You are on page 1of 25

A Project

on

“Role of Antiferromagnetism in High Tc


Superconductivity”
Submitted To

Post Graduate Department of Physics


Government (Autonomous) College, Angul

Submitted By

Twinkle Chanda

M. Sc. Physics, Exam Roll No: 21PGPHY001

Government (Autonomous) College, Angul, Odisha - 759122

Guided By
Dr. Basanta Kumar Sahoo
Asst. Professor in Physics (Stage – III)
Government (Autonomous) College, Angul, Odisha - 759122
GOVT. (AUTONOMOUS) COLLEGE, ANGUL

Certificate
This is to certify that the project work entitled "Role of Antiferromagnetism in

High Tc Superconductivity" submitted to the Post Graduate Department of

Physics, Government (Autonomous) College, Angul in partial fulfilment of

requirements for the Course Master of Science in Physics, is a work done by Miss

Twinkle Chanda under my supervision and guidance and that this work has not

been submitted elsewhere for the award of any degree.

Place: (Signature)
Date: Dr. Basanta Kumar Sahoo
Asst. Professor in Physics (Stage – III)
Govt. (Autonomous) College, Angul
Acknowledgement
I would like to convey my heartfelt gratitude to all the people who have helped and
inspired me during my project Work. This project would not have been possible
without the support from all.

First of all, I would like to thank my project guide and HOD Dr. Basanta Kumar
Sahoo, for his endless guidance and instruction during the project.

I further thank all other staff member of PG Department of Physics, Govt.


(Autonomous) College, Angul for all their help and support in completing the
project successfully.

My deepest gratitude goes to my family for their love and support all the way till
now. I also thank my friends here for their constant encouragement and the joyous
time we spent together. All the references I used during the project work including
the definitions, texts, formulae etc. are mentioned separately in the reference
section. I personally thank all the authors and the contributors.

Twinkle Chanda

ii
Government (Autonomous) College, Angul
Post Graduate Department of Physics

Declaration

I, Twinkle Chanda, do hereby declare that my project “Role of


Antiferromagnetism in High Tc Superconductivity” is my original work. It is
not submitted by any other institution or published any time before the purpose
what so ever. This declaration has compiled for fulfilling requirement for the award
of M.Sc. Physics of Government (Autonomous) College, Angul.

Date: Twinkle Chanda

Place: Angul 2nd Year M.Sc. Physics

Exam Roll No: 21PGPHY001

iii
Contents

 Magnetism

 Diamagnetism

 Paramagnetism

 Ferromagnetism

o Magnetic domains

 Antiferromagnetism

 Key Difference – Ferromagnetism vs Antiferromagnetism

 Antiferromagnetic Materials

 Antiferromagnetic energy gap in chromium

 Methods of Study Energy Gap

 Antiferromagnetism in Superconductors

 Hybridization

 External Magnetic Field

 Doping

 Difference between Spin Density Wave and Antiferromagnetism

 Antiferromagnetism in High Temperature Superconductivity

 Cuprate High-Temperature Superconductor

o Structure
iv
o Superconducting Mechanism

o Re-cuprating Models

 Iron based superconductor

 Heavy fermion superconductor

 Organic Superconductor

o One-dimensional Fabre and Bechgaard salts:

o Two-dimensional (BEDT-TTF) 2X

 Doped fullerenes

 Comparison Between Cuprates and Iron Pnictide

 Conclusion

 Reference

v
Abstract
Magnetism is the result of attraction, when two objects come together, or repulsion when two objects move
apart. A magnet is an object that has properties of magnetism. For example, a magnet might attract another
object. A magnetic field is an invisible area around a magnet where magnetism occurs. These are three types.
Such as diamagnetism, paramagnetism, ferromagnetism.

Equal magnitude of magnetic moment associated with unpaired electrons are aligned in opposite
directions, the net magnetic moment is zero is called antiferromagnetism. Example: MnO, Mn2O3 and
MnO2.Ferromagnetism is observed in transition metals and some of their compounds. In
antiferromagnetism, the magnetic moments point in the opposite direction. Due to this, the magnetic
susceptibility of a substance decreases to some extent. Antiferromagnetism is observed in the salts of ions like
Mn+2, Fe+3 and Gd+3. Antiferromagnetic materials occur commonly among transition metal compounds,
especially oxides. Examples include hematite, metals such as chromium, alloys such as iron manganese
(FeMn), and oxides such as nickel oxide (NiO). There are also numerous examples among high nuclearity
metal clusters. Antiferromagnetic materials improve the way information is written and read electrically in
devices. They are microscopic magnets with opposite orientations. Computers nowadays use Silicon
components, but they fail to be as efficient as antiferromagnetic materials.

High-transition temperature copper oxide superconductors have puzzled physicists since these materials
were discovered 10 years ago. Nagaosa's Perspective discusses a new theory described by Zhang (1089) in the
same issue in which superconductivity is related to another important property, antiferromagnetism, by an
elegant symmetry principle. Antiferromagnetism exists near superconductivity in the electronic phase diagram
of a variety of seemingly unrelated unconventional superconductors. Unconventional superconductors are
those which are not described by BCS theory. Antiferromagnetism (abbreviated AF) is where the spin on
adjacent magnetic sites points in opposite directions. Below are some classes of superconductors where
antiferromagnetism seems to have something to do with superconductivity. Note that each class comprises
many different materials, so in actuality, there are hundreds of compounds which share this
property.Cuprate High-temperature superconductors, Iron-based (pnictide) sort-of-high temperature
superconductors and Organic superconductor.

vi
Role of Antiferromagnetism in High Tc Superconductivity
-Twinkle Chanda

1. Magnetism:

Magnetism is the class of physics attributes that are mediated by a magnetic field, which refers to the
capacity to induce attractive and repulsive phenomena in other entities. Magnetism is one aspect of the
combined phenomena of electromagnetism. The magnetic state of a material depends on temperature, pressure,
and the applied magnetic field. A material may exhibit more than one form of magnetism as these variables
change. Magnetism, phenomenon associated with magnetic fields, which arise from the motion of electric
charges. This motion can take many forms. It can be an electric current in a conductor or charged particles
moving through space, or it can be the motion of an electron in an atomic orbital. Magnetism is also associated
with elementary particles, such as the electron, that have a property called spin.

