Skip To Main Content

You might also like

Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 15

Skip to Main content

 Journals & Books


Register Sign in
Electron Paramagnetic Resonance
Spectroscopy
Electron spin resonance reveals the presence of an S-alkylthiirane
radical cation in reactions of an alkyl vinyl sulfide with hydroxyl
radicals or the chlorine radical anion 〈81JCS(P2)1066〉.
From: Comprehensive Heterocyclic Chemistry, 1984
Related terms:
 Graphene
 Titanium Dioxide
 Photoluminescence
 Oxide
 Nanoparticles
 Impurity
 Protein
View all Topics
Download as PDF
Set alert
About this page
In-situ ESR for Studies of Paramagnetic Species on
Electrode Surfaces and Electron Spins Inside Electrode
Materials
Lin Zhuang, Juntao Lu, in In-situ Spectroscopic Studies of Adsorption
at the Electrode and Electrocatalysis, 2007
1 Introduction
Electron spin resonance (ESR), also known as electron paramagnetic
resonance (EPR), is a unique technique for detecting unpaired
electrons. Once an ESR signal is detected, this is very clear evidence
that the sample contains unpaired electrons, i.e.
uncompensated electron spins. This technique has been widely used
for the study of radicals. ESR can also be used to study other samples
containing unpaired electrons, such as transition metal ions and their
complexes, defects in solid materials, conduction electrons, etc. Since
the early stage of development, ESR has been closely related
to electrochemistry. There have been a number of reviews on the
application of ESR to electrochemistry [1–4]. Most previous studies
employing in-situ ESR in electrochemistry concerned radicals in
solutions. Reports of ESR studies of radicals on electrode surfaces
are limited due to a number of experimental difficulties. Electron spins
inside electrode materials are of interest and can also be studied by
in-situ ESR. In this chapter, studies of radicals on electrode surfaces
and of uncompensated electron spins inside electrodes are
summarized. We shall first briefly review the principles of ESR,
discuss the technical aspects of in-situ ESR, and then a number of
examples will be given.
View chapter Purchase book
Electron spin resonance studies of GaAs:Er,O
H. Ohta, ... Y. Fujiwara, in Rare Earth and Transition Metal Doping of
Semiconductor Materials, 2016
Abstract
Electron spin resonance (ESR) studies of GaAs codoped with Er and
O atoms (GaAs:Er,O) are presented. Detailed angular dependence
measurements of ESR revealed lower symmetry than C2v for Er-2O
centers. Moreover, the temperature dependence result and analyses
of line width suggested the existence of the exchange interaction for
Er-2O centers. From our ESR results and analyses for GaAs:Er,O with
and without charge carriers, we discuss the origin of A, B, and C ESR
centers, Er pair model, and the proposed model for the trap level.
View chapter Purchase book
ESR Insights into Macroradicals in UHMWPE
Muhammad Shah Jahan PhD, in UHMWPE Biomaterials Handbook
(Third Edition), 2016
Abstract
Electron spin resonance (ESR) is a technique that can directly detect
and quantify unpaired or odd electrons in atomic or molecular
systems. The materials that contain unpaired electrons are known
as paramagnetic materials since they exhibit a net magnetic moment
in an external magnetic field. For this reason, ESR is also known as
electron paramagnetic resonance (EPR). Although both names are
used in practice, a third name has been introduced to replace them. In
line with nuclear magnetic resonance (NMR), the new name is
electron magnetic resonance (EMR) since the magnetic resonance in
ESR/EPR results from the electron magnetic moment. In this chapter,
the term ESR is used.
View chapter Purchase book
Spectroscopic Characterization Protocols for
Interpenetrating Polymeric Networks
Jose James, ... Sabu Thomas, in Unsaturated Polyester Resins, 2019
10.3.4 Electron Spin Resonance Spectroscopic Analysis
ESR spectroscopy is another technique used for the characterization
of IPNs. ESR spectroscopy is widely used as a potential tool for
understanding the structure and heterogeneities in IPNs on a length
scale that is less than 5 nm. ESR line shapes and spectral simulation
data are used to arrive at some morphological conclusions along with
the help of SEM and TEM. ESR spectra of the spin probes in semi-
IPNs with different compositions, cross-link densities, and molecular
weights were measured in the temperature range of 173K–413K at
intervals of 10K or less. ESR spectra of probes measured as a
function of temperature for NR (polyisoprene) and for different semi-
IPNs of NR–PMMA systems [9] are shown in Fig. 10.6. The free
nitroxide radical, 4-hydroxy-2,2,6,6-tetramethylpiperidine-1-oxyl, was
used as a spin probe for the ESR measurements. In this study, the
main conclusions from the ESR investigations of PI–PMMA IPNs are
as follows.
Sign in to download full-size image
Figure 10.6.  ESR spectra of probes measured as a function of temperature in

