bacana play bonus

mitzvahceremonies.com:2024/9/11 3:29:59

  1. bacana play bonus
  2. jogos de matemática online

bacana play bonus

  
live      

Os pilotos ucranianos começaram a pilotar F-16, disse Volodymyro Zelenskiy afirmando que os caça de fabricação norte-americana chegaram há mais 💸 do 29 meses desde o início da invasão russa.

O líder ucraniano anunciou o uso de F-16, que Kyiv há muito 💸 tempo faz lobby para s.a quando ele conheceu pilotos militares bacana play bonus uma base aérea ladeada por dois dos jatos com 💸 mais duas sobrevoando a aeronave

"Os F-16 estão na Ucrânia. Nós fizemos isso, estou orgulhoso de nossos caras que dominam esses 💸 jatos e já começaram a usá-los para o nosso país", disse Zelenskiy bacana play bonus um local onde as autoridades pediram aos 💸 repórteres não divulgar por razões da segurança ”.

A chegada dos jatos é um marco para a Ucrânia após muitos meses 💸 de espera, embora ainda não esteja claro quantos estão disponíveis e quanto impacto terão no aprimoramento das defesas aéreas.

A Rússia 💸 tem alvejado bases que podem abrigá-las e prometeu derrubálos.

Os F-16 estavam na lista de desejos da Ucrânia há muito tempo, 💸 equipados com um canhão 20mm e podem transportar bombas.

Dois F-16 voaram sobre a base aérea enquanto Zelenskiy falava.


bacana play bonus
: Valentyn Ogirenko/Reuters

Falando 💸 com repórteres na pista de um aeródromo, Zelenskiy disse que ainda não há pilotos suficientes treinados para usar os F-16 💸 ou o suficiente dos próprios jatos.

"O positivo é que estamos esperando mais F-16s... muitos caras estão treinando agora", disse ele.

A 💸 Ucrânia já contou anteriormente com uma frota envelhecida de aviões da era soviética que são superados pela mais avançada e 💸 muito maior esquadra russa.

skip promoção newsletter passado

Nosso e-mail da manhã detalha as principais histórias do dia, dizendo o que está 💸 acontecendo.

Aviso de Privacidade:

As newsletters podem conter informações sobre instituições de caridade, anúncios on-line e conteúdo financiado por terceiros. Para mais 💸 informação consulte a nossa Política De Privacidade Utilizamos o Google reCaptcha para proteger nosso site; se aplica também à 💸 política do serviço ao cliente da empresa:

após a promoção da newsletter;

A Rússia usou essa vantagem para realizar ataques regulares de 💸 mísseis bacana play bonus alvos na Ucrânia e bater posições ucranianas da linha dianteira com milhares das bombas guiada, apoiando suas forças 💸 que estão avançando lentamente no leste.

"Este é o novo estágio de desenvolvimento da força aérea das forças armadas ucranianas", disse 💸 Zelenskiy.

"Fizemos muito para que as forças ucranianas fizessem a transição de um novo padrão da aviação, o combate ocidental", acrescentou 💸 ele citando centenas das reuniões e diplomacia implacável na obtenção dos F-16.

Ainda não está claro com quais mísseis os jatos 💸 estão equipados. Um alcance mais longo de míssil permitiria que eles tivessem um impacto maior no campo do batalha, dizem 💸 analistas militares

Zelenskiy disse que também espera fazer lobby entre os países vizinhos aliados para ajudar a interceptar mísseis russos sendo 💸 lançados na Ucrânia por meio de conversas no Conselho OTAN-Ucrânia.

"Esta é outra ferramenta, e eu quero tentar isso para que 💸 os países da Otan possam conversar com a Ucrânia sobre uma pequena coalizão de vizinhos derrubando mísseis inimigos", disse ele.

"Acho 💸 que essa decisão é provavelmente difícil para nossos parceiros, eles sempre têm medo de uma escalada excessiva mas estamos lutando 💸 contra isso."

  • casino holdem online free
  • No Brasil a educação em Educação Física (EiF) surgiu na década de 1950, período em que a educação física se 😆 tornou mais centralizado nos centros de educação de nível superior.

    Foi uma forma de educação do conhecimento superior, focada nas necessidades 😆 educativas e no desenvolvimento das competências cognitivas do indivíduo.

    A partir da década de 1980, surgiu a ideia de formar profissionais 😆 de gestão em educação física que explondessem o conhecimento das ciências e do comportamento social

    por meio da educação física.

    É nesse 😆 contexto, que surgiu, em 1980, o Fórum Educativo, que foi criado por representantes das entidades de educação para a educação 😆 física.

  • 7games baixar app de download
  • jogos de matemática online


    bacana play bonus

    Intrinsic quantum property of particles

    This article is about the concept in quantum

    mechanics. For the concept in classical mechanics, see 🧲 Rotation

    Spin is an intrinsic

    form of angular momentum carried by elementary particles, and thus by composite

    particles such as hadrons, 🧲 atomic nuclei, and atoms.[1][2]: 183–184 Spin should not be

    conceptualized as involving the "rotation" of a particle's "internal mass", as 🧲 ordinary

    use of the word may suggest: spin is a quantized property of waves.[3]

    The existence of

    electron spin angular momentum 🧲 is inferred from experiments, such as the Stern–Gerlach

    experiment, in which silver atoms were observed to possess two possible discrete

    🧲 angular momenta despite having no orbital angular momentum.[4] The existence of the

    electron spin can also be inferred theoretically from 🧲 the spin–statistics theorem and

    from the Pauli exclusion principle—and vice versa, given the particular spin of the

    electron, one may 🧲 derive the Pauli exclusion principle.