The most familiar effects occur in ferromagnetic materials, which are strongly attracted by magnetic
fields and can be magnetized to become permanent magnets, producing magnetic fields themselves.
Demagnetizing a magnet is also possible. Only a few substances are ferromagnetic; the most common ones
are iron, cobalt, and nickel and their alloys. The rare-earth metals neodymium and samarium are less common
examples. The prefix ferro- refers to iron because permanent magnetism was first observed in lodestone, a
form of natural iron ore called magnetite, Fe3O4.

All substances exhibit some type of magnetism. Magnetic materials are classified according to their bulk
susceptibility. Ferromagnetism is responsible for most of the effects of magnetism encountered in everyday
life, but there are actually several types of magnetism. Paramagnetic substances, such
as aluminium and oxygen, are weakly attracted to an applied magnetic field; diamagnetic substances, such
as copper and carbon, are weakly repelled; while antiferromagnetic materials, such as chromium and spin
glasses, have a more complex relationship with a magnetic field. The force of a magnet on paramagnetic,
diamagnetic, and antiferromagnetic materials is usually too weak to be felt and can be detected only by
laboratory instruments, so in everyday life, these substances are often described as non-magnetic.

The strength of a magnetic field almost always decreases with distance, though the exact mathematical
relationship between strength and distance varies. Different configurations of magnetic moments and electric
currents can result in complicated magnetic fields.

Magnetism, at its root, arises from two sources:

1 Electric current.

2 Spin magnetic moments of elementary particles.

The magnetic properties of materials are mainly due to the magnetic moments of their atoms orbiting electrons.
1
These are three types. Such as;

 Diamagnetism

 Paramagnetism

 Ferromagnetism

1.1 Diamagnetism:

Diamagnetism appears in all materials and is the tendency of a material oppose an applied magnetic
field, and therefore, to be repelled by a magnetic field. However, in a material with paramagnet. Thus despite
properties, the paramagnetic behaviour dominates despite its universal occurrence, diamagnetic field’s
behaviour is observed only in a purely diamagnetic material. In a diamagnetic material, there are no unpaired
electrons, so the intrinsic electron magnetic moments cannot produce any bulk effect. In these cases, the
magnetization arises from the electron orbital motions, which can be understood classically as follows:

When a material is put in a magnetic field, the electrons circling the nucleus will experience, in addition
to their coulomb attraction to the nucleus, a Lorentz force from the magnetic field. Depending on which
direction the electron is orbiting, this force may increase the centripetal force on the electrons, pulling them in
towards the nucleus, or it may decrease the force, pulling them away from the nucleus. This effect
systematically increases the orbital magnetic moments that were aligned moments that were aligned opposite
the field and decreases the ones aligned parallel to the field. This results in a small bulk magnetic moment,
with an opposite direction to the applied field.

All materials undergo this orbital response.


However, in paramagnetic and ferromagnetic
substances, the diamagnetic effect is overwhelmed by
the much stronger effects caused by the unpaired
electrons.

Figure 1: Hierarchy of magnetism

1.2 Para magnetism:

In a paramagnetic material there are unpaired


electrons; i.e., atomic or molecular orbitals with
exactly one electron in them. While paired electrons
are required by the Pauli Exclusion Principle to have their intrinsic ('spin') magnetic moments pointing in
opposite directions, causing their magnetic fields to cancel out, an unpaired electron is free to align its magnetic
moment in any direction. When an external magnetic field is applied, these magnetic moments will tend to
align themselves in the same direction as the applied field, thus reinforcing it.

2
1.3 Ferromagnetism:

Ferromagnetism is a property of certain materials (such as iron) that results in a significant, observable
magnetic permeability, and in many cases, a significant magnetic coercivity, allowing the material to form a
permanent magnet. Ferromagnetic materials are familiar metals that are noticeably attracted to a magnet, a
consequence of their substantial magnetic permeability. Ferromagnetism is a property of certain materials
(such as iron) that results in a significant, observable magnetic permeability, and in many cases, a significant
magnetic coercivity, allowing the material to form a permanent magnet. Ferromagnetic materials are familiar
metals that are noticeably attracted to a magnet, a consequence of their substantial magnetic permeability.
Every ferromagnetic substance has its own individual temperature, called the Curie temperature, or Curie
point, above which it loses its ferromagnetic properties. This is because the thermal tendency to disorder
overwhelms the energy-lowering due to ferromagnetic order.

Ferromagnetism only occurs in a few substances; common ones are iron, nickel, cobalt, their alloys, and
some alloys of rare-earth metals.

1.3.1 Magnetic domains:

The magnetic moments of atoms in a ferromagnetic material cause them


to behave something like tiny permanent magnets. They stick together and
align themselves into small regions of more or less uniform alignment called
magnetic domains or Weiss domains. Magnetic domains can be observed with a
magnetic force microscope to reveal magnetic domain boundaries that
resemble white lines in the sketch. There are many scientific experiments that
can physically show magnetic fields.

Figure 2: Magnetic domains boundaries (white lines) of ferromagnetic material (black rectangle)

When a domain contains too many molecules, it becomes unstable and divides
into two domains aligned in opposite directions so that they stick together more
stably. When magnetized strongly enough that the prevailing domain overruns
all others to result in only one single domain, the material is magnetically
saturated. When a magnetized ferromagnetic material is heated to the Curie point
temperature, the molecules are agitated to the point that the magnetic domains
lose the organization, and the magnetic properties they cause cease.

Figure 3: Effect of a magnet on the domain.

1.4 Antiferromagnetism:
Figure 4: Antiferromagnetic ordering

3
In an antiferromagnet, unlike a ferromagnet, there is a tendency for the intrinsic magnetic moments of
neighboring valence electrons to point in opposite directions. When all atoms are arranged in a substance so
that each neighbour is anti-parallel, the substance is antiferromagnetic.

