is approximately 10 Hz for cis isomers (dihedral angle ~0° or 120°) and 17 Hz for trans isomers (dihedral angle ~180°), allowing straightforward stereochemical assignment without additional techniques. For derivatives locked in conformations, such as tert-butylcyclohexane, axial-axial ^3J couplings are larger (~8-12 Hz) compared to axial-equatorial (~3-5 Hz) or equatorial-equatorial (~2-4 Hz) due to optimal torsional alignment in the former, confirming the preference for equatorial substituents. These variations stem directly from the cosine dependence in the Karplus framework, highlighting J-coupling's sensitivity to local geometry. In flexible molecules, rapid conformational interconversion leads to time-averaged J-coupling values, complicating static interpretations but revealing when studied via variable-temperature (VT) NMR. For instance, in monosubstituted cyclohexanes undergoing flips on the timescale at , observed ^3J averages (~7 Hz) reflect equal populations of axial and equatorial forms; cooling to -80°C slows exchange, signals and yielding distinct couplings for each conformer, from which activation barriers (~10-12 kcal/mol) can be calculated using line-shape . Similar VT-NMR approaches in peptides quantify rotamer populations around χ_1 angles by monitoring ^3J_{Hα-Hβ} changes, providing insights into side-chain essential for function. This averaging underscores J-coupling's role as a probe of motional timescales relative to the NMR experiment. Advanced applications extend J-coupling analysis through residual dipolar couplings (RDCs) in partially media, such as bicelle solutions, to determine absolute orientations and long-range conformations. RDCs manifest as modifications to scalar J-couplings (total splitting = J + D), where the dipolar component encodes internuclear vector orientations relative to the alignment tensor, enabling tensor fitting for global structure validation in proteins and small molecules. In protein NMR, ^1D_{NH} RDCs (~10-30 Hz in weak alignment) refine orientations and loop geometries beyond scalar J alone, as demonstrated in studies where multiple alignments resolve ambiguities in flexible regions. For carbohydrates, RDCs in fragments confirm anomeric configurations and chain helicity in solution. This method's power lies in its sensitivity to overall molecular alignment, complementing vicinal J for comprehensive conformational mapping. A key limitation of J-coupling-based conformational arises from signal averaging over multiple populated conformers, which can obscure individual contributions and lead to erroneous estimates if populations are unequal or is . In disordered proteins or flexible glycans, this results in effective J values that require population modeling or relaxation dispersion experiments for , potentially underestimating dynamic heterogeneity without supporting data like NOEs or RDCs.

Historical Context

Early Observations

The discovery of (NMR) in the post-World War II era laid the groundwork for observing subtle interactions like J-coupling. In 1946, and colleagues at detected NMR signals in liquids, while Edward M. Purcell and his team at Harvard observed them in solids, earning them the 1952 for these foundational experiments. These early successes prompted investigations into higher-resolution spectra, but initial instruments operated in continuous-wave () mode at low magnetic fields around 30 MHz for protons, limiting and due to magnet inhomogeneities and broad linewidths exceeding 1 Hz. The first reported observation of J-coupling occurred in 1950, when W. G. Proctor and F. C. Yu reported a multiplet in the 121Sb NMR spectrum of sodium hexafluoroantimonate (NaSbF6) in aqueous solution, later recognized as due to 121Sb-19F heteronuclear J-coupling. This marked the initial empirical evidence of spin-spin interactions through chemical bonds, though the researchers initially linked it to variations before fully interpreting it as coupling. Early spectrometers struggled with such , as field instabilities often obscured multiplets narrower than 10-20 Hz. In the early , H. S. Gutowsky and C. J. Hoffmann confirmed and expanded on these findings through studies of proton spectra in fluorine-containing compounds, including interpretations of 1H-19F couplings that revealed multiplet patterns. Separately, Gutowsky, D. W. McCall, and C. P. Slichter observed the first clear 1H-1H homonuclear J-coupling in the proton spectrum of , showing a triplet and quartet with ~7 Hz splittings at 30 MHz, attributed to vicinal proton interactions. These observations highlighted J-coupling's role in revealing molecular connectivity, despite challenges like rapid OH exchange broadening the alcohol proton signal. A pivotal 1951 publication by Gutowsky, alongside D. W. McCall and C. P. Slichter, provided a detailed analysis of J-coupling among nuclear magnetic dipoles in molecules, using ethanol and other liquids as examples to demonstrate how such interactions produce observable spectral fine structure and implications for molecular structure determination. This work solidified J-coupling as a key phenomenon in high-resolution NMR, transitioning the field from basic detection to structural analysis amid ongoing instrumental limitations.