    Spin is described

    mathematically as a vector for some particles such as photons, and as spinors 🧲 and

    bispinors for other particles such as electrons. Spinors and bispinors behave similarly

    to vectors: they have definite magnitudes and 🧲 change under rotations; however, they use

    an unconventional "direction". All elementary particles of a given kind have the same

    magnitude 🧲 of spin angular momentum, though its direction may change. These are

    indicated by assigning the particle a spin quantum number.[2]: 🧲 183–184

    The SI unit of

    spin is the same as classical angular momentum (i.e., N·m·s, J·s, or kg·m2·s−1). In

    practice, spin 🧲 is usually given as a dimensionless spin quantum number by dividing the

    spin angular momentum by the reduced Planck constant 🧲 ħ, which has the same dimensions

    as angular momentum. Often, the "spin quantum number" is simply called "spin".

    Relation

    to classical 🧲 rotation [ edit ]

    The very earliest models for electron spin imagined a

    rotating charged mass, but this model fails when 🧲 examined in detail: the required space

    distribution does not match limits on the electron radius: the required rotation speed

    exceeds 🧲 the speed of light. In the Standard Model, the fundamental particles are all

    considered "point-like": they have their effects through 🧲 the field that surrounds

    them.[5] Any model for spin based on mass rotation would need to be consistent with

    that 🧲 model.

    The classical analog for quantum spin is a circulation of energy or

    momentum-density in the particle wave field: "spin is 🧲 essentially a wave property".[3]

    This same concept of spin can be applied to gravity waves in water: "spin is generated

    🧲 by subwavelength circular motion of water particles".[6]

    Photon spin is the

    quantum-mechanical description of light polarization, where spin +1 and spin 🧲 −1

    represent two opposite directions of circular polarization. Thus, light of a defined

    circular polarization consists of photons with the 🧲 same spin, either all +1 or all −1.

    Spin represents polarization for other vector bosons as well.

    Relation to orbital

    angular 🧲 momentum [ edit ]

    As the name suggests, spin was originally conceived as the

    rotation of a particle around some axis. 🧲 Historically orbital angular momentum related

    to particle orbits.[7]: 131 While the names based on mechanical models have survived,

    the physical 🧲 explanation has not. Quantization fundamentally alters the character of

    both spin and orbital angular momentum.

    Since elementary particles are point-like,

    self-rotation 🧲 is not well-defined for them. However, spin implies that the phase of the

    particle depends on the angle as e 🧲 i S θ {\displaystyle e^{iS\theta }} , for rotation

    of angle θ around the axis parallel to the spin S. 🧲 This is equivalent to the

    quantum-mechanical interpretation of momentum as phase dependence in the position, and

    of orbital angular momentum 🧲 as phase dependence in the angular position.

    For fermions,

    the picture is less clear. Angular velocity is equal by Ehrenfest theorem 🧲 to the

    derivative of the Hamiltonian to its conjugate momentum, which is the total angular

    momentum operator J = L 🧲 + S. Therefore, if the Hamiltonian H is dependent upon the spin

    S, dH/dS is non-zero, and the spin causes 🧲 angular velocity, and hence actual rotation,

    i.e. a change in the phase-angle relation over time. However, whether this holds for

    🧲 free electron is ambiguous, since for an electron, S2 is constant, and therefore it is

    a matter of interpretation whether 🧲 the Hamiltonian includes such a term. Nevertheless,

    spin appears in the Dirac equation, and thus the relativistic Hamiltonian of the

    🧲 electron, treated as a Dirac field, can be interpreted as including a dependence in the

    spin S.[8] Under this interpretation, 🧲 free electrons also self-rotate, with the

    zitterbewegung effect understood as this rotation.

    Quantum number [ edit ]

    Spin obeys

    the mathematical laws 🧲 of angular momentum quantization. The specific properties of spin

    angular momenta include:

    Spin quantum numbers may take half-integer values.

    Although

    the direction 🧲 of its spin can be changed, the magnitude of the spin of an elementary

    particle cannot be changed.

    The spin of 🧲 a charged particle is associated with a

    magnetic dipole moment with a g -factor that differs from 1. (In the 🧲 classical context,

    this would imply the internal charge and mass distributions differing for a rotating

    object.[9])

    The conventional definition of the 🧲 spin quantum number is s = n/2, where n

    can be any non-negative integer. Hence the allowed values of s 🧲 are 0, 1/2, 1, 3/2, 2,

    etc. The value of s for an elementary particle depends only on the type 🧲 of particle and

    cannot be altered in any known way (in contrast to the spin direction described below).

    The spin 🧲 angular momentum S of any physical system is quantized. The allowed values of

    S are

    S = ℏ s ( s 🧲 + 1 ) = h 2 π n 2 ( n + 2 ) 2 = h 4 π n 🧲 ( n + 2 ) , {\displaystyle

    S=\hbar \,{\sqrt {s(s+1)}}={\frac {h}{2\pi }}\,{\sqrt {{\frac {n}{2}}{\frac

    {(n+2)}{2}}}}={\frac {h}{4\pi }}\,{\sqrt {n(n+2)}},}

    h

    ℏ = h 🧲 2 π {\textstyle \hbar

    ={\frac {h}{2\pi }}}

    s

    n

    Fermions and bosons [ edit ]

    whereis the Planck constant ,

    andis the reduced Planck 🧲 constant. In contrast, orbital angular momentum can only take

    on integer values of; i.e., even-numbered values of

    Those particles with half-integer

    🧲 spins, such as 1/2, 3/2, 5/2, are known as fermions, while those particles with integer

    spins, such as 0, 1, 🧲 2, are known as bosons. The two families of particles obey

    different rules and broadly have different roles in the 🧲 world around us. A key

    distinction between the two families is that fermions obey the Pauli exclusion

    principle: that is, 🧲 there cannot be two identical fermions simultaneously having the

    same quantum numbers (meaning, roughly, having the same position, velocity and 🧲 spin

    direction). Fermions obey the rules of Fermi–Dirac statistics. In contrast, bosons obey

    the rules of Bose–Einstein statistics and have 🧲 no such restriction, so they may "bunch

    together" in identical states. Also, composite particles can have spins different from

    their 🧲 component particles. For example, a helium-4 atom in the ground state has spin 0

    and behaves like a boson, even 🧲 though the quarks and electrons which make it up are all

    fermions.