Antiferromagnets have a zero net magnetic moment, meaning that no field is produced by them.
Antiferromagnets are less common compared to the other types of behavior and are mostly observed at low
temperatures. In varying temperatures, antiferromagnets can be seen to exhibit diamagnetic and ferromagnetic
properties. Antiferromagnetic structures were first shown through neutron diffraction of transition metal
oxides such as nickel, iron, and manganese oxides. The experiments, performed by Clifford Shull, gave the
first results showing that magnetic dipoles could be oriented in an antiferromagnetic structure.

Antiferromagnetic materials occur commonly among transition metal compounds, especially oxides.
Examples include hematite, metals such as chromium, alloys such as iron manganese (FeMn), and oxides
such as nickel oxide (NiO). There are also numerous examples among high nuclearity metal clusters.
Organic molecules can also exhibit antiferromagnetic coupling under rare circumstances, as seen in radicals
such as 5-dehydro-m-xylylene.

2. Key Difference – Ferromagnetism vs Antiferromagnetism:

Ferromagnetism and antiferromagnetism are two of the five classifications of magnetic properties. The
other three are diamagnetism, paramagnetism, and ferrimagnetism. The key difference between
ferromagnetism and antiferromagnetism is that ferromagnetism can be found in materials having their
magnetic domains aligned into the same direction whereas antiferromagnetism can be found in materials
having their magnetic domains aligned in opposite directions.

A magnetic domain or an atomic moment is a region in which the magnetic fields of atoms are grouped
together and aligned. Ferromagnetic materials are attracted to an external magnetic field and have a net
magnetic moment. But antiferromagnetic materials have a zero net magnetic moment.

3. Antiferromagnetic Materials:
When any material is placed in an applied magnetic field, it experiences the field and thus induces a
magnetic property in it. These magnetic properties are induced on the macroscopic level based on the
interaction between the individual dipole moments of an atom with the applied magnetic fields. The one that
shows the commonly known property that the material is attracted to the magnetic field wholly, is known as
the ferromagnet. In this, the magnetic moments of each atom are all aligned in the same direction.

In the case of antiferromagnets, the magnetic moments of each atom are arranged in such a way that
every second moment is in the opposite direction to the first. In simple words, in a pair of atoms, the moments
are arranged in such a way that they cancel out each other. That is if one is upwards the other will be
downwards.

4
Thus, in anti-ferromagnets, the net magnetic moment of the
substance is zero as the alternate orientations of moment cancel out
each other. This indicates that they do not produce any magnetic field
of their own. The anti-ferromagnetic nature can be destroyed by
heating at higher temperatures. Examples are Manganese oxide,
Chromium oxide, Ferrous oxide etc.

Figure 5: Band Gap Characters and Ferromagnetic/Antiferromagnetic Coupling in Group-IV Monolayers


Tuned by Chemical Species and Hydrogen Adsorption
Antiferromagnetic Neel
Configurations.
materials temperature
Table 1: Here are some examples of materials with NEEL
temperature Ferrous oxide 198k

3.1. Antiferromagnetic energy gap in chromium: Nickel fluoride 73k


A temperature-dependent dip has been observed in the
Chromium 311k
infrared reflectivity of chromium, which is attributed to
excitation of an electron across the antiferromagnetic energy Manganese oxide 116k
gap. At 80°K, the absorption maximum occurs at photon
Manganese 67k
energy 5.1kTN, which is larger than the value predicted by the
fluoride
simplest BCS-type model, 3.5kTN. Both the position and
shape of the adsorption can be explained if the effects of electron-phonon scattering are incorporated in the
model.

4. Methods of Study Energy Gap:


The interatomic distance can considerably influence the exchange interaction between two neighboring
magnetic atoms and then decides whether ferromagnetism or antiferromagnetism forms in magnetic materials.
We test the effect of the interatomic distance to the magnetic coupling of semihydrogenated group-IV
monolayers with chair configuration. For all structures with

chair configuration in this work, the interatomic distance between two neighboring magnetic atoms is
equal to the lattice constant. Total energies of FM and AFM coupling in each structure are recalculated with
elongated and shorten lattice constants within a range of ±5 %. In this range, SiC-H and GeC-H undergo FM-
AFM transitions. The difference of total energies between AFM and FM states is depicted in Fig. 8, where a
positive value means a ferromagnetic system and a negative value means an antiferromagnetic system. The
difference of total energies increases with elongated interatomic distance and decreases with shorten
interatomic distance. FM-AFM transition for SiC-H and GeC-H occurs at 3.180 and 3.223 Å, respectively.
This transition can prove that the interatomic distance between two neighboring magnetic atoms is a key factor
that influences whether ferromagnetic or antiferromagnetic coupling forms. Kaloni et al. found that decorating
5
silicene or germanene with 3d or 5d transition metal atoms can induce magnetism, and the formation of FM
or AFM coupling is controlled by the species and position of transition metal atoms.

5. Antiferromagnetism in Superconductors:

The reason for the weak mutual effects of superconductivity and antiferromagnetism is that there is no
average 412 Sov. Phys. Usp. 29 (5), May 1986 0038-5670/86/050412-14$01.80 © 1986 American Institute of
Physics 412 magnetization in an antiferromagnetic substance (in contrast with ferromagnetism), and the
magnetic moment changes direction over scale lengths small in comparison with the superconducting
correlation length. Superconductivity has a very weak effect on this order, since Cooper pairing causes almost
no change in the electron spin susceptibility at the wave vectors G ~a characteristic of antiferromagnetic order.
Here the exchange interaction remains the same [within a relative error (a/£0) ~TC/^F as in normal
antiferromagnetic materials. The short-wave part of the electromagnetic interaction also remains essentially
the same upon the appearance of superconductivity within (aV/i L£O)> where/I L is the London screening
depth. The reason is that effective screening of the field in a superconductor is possible only over distances of
the order of or greater than /I L. The effect of antiferromagnetism on superconductivity turns out to be stronger.
Two basic mechanisms for this effect can be identified: a) the spin splitting of electronic levels by the exchange
field set up by localized moments and the resulting appearance of a gap on a small part of the Fermi surface.
Such a gap reduces the total density of electron states.I4|1S Another important point is that the exchange field
makes the superconductivity gapless (in a pure compound); this effect is usually basically responsible for the
suppression of superconductivity at low temperatures, T< TN < Tc. b) The magnetic scattering of electrons16
by spin fluctuations above the Neel point and by spin waves below the Neel point. The most important
characterstics are hybridization, external magnetic field and doping.