(A) NR, (B) semi-IPN with 35 wt.% of PMMA, and (C) semi-IPN with 50/50

composition [9].

At lower temperatures, the ESR spectra approached the rigid limit


spectrum and as the temperature increased, the spectral lines got
narrower and the outer peaks shifted inward. In the range of 293K–
373K, complex spectra were obtained, which is an indication of the
presence of motionally distinct regions in semi-IPNs. The second
conclusion is that semi-IPN synthesis resulted in a dramatic change in
the motional behavior of both polymers and this was due to the
molecular level interpenetration between the two polymer chains. The
intimate mixing of the rigid PMMA chains with the highly mobile NR
networks strengthened the NR networks.
View chapter Purchase book
The influence of sterilization on octacalcium phosphate for
clinical applications
Kieran A. Murray, ... Cathriona O’Neill, in Octacalcium Phosphate
Biomaterials, 2020
4.2.5.1 Electron spin resonance
ESR analysis was performed using a continuous-wave X-band Bruker
EMXplus spectrometer with an ER 4131 VT temperature control
system. For absolute g-value determination a calibration using
diphenylpicrylhydrazyl at 0.1 mW (g=2.0036) was executed. Prior to
testing, all the nonirradiated and irradiated OCP samples were
prepared by placing them into Suprasil quartz tubes. The filled quartz
tubes were then loaded into the resonance cavity of the ESR
equipment. The analysis was performed at room temperature, with a
modulation amplitude of 0.1 mT and microwave power of 0.1, 1, and
10 mW. Testing was carried out in duplicate for each batch. The
assignment of the spectra to the corresponding radicals was
supported by computer simulations of the experimental spectra with
the “SimFonia” program (Bruker). The intensities of the spectra were
normalized according to signal I.
View chapter Purchase book
1,3,2-Dioxathiolane Oxides: Epoxide Equivalents and
Versatile Synthons
B.B. Lohray, Vidya Bhushan, in Advances in Heterocyclic Chemistry,
1997
D ESR Spectroscopy
Electron spin resonance (ESR) studies of various organic sulfites have
been conducted. The radical was generated by an aqueous solution
of titanium(III) chloride, hydrogen peroxide, and ethylene sulfite in the
cavity of an ESR spectrometer. The measurements were carried out
at pH < 2 or at pH 9 where nitromethane was used to generate the
nitro aci anion to trap the radical generated. Ethylene sulfite gave a
singlet that was as- signed on the basis of the g factor (2.0031)
to SO3−˙. The spectrum with a (2H) 0.173, a(2H 0.035 mT, g 2.003),
is attributed to splitting of EtOSO2·, which is likely to originate from
HOCH2CH2OSO2· (22) as shown in Scheme 1. The first step involved
the addition of the OH radical to the S = O bond to give 19, which
undergoes intramolecular H-atom abstraction by an alkoxy radical to
form 22 and 23 via 20.
Sign in to download full-size image
Scheme  1.