Theoretical Advancements

In the and , theoretical understanding of J-coupling advanced through Norman Ramsey's formulation, which linked the phenomenon to hyperfine interactions between nuclear spins and orbital currents in molecules. Ramsey's decomposed J into four contributions: the diamagnetic spin-orbit, paramagnetic spin-orbit, spin-dipolar, and Fermi contact terms, providing a foundational quantum mechanical framework for interpreting observed splittings. Among these, the Fermi contact term emerged as dominant for many vicinal and couplings, particularly in organic molecules, due to its dependence on s-electron density at the nuclei, which facilitates efficient spin polarization transmission through bonds. This era's progress was summarized in Herbert Gutowsky's 1969 review, which highlighted the Fermi contact mechanism's role in explaining the signs and magnitudes of J values across diverse systems. The 1960s saw empirical refinements that bridged theory and experiment, notably through Karplus's 1963 development of angle-dependent relations for vicinal proton couplings in compounds. The , J = A cos²θ + B cosθ + C (where θ is the H-C-C-H and A, B, C are empirically fitted constants), correlated J magnitudes with molecular conformation, enabling conformational analysis without full quantum computations. These curves, initially derived for hydrocarbons, were extended to heteroatom-substituted systems, emphasizing the Fermi contact term's sensitivity to bond angles and , and became a staple for interpreting spectral data in complex molecules. From the to the , ab initio quantum chemical methods enabled direct computation of J-coupling constants, shifting from empirical models to predictive theory. Early coupled Hartree-Fock (CHF) calculations reproduced experimental J values for small molecules like HD and H₂O, isolating contributions from each Ramsey term with basis sets of moderate size. The introduction of (DFT) in the mid-1990s further accelerated progress, offering efficient scaling for larger systems while maintaining accuracy comparable to post-Hartree-Fock methods; for instance, DFT with gauge-independent atomic orbitals (GIAOs) predicted vicinal ³J_HH couplings in alkanes within 1-2 Hz of experiment. Coupled-cluster approaches, such as CCSD, enhanced precision by incorporating electron correlation effects, achieving errors below 1 Hz for one-bond couplings in first-row hydrides and marking a milestone in theoretical NMR prediction. In the and beyond, advancements addressed relativistic effects critical for heavy-atom systems, where scalar relativistic Hamiltonians like the zeroth-order regular (ZORA) revealed enhancements in J up to several hundred percent due to contracted s-orbitals increasing Fermi . For example, in Sn- and Pb-containing compounds, relativistic corrections doubled certain one-bond J values compared to nonrelativistic results. Concurrently, integrated software packages such as Gaussian incorporated these methods into routine workflows, allowing users to compute full J tensors for biomolecules with hybrid DFT or coupled-cluster levels, often scaling to hundreds of atoms and supporting models for realistic predictions. In the and , models trained on quantum chemical data have enabled rapid prediction of J-couplings for large molecules, complementing traditional DFT and coupled-cluster methods, with applications in and as of 2025.

References

  1. [1]
    An Introduction to Biological NMR Spectroscopy - PMC
    Scalar coupling (or J-coupling) is a through-bond interaction between two nuclei (A and X) with a nonzero spin. It is an indirect interaction between the two ...