5.1 Hybridization:

Magnons in magnetically ordered low-dimensional materials potentially enable novel nano-spintronics.


The lifetime of magnons often depends on the strength of spin-lattice coupling and scattering with phonons.
Here, we found coherent magnon phonon hybridization in the class of transition metal phosphorous
trichalcogenides (MPX3, M = Fe, Mn, Ni and X = S, Se). From bulk to ultrathin layer, the antiferromagnetic
magnons are identified by Zeeman splitting with applied magnetic field. A model is developed to
quantitatively explain the nature of the emergent quasiparticles, as well as their center frequencies, scattering
intensities, and polarization selection. This strong coupling may result in finite Berry curvature in the
hybridization bands and possibly non-trivial topology.

Magnetic excitations in van der Waals (vdW) materials, especially in the two-dimensional (2D) limit,
are an exciting research topic from both the fundamental and applied perspectives. Using temperature-
dependent, magneto-Raman spectroscopy, we identify the hybridization of two-magnon excitations with two
separate Eg phonons in MnPSe3, a magnetic vdW material that could potentially host 2D antiferromagnetism.

6
Results from first principles calculations of the phonon and magnon spectra further support our identification.
The Raman spectra's rich temperature dependence through the magnetic transition displays an avoided-
crossing behavior in the phonons' frequency and a concurrent decrease in their lifetimes. We construct a model
based on the interaction between a discrete level and a continuum that reproduces these observations. The
strong magnon-phonon hybridization reported here highlights the need to understand its effects on spin
transport experiments in magnetic vdW materials.

5.1.1 Raman phonons of MnPSe3:

Figure 6: Raman phonons of MnPSe3.

As the temperature is lowered through TNéel, several


notable features appear in the frequency region between 75
and 140 cm−1: (i) The Eg1 and Eg2 phonons undergo
substantial changes in intensity and frequency, and (ii) a
new scattering intensity peak appears to split off from
Eg2 and continues to increase in frequency to approximately 130 cm−1 at 10 K. The new mode, with
Eg symmetry, cannot be attributed to a phonon according to the DFT calculations and the lack of structural
transition across TNéel. This mode is asymmetric, with a cutoff at high frequency, a peak intensity around 130
cm−1, and a low-intensity tail that merges with the Eg2 phonon. A previous study interpreted the origin of the
130 cm−1 peak as one-magnon scattering based on its strong frequency shift as a function of temperature. A
one-magnon scattering mode in an AFM is expected to split in an applied magnetic field due to its net magnetic
moment, ∆S = ±1. This behavior was observed in 3D AFMs such as MnF2 and FeF2 and more recently in vdW
AFMs. However, when we measure the low-temperature (T = 2 K) Raman spectra as a function of static
applied magnetic field, both parallel and perpendicular to the honeycomb plane, we find no change in the 130
cm−1 peak. The lack of magnetic field dependence and the similarities with the two-magnon scattering in
MnF2/FeF2 strongly suggests that the peak at 130 cm−1 is not due to one-
magnon scattering but is, instead, due to two-magnon scattering.

5.1.2 Temperature-dependent Raman spectra of MnPSe3:

The Raman spectra for crossed (VH) polarization collected through


the Néel transition at 74 K. The spectra are normalized to the peak intensity
of Eg2 and offset for clarity. Below TNéel, a new feature appears between 115
and 130 cm−1, while the E g1 and Eg2 phonons undergo changes in both
intensity and frequency.

Figure 7: Temperature-dependent Raman spectra of MnPSe3.

7
5.2 External Magnetic Field:

We describe the dynamics of an antiferromagnetic nano-oscillator in an external magnetic field of any


given time distribution. The oscillator is powered by a spin current originating from spin-orbit effects in a
neighboring heavy metal layer and is capable of emitting a THz signal in the presence of an additional easy-
plane anisotropy. We derive an analytical formula describing the interaction between such a system and an
external field, which can affect the output signal character. Interactions with magnetic pulses of different
shapes, with a sinusoidal magnetic field and with a sequence of rapidly changing magnetic fields are
discussed. We also perform numerical simulations based on the Landau-Lifshitz-Gilbert equation with spin-
transfer torque effects to verify the obtained results and find a very good quantitative agreement between
analytical and numerical predictions.

Figure 8: Schematic picture of the antiferromagnetic nano-oscillator


interacting with external magnetic field. Bright red arrows represent
the antiferromagnetic magnetization vectors →m1 and →m2, blue
arrows represent the influence of hard axis anisotropy
field →Hh and of the spin torque →τ (both constant in time). Green
arrow represents the external magnetic field →Hp (varying in time)
which is either parallel or antiparallel to the anisotropy field.

Figure 9: Magnetic field pulses of trapezoidal (a) and Gaussian-like (b)


shape and the respective antiferromagnetic oscillator responses to their
presence (c) and (d). For the trapezoidal shape, we used
slopes β1=5 mT/ps and β2=−2 mT/ps. For the Gaussian-like shape, we
used parameters tμ=37 ps and tσ=3 ps.

5.3 Doping:

In order to understand the role of Al on the magnetic properties of YBaCuO, magnetic neutron diffraction
studies have been performed on reduced YBa2Cu3_r Alx06+y single crystals with different Al content i
(0.06< i <0.19) and 0 content y (0.1S< y <0.36). All crystals showed a transition from the paramagnetic to the
AFI phase at T^ w400 K. The corresponding cell is doubled in a and b directions and equal to the chemical
cell along c, i.e. the moments on the Cu plane sites are antiferromagnetically long range and 3d ordered, while
no moment is allowed on the Cu chain sites by symmetry. The moments on nearestneighbor Cu plane layers,
which are antiferromagnetically coupled along the tetragonal caxis form an antiferromagne -tic bilayer system.
The average ordered moment is found to be0.58 (2) fiB and the spin-direction is lying within the ab plane.
These results are similar to those observed in the undoped system.