Alternatively, the alkoxy radical 22 may abstract a proton from an ‐α‐


SO3−˙ to form 23, which is trapped by CH2NO2−˙. The radical 23 is
suitable for fragmentation to form SO3−˙ [76JCS(P2)1040]. Similar
studies were carried out in the presence of 1BuO⋅ radical, which shows
the presence of sulfuranyloxy radical from ethylene sulfite
having g = 2.0044, a(1H) 0.24, a(2H) 0.0037 mT [77JCR(S)173].
When di-t-butyl peroxide is photolyzed in the presence of ethylene
sulfite in a solvent mixture of cyclopropane and ethylene oxide (1:1), a
spectrum of sulfuranyloxy radical is observed [a(1H) 2.38; a(2H)
0.37 G, g 2.0044 at 163 K]. The ESR spectrum of 21 shows a large
unique proton splitting, which is probably from one of the quasiapical
methyleneoxy protons. The presence of a five-membered ring may
increase the stability of 21 with respect to its acyclic
analog (77JMR509).
View chapter Purchase book
Nanocarbons: Preparation, assessments, and applications
in structural engineering, spintronics, gas sensing, EMI
shielding, and cloaking in X-band
Ashwini P. Alegaonkar, Prashant S. Alegaonkar, in Nanocarbon and
its Composites, 2019
7.4.2 Radical spin correlations: Electron spin resonance measurements and
analysis
ESR measurements were carried out to investigate spin transport at a
microwave frequency of ∼ 9.1 GHz (X-band) at room temperature.
The static magnetic field is swept over the range 300 − 370 mT
gradually at a spectrum-point time constant of 0.1 s and 6 kHz was the
amplitude of modulation frequency. The other parameters such as
field center, ESR line width sensitivity, resolution, and maximum
microwave output are 336 mT, ΔH = 0.05 mT, 7.0 × 109 spins/0.1 mT,
2.35 μT, and 985 − 1000 μW, respectively. For the GNCs and N-
GNCs, the first derivative of the paramagnetic absorption signal was
recorded. Among low and high concentration samples, the highest
was taken from a batch of N-GNCs. Fig. 7.16A shows the room-
temperature ESR spectra for GNCs. The first absorption derivative,
dY/dH, as a function of applied field shows a symmetric absorption
peak.

Sign in to download full-size image


Fig. 7.16.  Typical electron spin resonance spectra recorded at room

temperature. (left panel) Comparison of (a) GNC and (b) N-GNC line widths; (right

panel) enlarged feature for N-GNCs, indicated by an arrow.

For GNCs, symmetric and homogeneous broadening is observed. In


general, the ESR line shape gives information about magnetic
interactions in the measured sample. The line width, peak-to-peak
distance, and Hpp have an azimuth angular dependence, θ. So it has
the orientation of an applied field with respect to the orientation of the
GNC planes. For the bulk powder
specimen, Hpp(θ) = sin2(θ)H∥ + cos2(θ)H⊥. For GNCs, line width is found
to be 0.6756 mT. The measured Hpp consists of contributions
from H∥ as well as H⊥. From spin dynamics, the prospective important
parameter is the spin-spin relaxation time, Tss, correlated
by Hpp = (1/γeTss), where γe is the gyromagnetic ratio for electrons with a
value of 1.760859 × 1011/s T. For GNCs, the value Tss is 0.8406 ps.
The further magnitude of the g-factor is 1.99685, at which the
resonance has occurred under an applied microwave frequency and
characterizes the magnetic moment, μ, and γe, associated with
unpaired electrons in the material. The estimate of the g-factor is
found to be less than 2.0023, which is the g-factor for a typical
nondegenerated Pauli gas. This directs that the local magnetic
environment in GNCs is specifically different than the conventional
magnetic material. It is also clear that spin does not originate from
transition metal impurities, but only from carbon-inherited spin species
due to the small values of Hpp and a minor deviation in the g-factor.
Furthermore, the key factor in governing spintronic usability is the
spin-lattice relaxation time, Tsl, which describes the variation of the
nonthermal spin state around the lattice. The magnitude of Tsl is
computed as (1/Tsl) = (28.0 GHz)/THpp (where T is in Kelvin) and is
found to be 1.586 ns. The theoretical estimate for Tsl for materials
spintronic application is 1 − 100 ns; however 60 − 150 ps are the
experimental spin-transport measurements for graphene [99]. Using
the theory of spin relaxation and spin transport, the parameters were
calculated. Fig. 7.16B shows an enlarged ESR spectrum for N-GNCs.
A comparison of obtained parameters is given in Table 7.3.
Table 7.3 . Comparison of spin transport parameters for GNCs and N-GNCsa