    This has some profound consequences:

    Spin–statistics theorem [ edit ]

    The

    🧲 spin–statistics theorem splits particles into two groups: bosons and fermions, where

    bosons obey Bose–Einstein statistics, and fermions obey Fermi–Dirac statistics 🧲 (and

    therefore the Pauli exclusion principle). Specifically, the theory states that

    particles with an integer spin are bosons, while all 🧲 other particles have half-integer

    spins and are fermions. As an example, electrons have half-integer spin and are

    fermions that obey 🧲 the Pauli exclusion principle, while photons have integer spin and

    do not. The theorem relies on both quantum mechanics and 🧲 the theory of special

    relativity, and this connection between spin and statistics has been called "one of the

    most important 🧲 applications of the special relativity theory".[11]

    Magnetic moments [

    edit ]

    Schematic diagram depicting the spin of the neutron as the black 🧲 arrow and

    magnetic field lines associated with the neutron magnetic moment. The neutron has a

    negative magnetic moment. While the 🧲 spin of the neutron is upward in this diagram, the

    magnetic field lines at the center of the dipole are 🧲 downward.

    Particles with spin can

    possess a magnetic dipole moment, just like a rotating electrically charged body in

    classical electrodynamics. These 🧲 magnetic moments can be experimentally observed in

    several ways, e.g. by the deflection of particles by inhomogeneous magnetic fields in 🧲 a

    Stern–Gerlach experiment, or by measuring the magnetic fields generated by the

    particles themselves.

    The intrinsic magnetic moment μ of a 🧲 spin- 1/2 particle with

    charge q, mass m, and spin angular momentum S, is[12]

    μ = g s q 2 m 🧲 S , {\displaystyle

    {\boldsymbol {\mu }}={\frac {g_{\text{s}}q}{2m}}\mathbf {S} ,}

    where the dimensionless

    quantity g s is called the spin g-factor. For 🧲 exclusively orbital rotations it would be

    1 (assuming that the mass and the charge occupy spheres of equal radius).

    The electron,

    🧲 being a charged elementary particle, possesses a nonzero magnetic moment. One of the

    triumphs of the theory of quantum electrodynamics 🧲 is its accurate prediction of the

    electron g-factor, which has been experimentally determined to have the value

    −2.00231930436256(35), with the 🧲 digits in parentheses denoting measurement uncertainty

    in the last two digits at one standard deviation.[13] The value of 2 arises 🧲 from the

    Dirac equation, a fundamental equation connecting the electron's spin with its

    electromagnetic properties, and the deviation from −2 🧲 arises from the electron's

    interaction with the surrounding electromagnetic field, including its own

    field.[14]

    Composite particles also possess magnetic moments associated 🧲 with their

    spin. In particular, the neutron possesses a non-zero magnetic moment despite being

    electrically neutral. This fact was an 🧲 early indication that the neutron is not an

    elementary particle. In fact, it is made up of quarks, which are 🧲 electrically charged

    particles. The magnetic moment of the neutron comes from the spins of the individual

    quarks and their orbital 🧲 motions.

    Neutrinos are both elementary and electrically

    neutral. The minimally extended Standard Model that takes into account non-zero

    neutrino masses predicts 🧲 neutrino magnetic moments of:[15][16][17]

    μ ν ≈ 3 × 10 − 19 μ

    B m ν c 2 eV , {\displaystyle 🧲 \mu _{

    u }\approx 3\times 10^{-19}\mu _{\text{B}}{\frac

    {m_{

    u }c^{2}}{\text{eV}}},}

    where the μ ν are the neutrino magnetic moments, m ν are

    the 🧲 neutrino masses, and μ B is the Bohr magneton. New physics above the electroweak

    scale could, however, lead to significantly 🧲 higher neutrino magnetic moments. It can be

    shown in a model-independent way that neutrino magnetic moments larger than about 10−14

    🧲 μ B are "unnatural" because they would also lead to large radiative contributions to

    the neutrino mass. Since the neutrino 🧲 masses are known to be at most about 1 eV/c2,

    fine-tuning would be necessary in order to prevent large contributions 🧲 to the neutrino

    mass via radiative corrections.[18] The measurement of neutrino magnetic moments is an

    active area of research. Experimental 🧲 results have put the neutrino magnetic moment at

    less than 1.2×10−10 times the electron's magnetic moment.

    On the other hand elementary

    🧲 particles with spin but without electric charge, such as a photon or a Z boson, do not

    have a magnetic 🧲 moment.

    Curie temperature and loss of alignment [ edit ]

    In ordinary

    materials, the magnetic dipole moments of individual atoms produce magnetic 🧲 fields that

    cancel one another, because each dipole points in a random direction, with the overall

    average being very near 🧲 zero. Ferromagnetic materials below their Curie temperature,

    however, exhibit magnetic domains in which the atomic dipole moments spontaneously

    align locally, 🧲 producing a macroscopic, non-zero magnetic field from the domain. These

    are the ordinary "magnets" with which we are all familiar.

    In 🧲 paramagnetic materials,

    the magnetic dipole moments of individual atoms will partially align with an externally

    applied magnetic field. In diamagnetic 🧲 materials, on the other hand, the magnetic

    dipole moments of individual atoms align oppositely to any externally applied magnetic

    field, 🧲 even if it requires energy to do so.

    The study of the behavior of such "spin

    models" is a thriving area 🧲 of research in condensed matter physics. For instance, the

    Ising model describes spins (dipoles) that have only two possible states, 🧲 up and down,

    whereas in the Heisenberg model the spin vector is allowed to point in any direction.

    These models 🧲 have many interesting properties, which have led to interesting results in

    the theory of phase transitions.