At low temperatures a complete reordering into a second antiferromagnetic phase AFII occurs at a
temperature T2. In this phase the cell is doubled in a and b as well as in c direction, with, in contrast to the

8
AFI phase, ferromagnetic coupling of the bilayers along the c axis. The ordered moment of 0.55(2) \IQ found
on the Cu plane-sites is similar to that observed for the AFI phase, the spins lie in the ab plane, too. Although
a moment on the Cu chain-sites is allowed by symmetry in this phase, no moment is found on these sites within
the accuracy of our data, which is in agreement with recent NMR studies. From our results we find a tendency
that T2 increases with increasing Al content x for the same 0 stoichiometry. The highest T2 of IS K is observed
for a crystal with 1=0.19 and y=0.36. Further, crystals, which were reduced under extreme conditions
suggesting maximal Al cluster formation during the reduction process, show a decrease of T2. Therefore
besides the Al content the distribution of Al on the Cu chain sites is also an important parameter effecting
the AFII reordering.

Figure 10: (a) Doping x dependence of antiferromagnetic (AF)


and superconducting (SC) transition temperatures in CeCo (In
1−x Cdx) 5 for Cd-content x ≤ 0.02. The crystals investigated
here (arrow) exhibit both transitions. (b) B-T phase diagram of
CeCo (In 0.9925 Cd 0.0075) 5 obtained from magnetotransport
(•,), neutron scattering (△, ▽) and heat capacity (+,×)
measurements with B ⊥ c. (c) B-T diagram for B c from
magnetotransport and heat capacity. Indications of a transition
within the AF phase are found (⋄). Reserved for Publication Footnotes.

6. Difference Between Spin Density Wave and Antiferromagnetism:

The salient properties of spin-density-wave (SDW) states of an electron gas are reviewed.
Antiferromagnetism in metals can arise from SDW's whether or not the metal contains localized magnetic
moments (with spin degrees of freedom). If there are no local moments, a SDW ground state of the conduction
electron gas is required, containing one or more collective-electron magnetization waves. Large amplitude
SDW's are observed in Cr and, very likely, small amplitude SDW's will eventually be observed in a number
of other metals, including some of the nontransition elements. Localized d or f electrons can be
antiferromagnetically ordered by SDW's even though the conduction electron gas, in which they are imbedded,
has of itself a nonmagnetic ground state. Exchange interactions between the localized-electron spins and the
spin polarization density of excited SDW's are responsible for excitation of the latter, which in turn orient and
order the former. Indirect exchange interactions (Ruderman-Kittel) between localized spins are shown to be
equivalent to, and derivable from a multiple SDW excitation of the electron gas. However, the indirect-
interaction model is inappropriate for a discussion of long range order, since the SDW excitations are then
concentrated in a few discrete waves. Observed effects such as truncation of the Fermi surface by repopulation
of k space, the energetic importance of commensurateness between the SDW wavelength and lattice constant,
and the occurrence of multiply periodic states of long range order are natural and immediate consequences of
SDW theory.
9
7. Antiferromagnetism High Temperature in Superconductivity:

Figure 11: Antiferromagnetism exists near superconductivity in the electronic phase diagram of a variety of
seemingly unrelated unconventional superconductors. Unconventional superconductors are those which are
not described by BCS theory.

7.1 Cuprate High-Temperature Superconductor:

From 1986, many cuprate superconductors were identified, and can be put into three groups on a phase
diagram critical temperature vs. oxygen hole content and copper hole content:

Lanthanum barium- (LB-CO), Tc=-240 °C (35 K).

Yttrium barium- (YB-CO), Tc=-215 °C (93 K).

Bismuth strontium calcium- (BiSC-CO), Tc=-180 °C (95 K).

Thallium barium calcium- (TBC-CO), Tc=-150 °C (125 K).

Mercury barium calcium- (HGBC-CO) 1993, with Tc=-140 °C


(133 K), currently the highest cuprate critical temperature.

Figure 12: Schematic phase diagram of cuprate superconductors as a function of temperature T and hole
density p. At half-filling, p=0, the materials are antiferromagnetic (AF) Mott insulators. The doping range p
0.16 (p 0.16) is underdoped (overdoped), and the optimal doping is p 0.16.

10
7.1.1. Structure:
Cuprates are layered materials, consisting of superconducting planes of
copper oxide, separated by layers containing ions such as lanthanum,
barium, strontium, which act as a charge reservoir, doping electrons or
holes into the copper-oxide planes. Thus the structure is described as a
superlattice of superconducting CuO2 layers separated by spacer layers,
resulting in a structure often closely related to the perovskite structure.
Superconductivity takes place within the copper-oxide (CuO2) sheets, with
only weak coupling between adjacent CuO2 planes, making the properties
close to that of a two-dimensional material. Electrical currents flow within
the CuO2 sheets, resulting in a large anisotropy in normal conducting and
superconducting properties, with a much higher conductivity parallel to the CuO2 plane than in the
perpendicular direction.

7.1.2. Superconducting Mechanism:

In conventional superconductors, Bardeen-Cooper-Schrieffer (BCS) theory describes how phonon


vibrations coax electrons together into Cooper pairs. The material properties of those superconductors often
abide by “Matthias’s rules”—no magnetism, no oxides, no insulators. Apart from sulfur hydrides, no BCS
superconductor exceeds temperatures of 40 K. None of that has stopped doped copper oxides, whose parent
compounds are insulating antiferromagnets, from remaining superconducting at temperatures as high as 135
K. As further evidence against a BCS pairing mechanism, cuprate superconductors are mostly iCuprate
superconductors vary in their chemical formulas, but all contain the same essential building block: planes with
one copper atom sandwiched between two oxygen atoms. Hypotheses abound for the mechanism behind
cuprates’ superconductivity. Some theorists have suggested spin fluctuations; others believe phonons are the
answer. Less than a year after Georg Bednorz and Alex Müller’s discovery of cuprate superconductivity, Philip
Anderson proposed that the glue that binds electrons comes from super exchange, in which the spins of copper
atoms are coupled, creating a magnetic attraction among their electrons despite the nonmagnetic oxygen atom
in between. Sensitive to changes in phonon frequency.