Parameters GNCs N-GNCs


Resonance magnetic field (H ) r 327.163 mT 327.020 mT
g-Factor 1.99685 1.99856
Peak-to-peak (ΔH ) pp 0.6756 mT 0.4618 mT
Spin-spin relaxation time (T ) ss 0.8406 ps 1.2297 ps
Spin-lattice relaxation time (T ) sl 1.586 ns 2.3201 ns
Δg 5.45 × 10 − 3
4.24 × 10 − 3

Band width (Δ) 4.06 eV 3.79 eV


Parameters GNCs N-GNCs
Spin-orbit coupling constant (L ) i 22.12 meV 16.07 meV
Spin relaxation rate (Γ ) spin 7.81 × 10  eV
− 8
5.35 × 10  eV
− 8

Momentum relaxation rate (Γ) 2.63 meV 2.97 meV


Pseudo chemical potential (μ) 1.94 eV 1.79 eV
Density of states (ρ) 0.07469 stats/eV atom 0.06892 stats/eV atom
Pauli-spin susceptibility (χ at 300 K) 4.80 × 10 − 7
4.43 × 10 − 7

ESR line limit of detection 514.81 · L D 580.55 · L D

a
Measurements were performed at 300 K.

A comparative study of N-GNCs with GNCS shows that spin transport


is modified and reflected in measured relaxation parameters.
Magnitudes of Tss and Tsl are increased, respectively, to 1.2297 ps and
2.3201 ns. Thus, electron spin takes more time to regain its original
state. The possible reason is hindrance offered by a donor-loaded
nitrogen group. The extent of the SO coupling constant, Li, for N-
GNCs is reduced to 16.0 from 22.12 meV as that of GNCs. Basically,
in graphene, Li originates through intrinsic, Bychkov-Rashba (BR,
related to structural symmetry break), and ripples (related to inevitable
wrinkles/folding edges). The theoretical estimation for the intrinsic SO
coupling constant is between 1 μeV and 0.2 meV, and for BR,
10 − 36 μeV and V− 1 nm− 1. For the intrinsic component,
when μ˜ ≫ Γ, Γ(intrinsic) ≈ Li2/Γ, which is an Elliot-Yafet-like result [93],
where μ˜ is the pseudochemical potential, Γ is the momentum
relaxation rate, and Γ(intrinsic) is the intrinsic momentum relaxation
rate. Further, the ripple relaxation contribution becomes dominant only
when Γ ≫ μ˜. The computed values of Γ and μ˜ for our systems
indicate that intrinsic and BR could be the operative channels. But one
can neglect the intrinsic term because disorder is present in GNCs.
Once a vacancy is generated, the carbon atom that is the nearest
neighbor to a vacancy is displaced from its equilibrium position.
Using the theory of spin relaxation from experimentally obtained ESR
lines, the spin transport parameters were computed for maximum
concentration (∼ 0.5 mg/mL for GNCs and N-GNCs). As an effect, the
system with a vacancy undergoes a Jahn-Teller distortion and breaks
the honeycomb symmetry. The quantity of sites with damaged spatial
inversion symmetry is high in GNCs compared to pure graphene. In
the case of pure graphene, when Δ ≫ Li, where, Δ, is band gap, the
SO interaction couples π and σ band.
Hence, in GNCs the principal component in Li is BR with π-σ band
distortion. The decrease in the extent of Li may be attributed to a
reduction in mid gap between the π and σ bands, which is due to
alterations at the broken symmetry sites. This perturbation is due to
the nitrogen-loaded donor moieties. This is because of their strong
tendency to donate wandering electrons via a charge-transfer
interaction with the sp2 network. Further spin-relaxation time is
calculated using Tss = Ncol τ, where Γ ∼ 1/τ, τ is the momentum
scattering rate and Ncol is the spin-flip scattering probability (meV ps) of
the spin density wave propagating in crystal. [100]. Using this relation
of GNCs, the value of Ncol is estimated to be 2.22 meV ps and for N-
GNCs, the collisional probability is boosted to 3.65 meV ps. A similar
consistency is found with variations of density of states ρ(μ˜,Γspin) in
both systems, where Γspin is the spin-relaxation rate. It is remarkable
that enormous deviations have been observed between theoretical
and experimentally derived values of spin transport parameters for
GNC systems. The Pauli spin susceptibility, χspin, obtained via the ESR
technique (given in Table 7.3) as well as from VSM measurements are
well matched.
View chapter Purchase book
Tools and techniques for characterization and evaluation
of nanosensors
Zamaswazi P. Tshabalala, ... David E. Motaung, in Nanosensors for
Smart Cities, 2020
5.9.9 Electron paramagnetic resonance spectroscopy
Electron paramagnetic resonance (EPR) spectroscopy is valuable
instrumentation for the gas-sensing analysis [118–121]. It is used to
detect the superoxide radical O2−, which forms owing to the transfer
of electrons, trapped in VO, to the adsorbed oxygen molecules in the
SnO2 surface and atomic adsorbed oxygen (O−), also as paramagnetic
oxygen vacancies (VO), which are alleged to be significant species in