    Direction [ edit ]

    Spin projection 🧲 quantum number and

    multiplicity [ edit ]

    In classical mechanics, the angular momentum of a particle

    possesses not only a magnitude 🧲 (how fast the body is rotating), but also a direction

    (either up or down on the axis of rotation of 🧲 the particle). Quantum-mechanical spin

    also contains information about direction, but in a more subtle form. Quantum mechanics

    states that the 🧲 component of angular momentum for a spin-s particle measured along any

    direction can only take on the values[19]

    S i = 🧲 ℏ s i , s i ∈ { − s , − ( s − 1 ) , … ,

    🧲 s − 1 , s } , {\displaystyle S_{i}=\hbar s_{i},\quad s_{i}\in \{-s,-(s-1),\dots

    ,s-1,s\},}

    where S i is the spin component along 🧲 the i-th axis (either x, y, or z), s i

    is the spin projection quantum number along the i-th axis, 🧲 and s is the principal spin

    quantum number (discussed in the previous section). Conventionally the direction chosen

    is the z 🧲 axis:

    S z = ℏ s z , s z ∈ { − s , − ( s − 1 ) 🧲 , … , s − 1 , s } ,

    {\displaystyle S_{z}=\hbar s_{z},\quad s_{z}\in \{-s,-(s-1),\dots ,s-1,s\},}

    where S z

    is the 🧲 spin component along the z axis, s z is the spin projection quantum number along

    the z axis.

    One can see 🧲 that there are 2s + 1 possible values of s z . The number "2s +

    1" is the multiplicity 🧲 of the spin system. For example, there are only two possible

    values for a spin- 1/2 particle: s z = 🧲 + 1/2 and s z = − 1/2. These correspond to

    quantum states in which the spin component is pointing 🧲 in the +z or −z directions

    respectively, and are often referred to as "spin up" and "spin down". For a 🧲 spin- 3/2

    particle, like a delta baryon, the possible values are + 3/2, + 1/2, − 1/2, −

    3/2.

    Vector [ 🧲 edit ]

    A single point in space can rotate continuously without becoming

    tangled. Notice that after a 360-degree rotation, the spiral 🧲 flips between clockwise

    and counterclockwise orientations. It returns to its original configuration after

    spinning a full 720°.

    For a given quantum 🧲 state, one could think of a spin vector ⟨ S ⟩

    {\textstyle \langle S\rangle } whose components are the expectation 🧲 values of the spin

    components along each axis, i.e., ⟨ S ⟩ = [ ⟨ S x ⟩ , ⟨ 🧲 S y ⟩ , ⟨ S z ⟩ ] {\textstyle

    \langle S\rangle =[\langle S_{x}\rangle ,\langle S_{y}\rangle ,\langle S_{z}\rangle ]}

    . 🧲 This vector then would describe the "direction" in which the spin is pointing,

    corresponding to the classical concept of the 🧲 axis of rotation. It turns out that the

    spin vector is not very useful in actual quantum-mechanical calculations, because it

    🧲 cannot be measured directly: s x , s y and s z cannot possess simultaneous definite

    values, because of a 🧲 quantum uncertainty relation between them. However, for

    statistically large collections of particles that have been placed in the same pure

    🧲 quantum state, such as through the use of a Stern–Gerlach apparatus, the spin vector

    does have a well-defined experimental meaning: 🧲 It specifies the direction in ordinary

    space in which a subsequent detector must be oriented in order to achieve the 🧲 maximum

    possible probability (100%) of detecting every particle in the collection. For spin-

    1/2 particles, this probability drops off smoothly 🧲 as the angle between the spin vector

    and the detector increases, until at an angle of 180°—that is, for detectors 🧲 oriented

    in the opposite direction to the spin vector—the expectation of detecting particles

    from the collection reaches a minimum of 🧲 0%.

    As a qualitative concept, the spin vector

    is often handy because it is easy to picture classically. For instance,

    quantum-mechanical 🧲 spin can exhibit phenomena analogous to classical gyroscopic

    effects. For example, one can exert a kind of "torque" on an 🧲 electron by putting it in

    a magnetic field (the field acts upon the electron's intrinsic magnetic dipole

    moment—see the following 🧲 section). The result is that the spin vector undergoes

    precession, just like a classical gyroscope. This phenomenon is known as 🧲 electron spin

    resonance (ESR). The equivalent behaviour of protons in atomic nuclei is used in

    nuclear magnetic resonance (NMR) spectroscopy 🧲 and imaging.

    Mathematically,

    quantum-mechanical spin states are described by vector-like objects known as spinors.

    There are subtle differences between the behavior 🧲 of spinors and vectors under

    coordinate rotations. For example, rotating a spin- 1/2 particle by 360° does not bring

    it 🧲 back to the same quantum state, but to the state with the opposite quantum phase;

    this is detectable, in principle, 🧲 with interference experiments. To return the particle

    to its exact original state, one needs a 720° rotation. (The Plate trick 🧲 and Möbius

    strip give non-quantum analogies.) A spin-zero particle can only have a single quantum

    state, even after torque is 🧲 applied. Rotating a spin-2 particle 180° can bring it back

    to the same quantum state, and a spin-4 particle should 🧲 be rotated 90° to bring it back

    to the same quantum state. The spin-2 particle can be analogous to a 🧲 straight stick

    that looks the same even after it is rotated 180°, and a spin-0 particle can be

    imagined as 🧲 sphere, which looks the same after whatever angle it is turned

    through.