7.1.3 Re-cuprating Models:

Although BCS theory can be solved analytically—John Schrieffer famously solved the key equation for
the Cooper pairs on the subway—the theory behind High-Tc superconductors is more complex. To simplify
the picture, researchers have often turned to a one-band Hubbard model in which the cuprate is approximated
as a square lattice of spins. Anderson was able to use the model to show how super exchange might work;
others have even used it to predict where cuprate phase transitions take place. But the one-band Hubbard model
does not consider multiple electron orbitals between copper and oxygen because it essentially smashes the
oxygen and copper into one effective molecule.
11
Superconductivity in the cuprates is considered unconventional and is not explained by BCS theory.
Possible pairing mechanisms for cuprate superconductivity continue to be the subject of considerable debate
and further research. Similarities between the low-temperature antiferromagnetic state in undoped materials
and the low-temperature superconducting state that emerges upon doping, primarily the dx2-y2 orbital state
of the Cu2+ ions, suggest that electron-phonon coupling is less relevant in cuprates. Recent work on the Fermi
surface has shown that nesting occurs at four points in the antiferromagnetic Brillouin zone where spin waves
exist and that the superconducting energy gap is larger at these points. The weak isotope effects observed for
most cuprates contrast with conventional superconductors that are well described by BCS theory.

In 1987, Philip Anderson proposed that super exchange could act as a high-temperature superconductor
pairing mechanism. In 2016, Chinese physicists found a correlation between a cuprate's critical temperature
and the size of the charge transfer gap in that cuprate, providing support for the super exchange hypothesis. A
2022 study found that the varying density of actual Cooper pairs in a bismuth strontium calcium copper oxide
superconductor matched with numerical predictions based on super exchange.

7.2. Iron based superconductor:

Iron-based superconductors (FeSC) are iron-containing chemical compounds whose superconducting


properties were discovered in 2006. In 2008, led by recently discovered iron pnictide compounds (originally
known as oxypnictides). This new type of superconductors is based instead on conducting layers of iron and
a pnictide (chemical elements in group 15 of the periodic table, here typically arsenic (As) and phosphorus
(P)) and seems to show promise as the next generation of high temperature superconductors.

The oxypnictides such as LaOFeAs are often referred to as the '1111' pnictides.The crystalline material,
known chemically as LaOFeAs, stacks iron and arsenic layers, where the electrons flow, between planes of
lanthanum and oxygen. Replacing up to 11 percent of the oxygen with fluorine improved the compound — it
became superconductive at 26 kelvins, the team reports in the March 19, 2008 Journal of the American
Chemical Society. Subsequent research from other groups suggests that replacing the lanthanum in LaOFeAs
with other rare earth elements such as cerium, samarium, neodymium and praseodymium leads to
superconductors that work at 52 kelvins.

Iron pnictide superconductors crystallize into the [FeAs] layered structure alternating with spacer or
charge reservoir block. The compounds can thus be classified into “1111” system RFeAsO (R: the rare earth
element) including LaFeAsO, SmFeAsO, PrFeAsO, etc.; “122” type BaFe2As2, SrFe2As2 or CaFe2As2
“111” type LiFeAs, NaFeAs, and LiFeP. Doping or applied pressure will transform the compounds into
superconductors.

7.2.1. Superconductivity at high temperature:

High-temperature superconductors (abbreviated High-Tc or HTS) are defined as materials that behave
as superconductors at temperatures above 77 K (−196.2 °C; −321.1 °F), the boiling point of liquid nitrogen.

12
The main class of high-temperature superconductors is copper oxides combined with other metals,
especially the Rare-earth barium copper oxides (REBCOs) such as Yttrium barium copper oxide (YBCO).
The second class of high-temperature superconductors in the practical classification is the iron-based
compounds Magnesium diboride is sometimes included in high-temperature superconductors: It is relatively
simple to manufacture, but it superconducts only below 43 K, which makes it unsuitable for liquid nitrogen
cooling (approximately 30 K below nitrogen triple point temperature). Some extremely-high pressure
superhydride compounds are usually categorized as high-
temperature superconductors.

Superconducting transition temperatures are listed in


the tables (some at high pressure). BaFe1.8Co0.2As2 is
predicted to have an upper critical field of 43 tesla from
the measured coherence length of 2.8 nm.

Figure 13: Phase diagram of the 122 family of ferro-


pnictides complemented by the 122(Se) family as a
generalized phase diagram for the iron based superconductors.

7.3. Heavy fermion superconductor:

Heavy fermion superconductors are a type of unconventional superconductor. The first heavy fermion
superconductor, CeCu2Si2, was discovered by Frank Steglich in 1978.

Table 2: over 30 heavy fermion superconductors were found (in materials based on Ce, U), with a critical
temperature up to 2.3 K (in CeCoIn5).

Material 𝑇𝑐 (K) Comments

CeCu2Si2 0.7 First unconventional superconductor

CeCoIn5 2.3 Highest TC of all Ce-based heavy fermions

CePt3Si 0.75 First heavy-fermions superconductor with non-centrosymmetric crystal structure

CeIn3 0.2 Superconducting only at high pressure

UBe13 0.85 p-wave superconductor

UPt3 0.48 Several distinct superconducting phases

URu2Si2 1.3 Mysterious hidden-order phase below 17k

UPd2A13 2.0 Antiferromagnetism below 14k

13
Uni2A13 1.1 Antiferromagnetism below 5k

Heavy Fermion materials are intermetallic compounds, containing rare earth or actinide elements. The
f-electrons of these atoms hybridize with the normal conduction electrons leading to quasiparticles with an
enhanced effective mass.