the gas response of SMOx. Nonetheless, the paramagnetic species
are not observable during operando. [122,123] Previous studies have
carried out EPR ex situ analyses on several SMOx to detect
superoxide radical O2−, VO, and O− and link these properties with the
gas-sensing performances [52,118–121,124]. However, with regard to
catalysis, differentiation between the synergetic and direct
mechanisms, which may be appropriate in redox reactions and their
comparative contributions, is executed through in situ EPR
spectroscopic investigations that facilitate the classification of the
valence states of the metal ions/atoms.
View chapter Purchase book
ANALYTICAL METHODS IN UV DEGRADATION AND
STABILIZATION STUDIES
In Handbook of UV Degradation and Stabilization (Third Edition), 2020
11.12.1 ESR
Electron spin resonance addresses the very base of the changes
since most of them involve radical mechanisms and radicals can be
directly observed using ESR. ESR permits identification of radicals
and monitoring of their concentration in the sample (especially when
the specimen mounted in an instrument is exposed to a controlled UV
radiation). Gerlock et al.33–35 developed an original approach to testing
using ESR. Knowing that a hindered amine light stabilizer, HAS, can
produce a durable nitroxide radical, they doped polymer coatings with
HAS or hindered amine nitroxide, HAN, and measured nitroxide
concentration. Nitroxide scavenged the radicals generated by the
photolysis of the coating, converting short-lived species to stable
radicals. This approach made possible the quantitative evaluation of
radical formation rate.
The effect of diazo pigment on stabilization of polycarbonate in the
presence of HAS has been elucidated with the help of ESR. 36 The
likely mechanism of stabilization by the diazo pigment is the
absorption of UV radiation and energy conversion to non-aggressive
energy forms for PC, such as heat.36
It was established that the concentration of nitroxide in a sample is
usually low (less than 10% of added HAS concentration). 37 The
nitroxide concentration in the matrix is an integral sum of >NO . arising
in the primary step from the parent HAS and >NO formed in the
regeneration step from >NOP.37
Spatial distribution of HAS-developed nitroxides can be determined
experimentally using an ESR imaging technique (ESRI).37 The results
of studies by imaging technique helped to show experimentally the
superposition of the thermal stress on the “dark” side of the sample as
well as to validate experimentally the effect of oxygen available on
both surfaces which, combined with the residual radiation penetrating
through the sample depth to the “dark” (nonirradiated) surface, caused
degradation on the back side of the sample. 37
Spatial and temporal effects of the aging process were studied by
electron spin resonance and ESR imaging (ESRI) of HAS-derived
nitroxide radicals, by comparison of thermal transitions in original and
aged samples.38 Nondestructive (“virtual”) slicing of the 2D spectral-
spatial ESR images resulted in a series of ESR spectra that reflected
the variation of the F/S intensity ratio with sample depth. 38 HAS-NO
was detected on the nonirradiated side due to the small, but not
negligible, intensity of UV radiation transmitted through the plaque; on
this side the nitroxide concentration remained high because it was not
consumed in stabilization processes, as the local rate of degradation
was low.38
ESR was instrumental in realizing the well-designed concept of
sunscreen.39 Sunscreen protects the skin from the generation of free
radicals, whereas antioxidants remove the generated radicals. 39 A UV
filter (uncoated physical or chemical) that catalyzes the generation of
free radicals under UV irradiation should not be used. 39
View chapter Purchase book
UV DEGRADATION & STABILIZATION OF INDUSTRIAL
PRODUCTS
In Handbook of UV Degradation and Stabilization (Third Edition), 2020
8.22.1 IMPORTANT CHANGES AND DEGRADATION MECHANISMS
Electron paramagnetic resonance spectroscopy was used to monitor
formation of free radicals in human hair upon UV irradiation. 1 The EPR
spectra of brown hair were dominated by melanin signal, and those of
white hair were keratin-derived.1 The decay of UV induced keratin
radicals was enhanced by increased ambient humidity because
humidity had swollen the hair increasing molecular mobility fostering
radical reactions.1 This interpretation is consistent with the increased
UV-triggered protein damage in hair at high humidity as demonstrated
by the protein loss (Figure 8.8).1