    Mathematical formulation [ edit ]

    Operator [ edit ]

    Spin obeys 🧲 commutation

    relations[20] analogous to those of the orbital angular momentum:

    [ S ^ j , S ^ k ] = i

    🧲 ℏ ε j k l S ^ l , {\displaystyle \left[{\hat {S}}_{j},{\hat {S}}_{k}\right]=i\hbar

    \varepsilon _{jkl}{\hat {S}}_{l},}

    ε jkl

    S ^ 2 {\displaystyle 🧲 {\hat {S}}^{2}}

    S ^ z

    {\displaystyle {\hat {S}}_{z}}

    S

    : 166

    S ^ 2 | s , m s ⟩ = ℏ 2 s 🧲 ( s + 1 ) | s , m s ⟩

    , S ^ z | s , m s 🧲 ⟩ = ℏ m s | s , m s ⟩ . {\displaystyle {\begin{aligned}{\hat

    {S}}^{2}|s,m_{s}\rangle &=\hbar ^{2}s(s+1)|s,m_{s}\rangle ,\\{\hat

    {S}}_{z}|s,m_{s}\rangle &=\hbar 🧲 m_{s}|s,m_{s}\rangle .\end{aligned}}}

    whereis the

    Levi-Civita symbol . It follows (as with angular momentum ) that the eigenvectors

    ofand(expressed as kets in 🧲 the total basis ) are

    The spin raising and lowering

    operators acting on these eigenvectors give

    S ^ ± | s , 🧲 m s ⟩ = ℏ s ( s + 1 ) − m s ( m

    s ± 1 ) 🧲 | s , m s ± 1 ⟩ , {\displaystyle {\hat {S}}_{\pm }|s,m_{s}\rangle =\hbar {\sqrt

    {s(s+1)-m_{s}(m_{s}\pm 1)}}|s,m_{s}\pm 1\rangle ,}

    S ^ 🧲 ± = S ^ x ± i S ^ y

    {\displaystyle {\hat {S}}_{\pm }={\hat {S}}_{x}\pm i{\hat {S}}_{y}}

    : 166

    where

    But

    unlike orbital 🧲 angular momentum, the eigenvectors are not spherical harmonics. They are

    not functions of θ and φ. There is also no 🧲 reason to exclude half-integer values of s

    and m s .

    All quantum-mechanical particles possess an intrinsic spin s {\displaystyle

    s} 🧲 (though this value may be equal to zero). The projection of the spin s

    {\displaystyle s} on any axis is 🧲 quantized in units of the reduced Planck constant,

    such that the state function of the particle is, say, not ψ 🧲 = ψ ( r ) {\displaystyle

    \psi =\psi (\mathbf {r} )} , but ψ = ψ ( r , s 🧲 z ) {\displaystyle \psi =\psi (\mathbf

    {r} ,s_{z})} , where s z {\displaystyle s_{z}} can take only the values of 🧲 the

    following discrete set:

    s z ∈ { − s ℏ , − ( s − 1 ) ℏ , … 🧲 , + ( s − 1 ) ℏ , + s ℏ } .

    {\displaystyle s_{z}\in \{-s\hbar ,-(s-1)\hbar ,\dots ,+(s-1)\hbar 🧲 ,+s\hbar \}.}

    One

    distinguishes bosons (integer spin) and fermions (half-integer spin). The total angular

    momentum conserved in interaction processes is then 🧲 the sum of the orbital angular

    momentum and the spin.

    Pauli matrices [ edit ]

    The quantum-mechanical operators

    associated with spin- 1/2 🧲 observables are

    S ^ = ℏ 2 σ , {\displaystyle {\hat {\mathbf

    {S} }}={\frac {\hbar }{2}}{\boldsymbol {\sigma }},}

    S x = ℏ 🧲 2 σ x , S y = ℏ 2 σ y , S z

    = ℏ 2 σ z . 🧲 {\displaystyle S_{x}={\frac {\hbar }{2}}\sigma _{x},\quad S_{y}={\frac

    {\hbar }{2}}\sigma _{y},\quad S_{z}={\frac {\hbar }{2}}\sigma _{z}.}

    where in Cartesian

    components

    For the special case of 🧲 spin- 1/2 particles, σ x , σ y and σ z are the three

    Pauli matrices:

    σ x = ( 0 🧲 1 1 0 ) , σ y = ( 0 − i i 0 ) , σ z = ( 🧲 1 0 0 − 1 ) .

    {\displaystyle \sigma _{x}={\begin{pmatrix}0&1\\1&0\end{pmatrix}},\quad \sigma

    _{y}={\begin{pmatrix}0&-i\\i&0\end{pmatrix}},\quad \sigma

    _{z}={\begin{pmatrix}1&0\\0&-1\end{pmatrix}}.}

    Pauli exclusion principle [ edit ]

    For

    systems 🧲 of N identical particles this is related to the Pauli exclusion principle,

    which states that its wavefunction ψ ( r 🧲 1 , σ 1 , … , r N , σ N ) {\displaystyle \psi

    (\mathbf {r} _{1},\sigma _{1},\dots ,\mathbf 🧲 {r} _{N},\sigma _{N})} must change upon

    interchanges of any two of the N particles as

    ψ ( … , r i 🧲 , σ i , … , r j , σ j , … ) =

    ( − 1 ) 2 🧲 s ψ ( … , r j , σ j , … , r i , σ i , … 🧲 ) . {\displaystyle \psi (\dots

    ,\mathbf {r} _{i},\sigma _{i},\dots ,\mathbf {r} _{j},\sigma _{j},\dots )=(-1)^{2s}\psi

    (\dots ,\mathbf {r} _{j},\sigma _{j},\dots ,\mathbf 🧲 {r} _{i},\sigma _{i},\dots

    ).}

    Thus, for bosons the prefactor (−1)2s will reduce to +1, for fermions to −1. In

    quantum mechanics 🧲 all particles are either bosons or fermions. In some speculative

    relativistic quantum field theories "supersymmetric" particles also exist, where linear

    🧲 combinations of bosonic and fermionic components appear. In two dimensions, the

    prefactor (−1)2s can be replaced by any complex number 🧲 of magnitude 1 such as in the

    anyon.

    The above permutation postulate for N-particle state functions has most

    important consequences in 🧲 daily life, e.g. the periodic table of the chemical

    elements.