From specific heat measurements (ΔC/C (TC) one knows that the Cooper pairs in the superconducting
state are also formed by the heavy quasiparticles. In contrast to normal superconductors it cannot be described
by BCS-Theory. Due to the large effective mass, the Fermi velocity is reduced and comparable to the inverse
Debye frequency. This leads to the failing of the picture of electrons polarizing the lattice as an attractive force.

Some heavy fermion superconductors are candidate materials for the Fulde-Ferrell-Larkin-Ovchinnikov
(FFLO) phase. In particular, there has been evidence that CeCoIn5 close to the critical field is in an FFLO stat.

7.4. Organic Superconductor:

An organic superconductor is a synthetic organic compound that exhibits superconductivity at low


temperatures. Many materials may be characterized as organic superconductors. These include the Bechgaard
salts and Fabre salts which are both quasi-one-dimensional, and quasi-two-dimensional materials such as k-
BEDT-TTF2X charge-transfer complex, λ-BETS2X compounds, graphite intercalation compounds and three-
dimensional materials such as the alkali-doped fullerenes.

Organic superconductors are of special interest not only for scientists, looking for room-temperature
superconductivity and for model systems explaining the origin of superconductivity but also for daily life
issues as organic compounds are mainly built of carbon and hydrogen which belong to the most common
elements on earth in contrast to copper or osmium.

7.4.1. One-dimensional Fabre and Bechgaard salts:

Table-3: some examples of One-dimensional Fabre Material 𝑇𝑐 (K) pext (kbar)


and Bechgaard salts
(TMTSF)2SbF6 0.36 10.5
Fabre-salts are composed of
tetramethyltetrathiafulvalene (TMTTF) and (TMTSF)2PF6 1.1 6.5

Bechgaard salts of tetramethyltetraselenafulvalene


(TMTSF)2AsF6 1.1 9.5
(TMTSF). These two organic molecules are similar
except for the sulfur-atoms of TMTTF being replaced (TMTSF)2ReO4 1.2 9.5
by selenium-atoms in TMTSF. The molecules are
(TMTSF)2TaF6 1.35 11
stacked in columns (with a tendency to dimerization)
which are separated by anions. Typical anions are, for (TMTTF)2Br 0.8 26
example, octahedral PF6, AsF6 or tetrahedral ClO4 or

14
ReO4. A selection of the transition temperature and corresponding external pressure of several one-
dimensional organic superconductors is shown in the table above.

7.4.2.Two-dimensional (BEDT-TTF) 2X:

BEDT-TTF is the short form of bisethylenedithio-tetrathiafulvalene commonly abbreviated with ET.


These molecules form planes which are separated by anions. The pattern of the molecules in the planes is not
unique but there are several different phases Material 𝑇𝑐 (K) pext (kbar)
growing, depending on the anion and the growth
conditions. Important phases concerning βH-(ET)2I3 1.5 0
superconductivity are the α- and θ- phase with the
θ-(ET)2I3 3.6 0
molecules ordering in a fishbone structure and the
β- and especially κ-phase which order in a k-(ET)2I3 3.6 0
checkerboard structure with molecules being
dimerized in the κ-phase. α-(ET)2KHg(SCN)4 0.3 0

The following table shows only a few exemplary α-(ET)2KHg(SCN)4 1.2 1.2
superconductors of this class.
β’’-(ET)2SF5CH2CF2SO3 5.3 0
Table 4: some examples of two-dimensional
(BEDT-TTF) 2X κ-(ET)2Cu[N(CN)2]Cl 12.8 0.3

Some two-dimensional organic κ-(ET)2Cu[N(CN)2]Cl 13.1 0.3


superconductors of the κ-(ET)2X and λ(BETS)2X deuterated
families are candidates for the Fulde-Ferrell-
Larkin-Ovchinnikov (FFLO) phase when κ-(ET)2Cu[N(CN)2]Br 11.2 0
superconductivity is suppressed by an external deuterated
magnetic field.
κ-(ET)2Cu(NCS)2 10.4 0
7.5. Doped fullerenes:
κ-(ET)4Hg2.89Cl8 1.8 12
Superconducting fullerenes based on C60 are
fairly different from other organic κH-(ET)2Cu(CF3)4·TCE 9.2 0
superconductors. The building molecules are no
κH-(ET)2Ag(CF3)4·TCE 11.1 0
longer manipulated hydrocarbons but
pure carbon molecules. In addition these
molecules are no longer flat but bulky which gives rise to a three-dimensional, isotropic superconductor. The
pure C60 grows in an fcc-lattice and is an insulator. By placing alkali atoms in the interstitials the crystal
becomes metallic and eventually superconducting at low temperatures. Unfortunately, C60 crystals are not
stable at ambient atmosphere. They are grown and investigated in closed capsules, limiting the measurement
techniques possible. The highest transition temperature measured so far was TC = 33 K for Cs2RbC60.The
15
highest measured transition temperature of an
DoMaterial 𝑇𝑐 (K) pext (mbar)
organic superconductor was found in 1995 in
Cs3C60 pressurized with 15 kbar to be TC = 40 K. Under K3C60 18 0
pressure this compound shows a unique behavior.
Rb3C60 30.7 0
Usually the highest TC is achieved with the lowest
pressure necessary to drive the transition. Further K2CsC60 24 0
increase of the pressure usually reduces the transition
temperature. However, in Cs3C60 superconductivity K2RbC60 21.5 0

sets in at very low pressures of several 100 bar, and the


K5C60 8.4 0
transition temperature keeps increasing with increasing
pressure. This indicates a completely different Sr6C60 6.8 0
mechanism then just broadening of the bandwidth.
(NH3)4Na2CsC60 29.6 0
Table 5: some examples of doped fullerences materials
(NH3)K3C60 28 14.8
8. Comparison Between Cuprates and Iron
Pnictide:
Table 6: Comparison between cuprates and iron printed

Iron based superconductor 𝑀𝑔𝐵2 cuprates

Parent material Antiferromagnetic Pauli Antiferromagnetic Mott


semimetal (Bad metal) paramagnetic insulator (Ts~400K)
(Ts ~ 150 K) metal

Fermi level Fe 3d 5-orbitals B 2p 2-orbitals Cu 3d single orbitals

Max Tc(k) 56 K(for 1111 type) 40 K ~140K

38K (for 122 type)

Correlation effect Long range (non local) None nearly free Strong local electronic
electron interaction
Magnetic correlation

Impurity Robust Sensitive Sensitive

Superconducting gap symmetry Extended s~Wave(+-or S~Wave d~Wave


++)

16
9. Conclusion:
Magnetism is the result of attraction, when two objects come together, or repulsion when two objects
move apart. A magnet is an object that has properties of magnetism. For example, a magnet might attract
another object. A magnetic field is an invisible area around a magnet where magnetism occurs. These are three
types. Such as diamagnetism, paramagnetism, ferromagnetism.