  2. [2]
    Single Quantum Coherence, J-coupling
    Therefore, J-coupling simply adds or subtracts πJ from the main energy term, leading to a splitting of J of the original resonance. The two spin coupling ...
  3. [3]
    NMR Spectroscopy - MSU chemistry
    Using this terminology, a vicinal coupling constant is 3J and a geminal constant is 2J. The following general rules summarize important requirements and ...
  4. [4]
    NMR Coupling Constants - Chemical Instrumentation Facility
    NMR Coupling Constants ; Aliphatic, H-C-H, geminal · J · -15 - -10 ; Aliphatic, H-C-C-H · J · 6-8 ; Aldehyde, H-C-CO-H · J · 2-3 ; Alkene, H-C-H, geminal · J · 0-3.Missing: vicinal | Show results with:vicinal
  5. [5]
    NMR Spectroscopy :: Hans Reich NMR Collection - Content
    Feb 14, 2020 · Interpretation of J-coupling constants a) Two bond (gem) coupling, MO theory of coupling b) Three bond (vicinal) coupling - Karplus curves c ...J-Coupling · 5.04 - Geminal J · Coupling constants · NMR Bibliography
  6. [6]
    [PDF] COUPLING CONSTANTS AND STRUCTURE: VICINAL COUPLINGS
    "Orientational Dependence of. Vicinal Proton-Proton NMR Coupling Constants: The ... • When Aν approaches J, the coupling is called “strong” and the spectra are ...
  7. [7]
    NMR J-coupling constants in cisplatin derivatives studied ... - PubMed
    Jun 6, 2011 · Key factors contributing to the magnitude of coupling constants are elucidated, with the most significant being the presence of solvent as well ...
  8. [8]
    NMR Spectroscopy - MSU chemistry
    ... coupling of the OH proton is apparent. The vicinal coupling (J = 7 Hz) of the methyl and methylene hydrogens is typical of ethyl groups. In DMSO-d6 solution ...Missing: definition | Show results with:definition
  9. [9]
    Simultaneous determination of multiple coupling networks by high ...
    Nov 15, 2021 · J coupling constitutes an important NMR parameter for molecular-level composition analysis and conformation elucidation. Dozens of J-based ...
  10. [10]
    Protein NMR. J(HN-HA) Coupling Constants - IMSERC
    J(HN-HA) coupling constants are a valuable NMR restraint to distinguish secondary protein structures, with values like 4.8 Hz for alpha-helix and 8.5 Hz for ...
  11. [11]
    Detection of J-Couplings at Zero Magnetic Field Using Atomic ...
    Aug 8, 2014 · APPLICATIONS OF TECHNOLOGY: Drug design; Catalyst evaluation and optimization; Research tool for pharmaceutical and biomedical industries ...
  12. [12]
    Electron Coupled Interactions between Nuclear Spins in Molecules
    In addition to the direct magnetic interaction between two nuclear spins in a molecule, the nuclei can have an effective mutual interaction.
  13. [13]
    Theory and calculation of nuclear spin–spin coupling constants
    hFC and hSD are the Fermi-contact and spin dipole interactions, respectively. ... JMN,uv are generally given as the sum of the five contributions. The ...
  14. [14]
    Review on DFT and ab initio Calculations of Scalar Coupling ... - MDPI
    K, is the reduced nuclear spin-spin coupling constant defined as Kjk = 4π2Jjk/hγjγk, [2] where h is the Planck constant and γi,γj are the magnetogyric ...
  15. [15]
    Heteronuclear J-coupling - Benchtop NMR Blog - Nanalysis
    Nov 25, 2020 · Typical 1J13C-1H coupling constants fall within the 115 – 270 Hz range and are dependent on bond strength, bond angle, and amount of s ...