    Rotations [ edit ]

    As described above, quantum mechanics states that

    components 🧲 of angular momentum measured along any direction can only take a number of

    discrete values. The most convenient quantum-mechanical description 🧲 of particle's spin

    is therefore with a set of complex numbers corresponding to amplitudes of finding a

    given value of 🧲 projection of its intrinsic angular momentum on a given axis. For

    instance, for a spin- 1/2 particle, we would need 🧲 two numbers a ±1/2 , giving

    amplitudes of finding it with projection of angular momentum equal to + ħ/2 and 🧲 − ħ/2,

    satisfying the requirement

    | a + 1 / 2 | 2 + | a − 1 / 2 | 🧲 2 = 1. {\displaystyle

    |a_{+1/2}|^{2}+|a_{-1/2}|^{2}=1.}

    For a generic particle with spin s, we would need 2s

    + 1 such parameters. Since 🧲 these numbers depend on the choice of the axis, they

    transform into each other non-trivially when this axis is rotated. 🧲 It is clear that the

    transformation law must be linear, so we can represent it by associating a matrix with

    🧲 each rotation, and the product of two transformation matrices corresponding to

    rotations A and B must be equal (up to 🧲 phase) to the matrix representing rotation AB.

    Further, rotations preserve the quantum-mechanical inner product, and so should our

    transformation matrices:

    ∑ 🧲 m = − j j a m ∗ b m = ∑ m = − j j ( ∑ n 🧲 = − j j U n m a n )

    ∗ ( ∑ k = − j j U k 🧲 m b k ) , {\displaystyle \sum _{m=-j}^{j}a_{m}^{*}b_{m}=\sum

    _{m=-j}^{j}\left(\sum _{n=-j}^{j}U_{nm}a_{n}\right)^{*}\left(\sum

    _{k=-j}^{j}U_{km}b_{k}\right),}

    ∑ n = − j j ∑ k = − 🧲 j j U n p ∗ U k q = δ p q .

    {\displaystyle \sum _{n=-j}^{j}\sum _{k=-j}^{j}U_{np}^{*}U_{kq}=\delta

    _{pq}.}

    Mathematically speaking, 🧲 these matrices furnish a unitary projective

    representation of the rotation group SO(3). Each such representation corresponds to a

    representation of 🧲 the covering group of SO(3), which is SU(2).[21] There is one

    n-dimensional irreducible representation of SU(2) for each dimension, though 🧲 this

    representation is n-dimensional real for odd n and n-dimensional complex for even n

    (hence of real dimension 2n). For 🧲 a rotation by angle θ in the plane with normal vector

    θ ^ {\textstyle {\hat {\boldsymbol {\theta }}}} ,

    U = 🧲 e − i ℏ θ ⋅ S , {\displaystyle

    U=e^{-{\frac {i}{\hbar }}{\boldsymbol {\theta }}\cdot \mathbf {S} },}

    θ = θ θ 🧲 ^

    {\textstyle {\boldsymbol {\theta }}=\theta {\hat {\boldsymbol {\theta }}}}

    S

    Proof

    Working in the coordinate system where θ ^ = z ^ 🧲 {\textstyle {\hat {\theta }}={\hat

    {z}}} , we would like to show that S x and S y are rotated into 🧲 each other by the angle

    θ. Starting with S x . Using units where ħ = 1: S x → 🧲 U † S x U = e i θ S z S x e − i θ

    S z = 🧲 S x + ( i θ ) [ S z , S x ] + ( 1 2 ! ) 🧲 ( i θ ) 2 [ S z , [ S z , S x ] ] + ( 1 🧲 3

    ! ) ( i θ ) 3 [ S z , [ S z , [ S z , 🧲 S x ] ] ] + ⋯ {\displaystyle

    {\begin{aligned}S_{x}\rightarrow U^{\dagger }S_{x}U&=e^{i\theta S_{z}}S_{x}e^{-i\theta

    S_{z}}\\&=S_{x}+(i\theta )\left[S_{z},S_{x}\right]+\left({\frac {1}{2!}}\right)(i\theta

    )^{2}\left[S_{z},\left[S_{z},S_{x}\right]\right]+\left({\frac {1}{3!}}\right)(i\theta

    )^{3}\left[S_{z},\left[S_{z},\left[S_{z},S_{x}\right]\right]\right]+\cdots

    \end{aligned}}} Using 🧲 the spin operator commutation relations, we see that the

    commutators evaluate to i S y for the odd terms in 🧲 the series, and to S x for all of

    the even terms. Thus: U † S x U = S 🧲 x [ 1 − θ 2 2 ! + ⋯ ] − S y [ θ − θ 3 3 🧲 ! ⋯ ] = S x

    cos ⁡ θ − S y sin ⁡ θ , {\displaystyle {\begin{aligned}U^{\dagger

    }S_{x}U&=S_{x}\left[1-{\frac {\theta 🧲 ^{2}}{2!}}+\cdots \right]-S_{y}\left[\theta

    -{\frac {\theta ^{3}}{3!}}\cdots \right]\\&=S_{x}\cos \theta -S_{y}\sin \theta

    ,\end{aligned}}} s )[22] : 164 as expected. Note that since we 🧲 only relied on the spin

    operator commutation relations, this proof holds for any dimension (i.e., for any

    principal spin quantum 🧲 number

    where, andis the vector of spin operators

    A generic

    rotation in 3-dimensional space can be built by compounding operators of this 🧲 type

    using Euler angles:

    R ( α , β , γ ) = e − i α S x e − 🧲 i β S y e − i γ S z .