Equal magnitude of magnetic moment associated with unpaired electrons are aligned in opposite
directions, the net magnetic moment is zero is called antiferromagnetism. Example: MnO, Mn2O3 and
MnO2.Ferromagnetism is observed in transition metals and some of their compounds. In
antiferromagnetism, the magnetic moments point in the opposite direction. Due to this, the magnetic
susceptibility of a substance decreases to some extent. Antiferromagnetism is observed in the salts of ions like
Mn+2, Fe+3 and Gd+3. Antiferromagnetic materials occur commonly among transition metal compounds,
especially oxides. Examples include hematite, metals such as chromium, alloys such as iron manganese
(FeMn), and oxides such as nickel oxide (NiO). There are also numerous examples among high nuclearity
metal clusters. Antiferromagnetic materials improve the way information is written and read electrically in
devices. They are microscopic magnets with opposite orientations. Computers nowadays use Silicon
components, but they fail to be as efficient as antiferromagnetic materials.

High-transition temperature copper oxide superconductors have puzzled physicists since these materials
were discovered 10 years ago. Nagaosa's Perspective discusses a new theory described by Zhang (1089) in the
same issue in which superconductivity is related to another important property, antiferromagnetism, by an
elegant symmetry principle. Antiferromagnetism exists near superconductivity in the electronic phase diagram
of a variety of seemingly unrelated unconventional superconductors. Unconventional superconductors are
those which are not described by BCS theory. Antiferromagnetism (abbreviated AF) is where the spin on
adjacent magnetic sites points in opposite directions. Below are some classes of superconductors where
antiferromagnetism seems to have something to do with superconductivity. Note that each class comprises
many different materials, so in actuality, there are hundreds of compounds which share this
property.Cuprate High-temperature superconductors, Iron-based (pnictide) sort-of-high temperature
superconductors and Organic superconductor.

10. Reference:

1. The magnetism of matter, Feynman Lectures in Physics Ch 34.

2. "Diamagnetic Levitation". High Field Laboratory. Radboud University Nijmegen. 2011.

3. Paramagnetism. Encyclopædia Britannica.

4. Aharoni, Amikam (1996). Introduction to the Theory of Ferromagnetism. Clarendon Press. ISBN 0-19-
851791-2.

17
5. Álvarez, Nadia (2016). Dominios magnéticos y respuesta dinámica en aleaciones ferromagnéticas de
FEP [Magnetic domains and dynamical response in ferromagnetic FePt alloys] (PhD) (in Spanish).
Universidad Nacional de Cuyo. Docket 564.

6. M. Louis Néel (1948). "Propriétées magnétiques des ferrites; Férrimagnétisme et


antiferromagnétisme" (PDF). Annales de Physique. 12 (3): 137–198. Bibcode:1948AnPh...12..137N.

7. František, Hrouda (September 1, 2002). "Low-field variation of magnetic susceptibility and its effect on the
anisotropy of magnetic susceptibility of rocks". Geophysical Journal International. Oxford University
Press. 150 (3): 715–723.

8. Hsu, F. C. et al. Superconductivity in the PbO-type structure α-FeSe. Proc. Natl Acad. Sci. USA 105,
14262–14264 (2008).

9. McQueen, T. M. et al. Extreme sensitivity of superconductivity to stoichiometry in Fe1+δSe. Phys.


Rev. B 79, 014522 (2009).

10. Liu, D. F. et al. Electronic origin of high-temperature superconductivity in single-layer FeSe


superconductor. Nature Commun. 3, 931 (2012).

11. He, S. L. et al. Phase diagram and electronic indication of high-temperature superconductivity at 65 K in
single-layer FeSe films. Nature Mater. 12, 605–610 (2013).

12. Tan, S. Y. et al. Interface-induced superconductivity and strain-dependent spin density waves in
FeSe/SrTiO3 thin films. Nature Mater. 12, 634–640 (2013).

13.Li, W. et al. Phase separation and magnetic order in K-doped iron selenide superconductor. Nature
Phys. 8, 126–130 (2012).

14.Bao, W. et al. A novel large moment antiferromagnetic order in K0.8Fe1.6Se2 superconductor. Chin.
Phys. Lett. 28, 086104 (2011).

15.J. G. Bednorz; K. A. Mueller (1986). "Possible high TC superconductivity in the Ba-La-Cu-O system". Z.
Phys. B. 64 (2): 189–93. Bibcode:1986ZPhyB..64..189B.

16. Mark Buchanan (2001). "Mind the pseudogap". Nature. 409 (6816): 8–1. doi:10.1038/35051238

17. Sheng, Z. Z.; Hermann A. M. (1988). "Bulk superconductivity at 120 K in the Tl–Ca/Ba–Cu–O
system". Nature. 332 (6160): 138–39. Bibcode:1988Natur.332..138S.

18. Schilling, A.; Cantoni, M.; Guo, J. D.; Ott, H. R. (1993). "Superconductivity above 130 K in the Hg–Ba–
Ca–Cu–O system". Nature. 363 (6424): 56–8. Bibcode:1993Natur.363...56S.

19. Lee, Patrick A. (2008). "From high temperature superconductivity to quantum spin liquid: progress in
strong correlation physics". Reports on Progress in Physics. 71: 12501. arXiv:0708.2115.

18

You might also like