  16. [16]
    19F Coupling Constants Table - Organofluorine / Alfa Chemistry
    Typical J-Value (Hz), Description. 19F-1H, CH3F, 45 to 50, Direct coupling; largest J due to high gyromagnetic ratio of 1H. 19F-13C, CF4, 240 to 320, Strong ...
  17. [17]
    Solvent effects on the 1 H 1 H, 1 H 19 F and 19 F 19 F coupling ...
    The 1H and 19F NMR spectra of fluorobenzene and 1,2-difluorobenzene were measured in several pure solvents and in binary solvent mixtures to study the solvent ...
  18. [18]
    Coupling Constant - an overview | ScienceDirect Topics
    For example, vicinal couplings for ethane derivatives are usually < 10 Hz, while in ethylene derivatives 3 J(cis) ≈ 10 and 3 J(trans) ≈ 17 Hz.
  19. [19]
    Illustrated Glossary of Organic Chemistry - n+1 rule
    n+1 rule: When splitting is first-order, the NMR signal for a nucleus having n neighbors is split into n+1 lines. TheH-NMR spectrum of 2-methoxybutane ...
  20. [20]
    (n+1) Rule - Chemistry LibreTexts
    Jan 29, 2023 · The (n+1) rule predicts peak multiplicity in NMR spectra. If a nucleus is coupled to n equivalent nuclei, the peak multiplicity is n+1.Missing: homonuclear | Show results with:homonuclear
  21. [21]
    NMR Spectroscopy :: 5-HMR-3 Spin-Spin Splitting: J-Coupling
    The scalar coupling J is a through-bond interaction, in which the spin of one nucleus perturbs (polarizes) the spins of the intervening electrons, and the ...
  22. [22]
    NMR Spectroscopy :: 5-HMR-9 Second Order Effects in Coupled ...
    When Δν /J < 1 then second order effects become very pronounced, often preventing detailed manual interpretation of multiplets, or giving incorrect coupling ...
  23. [23]
    4.7: NMR Spectroscopy - Chemistry LibreTexts
    Aug 28, 2022 · Different spin states interact through chemical bonds in a molecule to give rise to this coupling, which occurs when a nuclei being examined is ...
  24. [24]
    NMR Spectroscopy :: 5-HMR-15 The AA'BB' Pattern
    The AA'BB' pattern in NMR occurs when the 1:1 doublet splits into two lines, creating a mirror image relationship, and is often seen in X-CH2CH2-Y fragments.
  25. [25]
    NMR Spectroscopy :: 5-HMR-16 Virtual Coupling
    Virtual coupling refers to an NMR phenomenon in which apparently first-order multiplets contain false coupling information.
  26. [26]
    Virtual Coupling - University of Ottawa NMR Facility Blog
    Aug 11, 2011 · Virtual coupling occurs when the observed nucleus appears to be coupled to both of the other two nuclei even though it is only coupled to one of them.Missing: symmetric | Show results with:symmetric
  27. [27]
    DEVELOPMENTS IN THE KARPLUS EQUATION AS THEY RELATE ...
    Responding to some criticism of the accuracy of dihedral angles calculated from coupling constants, Karplus published a second paper on the subject in 1963, in ...
  28. [28]
    Application of NMR Spectroscopy in Isotope Effects Studies
    Similarly to isotope effects on chemical shifts the magnitude of isotope effect on coupling constant are proportional to the fractional changes in masses and ...
  29. [29]
    1H NMR Coupling Constants - Organic Chemistry Data
    4 J HH - 9 J HH : Long Range Acetylenic Couplings <sup>4 </sup><i>J</i><sub>HH</sub> - <sup>9 </sup><i>J</i><sub>HH</sub>
  30. [30]
    Signs of Proton Coupling Constants - ACS Publications
    ... Coupling Constant for Vicinal Hydrogens and its Sign Compared to that for Geminal Hydrogens. ... coupling constants from the temperature dependence of N.M.R. ...