    {\displaystyle {\mathcal {R}}(\alpha ,\beta ,\gamma )=e^{-i\alpha S_{x}}e^{-i\beta

    S_{y}}e^{-i\gamma S_{z}}.}

    An 🧲 irreducible representation of this group of operators is

    furnished by the Wigner D-matrix:

    D m ′ m s ( α , 🧲 β , γ ) ≡ ⟨ s m ′ | R ( α , β , γ ) |

    s 🧲 m ⟩ = e − i m ′ α d m ′ m s ( β ) e − i 🧲 m γ , {\displaystyle D_{m'm}^{s}(\alpha

    ,\beta ,\gamma )\equiv \langle sm'|{\mathcal {R}}(\alpha ,\beta ,\gamma )|sm\rangle

    =e^{-im'\alpha }d_{m'm}^{s}(\beta )e^{-im\gamma },}

    d m ′ 🧲 m s ( β ) = ⟨ s m ′ | e − i β

    s y | s m 🧲 ⟩ {\displaystyle d_{m'm}^{s}(\beta )=\langle sm'|e^{-i\beta s_{y}}|sm\rangle

    }

    γ = 2π

    α = β = 0

    z

    D m ′ m s ( 0 , 🧲 0 , 2 π ) = d m ′ m s ( 0 ) e − i m 2 π 🧲 = δ m ′ m

    ( − 1 ) 2 m . {\displaystyle D_{m'm}^{s}(0,0,2\pi )=d_{m'm}^{s}(0)e^{-im2\pi }=\delta

    _{m'm}(-1)^{2m}.}

    whereis Wigner's small d-matrix 🧲 . Note that forand; i.e., a full

    rotation about theaxis, the Wigner D-matrix elements become

    Recalling that a generic

    spin state 🧲 can be written as a superposition of states with definite m, we see that if

    s is an integer, the 🧲 values of m are all integers, and this matrix corresponds to the

    identity operator. However, if s is a half-integer, 🧲 the values of m are also all

    half-integers, giving (−1)2m = −1 for all m, and hence upon rotation by 🧲 2π the state

    picks up a minus sign. This fact is a crucial element of the proof of the

    spin–statistics 🧲 theorem.

    Lorentz transformations [ edit ]

    We could try the same

    approach to determine the behavior of spin under general Lorentz transformations, 🧲 but

    we would immediately discover a major obstacle. Unlike SO(3), the group of Lorentz

    transformations SO(3,1) is non-compact and therefore 🧲 does not have any faithful,

    unitary, finite-dimensional representations.

    In case of spin- 1/2 particles, it is

    possible to find a construction 🧲 that includes both a finite-dimensional representation

    and a scalar product that is preserved by this representation. We associate a

    4-component 🧲 Dirac spinor ψ with each particle. These spinors transform under Lorentz

    transformations according to the law

    ψ ′ = exp ⁡ 🧲 ( 1 8 ω μ ν [ γ μ , γ ν ] ) ψ ,

    {\displaystyle \psi '=\exp {\left({\tfrac 🧲 {1}{8}}\omega _{\mu

    u }[\gamma _{\mu },\gamma

    _{

    u }]\right)}\psi ,}

    γ ν

    ω μν

    ⟨ ψ | ϕ ⟩ = ψ ¯ ϕ = ψ 🧲 † γ 0 ϕ {\displaystyle \langle

    \psi |\phi \rangle ={\bar {\psi }}\phi =\psi ^{\dagger }\gamma _{0}\phi }

    Measurement

    of spin along 🧲 the x , y , or z axes [ edit ]

    whereare gamma matrices , andis an

    antisymmetric 4 × 4 🧲 matrix parametrizing the transformation. It can be shown that the

    scalar productis preserved. It is not, however, positive-definite, so the

    🧲 representation is not unitary.

    Each of the (Hermitian) Pauli matrices of spin- 1/2

    particles has two eigenvalues, +1 and −1. The 🧲 corresponding normalized eigenvectors

    are

    ψ x + = | 1 2 , + 1 2 ⟩ x = 1 2 ( 🧲 1 1 ) , ψ x − = | 1 2 , − 1 2 ⟩ x = 1 2 🧲 ( 1 − 1 )

    , ψ y + = | 1 2 , + 1 2 ⟩ y = 🧲 1 2 ( 1 i ) , ψ y − = | 1 2 , − 1 2 ⟩ y 🧲 = 1 2 ( 1 − i ) ,

    ψ z + = | 1 2 , + 1 2 🧲 ⟩ z = ( 1 0 ) , ψ z − = | 1 2 , − 1 2 ⟩ 🧲 z = ( 0 1 ) .

    {\displaystyle {\begin{array}{lclc}\psi _{x+}=\left|{\frac {1}{2}},{\frac

    {+1}{2}}\right\rangle _{x}=\displaystyle {\frac {1}{\sqrt

    {2}}}\!\!\!\!\!&{\begin{pmatrix}{1}\\{1}\end{pmatrix}},&\psi _{x-}=\left|{\frac

    {1}{2}},{\frac {-1}{2}}\right\rangle _{x}=\displaystyle 🧲 {\frac {1}{\sqrt

    {2}}}\!\!\!\!\!&{\begin{pmatrix}{1}\\{-1}\end{pmatrix}},\\\psi _{y+}=\left|{\frac

    {1}{2}},{\frac {+1}{2}}\right\rangle _{y}=\displaystyle {\frac {1}{\sqrt

    {2}}}\!\!\!\!\!&{\begin{pmatrix}{1}\\{i}\end{pmatrix}},&\psi _{y-}=\left|{\frac

    {1}{2}},{\frac {-1}{2}}\right\rangle _{y}=\displaystyle {\frac {1}{\sqrt

    {2}}}\!\!\!\!\!&{\begin{pmatrix}{1}\\{-i}\end{pmatrix}},\\\psi _{z+}=\left|{\frac

    {1}{2}},{\frac {+1}{2}}\right\rangle 🧲 _{z}=&{\begin{pmatrix}1\\0\end{pmatrix}},&\psi

    _{z-}=\left|{\frac {1}{2}},{\frac {-1}{2}}\right\rangle

    _{z}=&{\begin{pmatrix}0\\1\end{pmatrix}}.\end{array}}}

    (Because any eigenvector

    multiplied by a constant is still an eigenvector, there is ambiguity about the 🧲 overall

    sign. In this article, the convention is chosen to make the first element imaginary and

    negative if there is 🧲 a sign ambiguity. The present convention is used by software such

    as SymPy; while many physics textbooks, such as Sakurai 🧲 and Griffiths, prefer to make

    it real and positive.)