Sign in to download full-size image


Figure 8.8.  Summary of radiation-induced autoxidation processes leading to

polypeptide cleavage, disulfide cleavage, and side chain degradation in proteins. R

= aminoacid side chain or a-carbon of amino acid in polypeptide.

[Adapted, by permission, from Groves, P; Marsh, J M; Sun, Y; Chaudhary, T;

Chechik, V, Free Radical Biol. Med., 121, 20-5, 2018.]Copyright © 2018

Ultraviolet radiation is harmful to human epidermal keratinocytes of


the epidermal skin layer, as well as to hair-follicle-associated
keratinocytes.2 Following UV irradiation, cells in minimal growth
medium showed a 72% reduction in metabolic activity, while l-cystine
and thiamine-treated cells showed only a 12-18% reduction. 2 The
prevention of the UV-induced reduction in metabolic activity was not
simply due to filtering UVR by the l-cystine- and thiamine-mediated
reduction, but this effect persisted even when l-cystine and thiamine
were washed out during UV irradiation.2
The human hair fibers were progressively damaged by exposure to
both UV and ionized gamma radiation. 3 A decrease in the
fluorescence intensity resulted from depletion of the aminoacid
tryptophan and significant reduction in a number of free SH-groups in
hair proteins.3 Hair damage was dose-dependent for exposures
between 0 and 10.0 Gy and 0-20 J/cm2 of UV radiation.3
View chapter Purchase book

Recommended publications:

 Journal of Alloys and Compounds


Journal

 Journal of Non-Crystalline Solids


Journal

 Materials Science and Engineering: C


Journal
 Materials Research Bulletin
Journal
Browse Journals & Books

 About ScienceDirect
 Remote access
 Shopping cart
 Advertise
 Contact and support
 Terms and conditions
 Privacy policy
We use cookies to help provide and enhance our service and tailor
content and ads. By continuing you agree to the use of cookies.
Copyright © 2021 Elsevier B.V. or its licensors or
contributors. ScienceDirect ® is a registered trademark of Elsevier
B.V.

You might also like