  31. [31]
    NMR spin–spin coupling constants: bond angle dependence of the ...
    NMR spin–spin coupling constants: bond angle dependence of the sign and magnitude of the vicinal 3JHF coupling. R. V. Viesser, L. C. Ducati, J. Autschbach and ...
  32. [32]
    [PDF] 5.3 Spin-Spin Splitting: J-Coupling - Organic Chemistry Data
    The constants Jo and K are used to correct for substituent effects in more sophisticated uses of the Karplus equation, different Jo values are also used for ...Missing: γ_I γ_S
  33. [33]
    E.COSY and the Relative Signs of Coupling Constants
    Jun 4, 2010 · The E.COSY 1 (Exclusive COrrelation SpectroscopY) technique is one method which can be used to determine the relative signs of coupling constants.
  34. [34]
    Isotope effects on spin-spin coupling - ACS Publications
    Isotope Effects on Chemical Shifts and Coupling Constants. 1996https://doi ... Signs of 31 P 13 C coupling constants and isotope effects. Magnetic ...
  35. [35]
    The absolute sign of J coupling constants determined using the ...
    Aug 31, 2004 · A novel methodology using the order matrix calculation to determine the absolute sign of spin–spin couplings based on the structure of ...
  36. [36]
    None
    ### Summary of Vector Model and Product Operator Formalism for J-Coupling in NMR
  37. [37]
    [PDF] Plroduct Operator Formalism
    According to the vector model, spin-spin coupling causes a divergence of I-spin vectors at rates *J, with respect to a hypothetical vector precessing at the.
  38. [38]
    Accurate Measurement of Small J Couplings - ScienceDirect.com
    The measurement of J coupling constants for scalar-coupled spin systems is usually feasible through the conventional SQC high-resolution spectra. However, it is ...<|separator|>
  39. [39]
    [PDF] Measurement of Homonuclear Proton Couplings from Regular 2D ...
    Here, we describe a simple but effective method to directly obtain quantitative J-coupling information from cross peaks in regular phase-sensitive COSY spectra ...
  40. [40]
    2D J-Resolved Experiment - IMSERC
    The 2D J-resolved experiment allows to separate the coupling constant and chemical shift informations in separate dimensions of a 2D spectrum.
  41. [41]
    Correlation of connected transitions by two-dimensional NMR ...
    E.COSY restricts coherence transfer to connected transitions in the energy level diagram, simplifying cross-peak multiplets in 2D NMR spectra.Missing: original | Show results with:original
  42. [42]
    Measurement of J(H,H) and long‐range J(X,H) coupling constants in ...
    New methods for the measurement of homonuclear J(H,H) and heteronuclear long-range J(C,H) coupling constants are presented.
  43. [43]
    Spin Simulation for NMR - Mestrelab Resources
    Jun 6, 2015 · This module of Mnova is an efficient simulator for high resolution NMR spectra. Follow the menu 'Prediction/Spin Simulation' which will display the ...
  44. [44]
    [PDF] SpinWorks Documentation, Version 4.2.7 (2017/08/24)
    SpinWorks has two functions: The first is to provide easy basic off-line processing of 1D, 2D and 3D NMR data on personal computers.<|separator|>
  45. [45]
    Multiplet analysis by strong-coupling-artifact-suppression 2D J ...
    Jul 15, 2021 · Unfortunately, conventional 2D J-resolved experiments generally encounter challenges of insufficient spectral resolution and strong coupling ...
  46. [46]
    Temperature dependence of nuclear magnetic resonance coupling ...
    Temperature dependence of nuclear magnetic resonance coupling constants and chemical shifts of the vinyl halides and some vinyl esters. Click to copy article ...
  47. [47]
    Analyzing Coupling Constants - Oregon State University
    This is important to structure elucidation because there should be a single proton coupled to this one with J = 8 Hz, but two protons coupled with J = 2.1 Hz.
  48. [48]
    Heteronuclear spin decoupling in solid-state NMR under magic ...
    The best decoupling is achieved by multiple-pulse sequences at high RF fields and under fast MAS. Two such sequences, two-pulse phase-modulated decoupling (TPPM) ...
  49. [49]
    Selective Homodecoupled 1D-1H NMR Experiment for Unravelling ...
    Jul 14, 2022 · The selective homonuclear decoupling experiment is suitable for obtaining coupling constants and pattern recognition in routine structural ...Supporting Information · Author Information · Acknowledgments · References
  50. [50]
    A new decoupling method for accurate quantification of polyethylene ...
    This new decoupling method produced the cleanest 13 C NMR spectra for polyethylene copolymer composition and triad sequence distribution analyses.
  51. [51]
    Optimization of 1H decoupling eliminates sideband artifacts in 3D ...
    TROSY-based triple resonance experiments are essential for protein backbone assignment of large biomolecular systems by solution NMR spectroscopy.
  52. [52]
    Perspectives of adiabatic decoupling in liquids - ScienceDirect.com
    It is shown that adiabatic decoupling is extremely flexible and achieves very high figure of merit (decoupling bandwidth to RF amplitude ratio)
  53. [53]
    [PDF] NMR Techniques in Organic Chemistry: a quick guide [1] [2]
    In proton spectroscopy, such coupling will usually occur over two or three bonds (geminal or vicinal coupling respectively) and its presence provides direct ...
  54. [54]
    Accurate measurements of homonuclear HN-H.alpha. coupling ...
    Accurate measurements of homonuclear HN-H.alpha. coupling constants in polypeptides using heteronuclear 2D NMR experiments.Missing: study | Show results with:study
  55. [55]
    Vicinal Proton Coupling in Nuclear Magnetic Resonance
    Molecular structure refinement of a ß-Heptapeptide based on residual dipolar couplings: The challenge of extracting structural information from measured RDCs.
  56. [56]
    Experimental Calibration of a Karplus Relationship in Order to Study ...
    Experimental Calibration of a Karplus Relationship in Order to Study the Conformations of Peptides by Nuclear Magnetic Resonance. Click to copy article link ...
  57. [57]
    NMR Spectroscopy :: 5-HMR-5 Vicinal Proton-Proton Coupling 3JHH
    Vicinal proton-proton coupling (3JHH) is a useful H-H coupling that provides information about the spatial orientation between two protons. Its size is ...
  58. [58]
    Conformational dynamics detected by nuclear magnetic resonance ...
    Conformational dynamics detected by nuclear magnetic resonance NOE values and J coupling constants ... 1D and 2D Variable Temperature NMR Study. The ...
  59. [59]
    Nuclear Magnetic Resonance Spectroscopy. Variable-Temperature ...
    A variable-temperature, proton noise-decoupled, 13C nmr spectral study of the chair-chair interconversion in 1,1,3,3-tetramethylcyclohexane has demonstrated the ...
  60. [60]
    Residual dipolar couplings in NMR structure analysis - PubMed - NIH
    Residual dipolar couplings (RDCs) have recently emerged as a new tool in nuclear magnetic resonance (NMR) with which to study macromolecular structure and ...Missing: paper Bax
  61. [61]
    NMR Studies of Dynamic Biomolecular Conformational Ensembles
    Abstract. Multidimensional heteronuclear NMR approaches can provide nearly complete sequential signal assignments of isotopically enriched biomolecules.
  62. [62]
    The Dependence of a Nuclear Magnetic Resonance Frequency ...
    The dependence of a nuclear magnetic resonance frequency upon chemical compound. WG Proctor and FC Yu. Department of Physics, Stanford University, Stanford, ...Missing: first observation J- coupling HF molecules
  63. [63]
    NMR - Gaussian.com
    Jan 5, 2017 · In the second step, the other three terms in the spin-spin coupling are calculated with the unmodified basis set specified in the route section.