    By the postulates of quantum mechanics, an experiment designed

    to measure the 🧲 electron spin on the x, y, or z axis can only yield an eigenvalue of the

    corresponding spin operator (S 🧲 x , S y or S z ) on that axis, i.e. ħ/2 or – ħ/2. The

    quantum state of 🧲 a particle (with respect to spin), can be represented by a

    two-component spinor:

    ψ = ( a + b i c 🧲 + d i ) . {\displaystyle \psi

    ={\begin{pmatrix}a+bi\\c+di\end{pmatrix}}.}

    When the spin of this particle is measured

    with respect to a given 🧲 axis (in this example, the x axis), the probability that its

    spin will be measured as ħ/2 is just | 🧲 ⟨ ψ x + | ψ ⟩ | 2 {\displaystyle {\big |}\langle

    \psi _{x+}|\psi \rangle {\big |}^{2}} . Correspondingly, the 🧲 probability that its spin

    will be measured as – ħ/2 is just | ⟨ ψ x − | ψ ⟩ 🧲 | 2 {\displaystyle {\big |}\langle

    \psi _{x-}|\psi \rangle {\big |}^{2}} . Following the measurement, the spin state of

    the particle 🧲 collapses into the corresponding eigenstate. As a result, if the

    particle's spin along a given axis has been measured to 🧲 have a given eigenvalue, all

    measurements will yield the same eigenvalue (since | ⟨ ψ x + | ψ x 🧲 + ⟩ | 2 = 1

    {\displaystyle {\big |}\langle \psi _{x+}|\psi _{x+}\rangle {\big |}^{2}=1} , etc.),

    provided that no measurements 🧲 of the spin are made along other axes.

    Measurement of

    spin along an arbitrary axis [ edit ]

    The operator to measure 🧲 spin along an arbitrary

    axis direction is easily obtained from the Pauli spin matrices. Let u = (u x , 🧲 u y , u

    z ) be an arbitrary unit vector. Then the operator for spin in this direction is

    🧲 simply

    S u = ℏ 2 ( u x σ x + u y σ y + u z σ z 🧲 ) . {\displaystyle S_{u}={\frac {\hbar

    }{2}}(u_{x}\sigma _{x}+u_{y}\sigma _{y}+u_{z}\sigma _{z}).}

    The operator S u has

    eigenvalues of ± ħ/2, just like the 🧲 usual spin matrices. This method of finding the

    operator for spin in an arbitrary direction generalizes to higher spin states, 🧲 one

    takes the dot product of the direction with a vector of the three operators for the

    three x-, y-, 🧲 z-axis directions.

    A normalized spinor for spin- 1/2 in the (u x , u y ,

    u z ) direction (which 🧲 works for all spin states except spin down, where it will give

    0/0) is

    1 2 + 2 u z ( 🧲 1 + u z u x + i u y ) . {\displaystyle {\frac {1}{\sqrt

    {2+2u_{z}}}}{\begin{pmatrix}1+u_{z}\\u_{x}+iu_{y}\end{pmatrix}}.}

    The above spinor is

    obtained 🧲 in the usual way by diagonalizing the σ u matrix and finding the eigenstates

    corresponding to the eigenvalues. In quantum 🧲 mechanics, vectors are termed "normalized"

    when multiplied by a normalizing factor, which results in the vector having a length of

    🧲 unity.

    Compatibility of spin measurements [ edit ]

    Since the Pauli matrices do not

    commute, measurements of spin along the different axes 🧲 are incompatible. This means

    that if, for example, we know the spin along the x axis, and we then measure 🧲 the spin

    along the y axis, we have invalidated our previous knowledge of the x axis spin. This

    can be 🧲 seen from the property of the eigenvectors (i.e. eigenstates) of the Pauli

    matrices that

    | ⟨ ψ x ± | ψ 🧲 y ± ⟩ | 2 = | ⟨ ψ x ± | ψ z ± ⟩ | 2 = | 🧲 ⟨ ψ y ± | ψ z ± ⟩ |

    2 = 1 2 . {\displaystyle {\big |}\langle \psi _{x\pm 🧲 }|\psi _{y\pm }\rangle {\big

    |}^{2}={\big |}\langle \psi _{x\pm }|\psi _{z\pm }\rangle {\big |}^{2}={\big |}\langle

    \psi _{y\pm }|\psi _{z\pm }\rangle {\big 🧲 |}^{2}={\tfrac {1}{2}}.}

    So when physicists

    measure the spin of a particle along the x axis as, for example, ħ/2, the particle's

    🧲 spin state collapses into the eigenstate | ψ x + ⟩ {\displaystyle |\psi _{x+}\rangle }

    . When we then subsequently 🧲 measure the particle's spin along the y axis, the spin

    state will now collapse into either | ψ y + 🧲 ⟩ {\displaystyle |\psi _{y+}\rangle } or |

    ψ y − ⟩ {\displaystyle |\psi _{y-}\rangle } , each with probability 1/2. 🧲 Let us say, in

    our example, that we measure − ħ/2. When we now return to measure the particle's spin

    🧲 along the x axis again, the probabilities that we will measure ħ/2 or − ħ/2 are each

    1/2 (i.e. they 🧲 are | ⟨ ψ x + | ψ y − ⟩ | 2 {\displaystyle {\big |}\langle \psi

    _{x+}|\psi _{y-}\rangle {\big 🧲 |}^{2}} and | ⟨ ψ x − | ψ y − ⟩ | 2 {\displaystyle {\big

    |}\langle \psi _{x-}|\psi _{y-}\rangle 🧲 {\big |}^{2}} respectively). This implies that

    the original measurement of the spin along the x axis is no longer valid, 🧲 since the

    spin along the x axis will now be measured to have either eigenvalue with equal

    probability.

    Higher spins [ 🧲 edit ]

    The spin- 1/2 operator S = ħ/2σ forms the

    fundamental representation of SU(2). By taking Kronecker products of this

    🧲 representation with itself repeatedly, one may construct all higher irreducible

    representations. That is, the resulting spin operators for higher-spin systems 🧲 in three

    spatial dimensions can be calculated for arbitrarily large s using this spin operator

    and ladder operators. For example, 🧲 taking the Kronecker product of two spin- 1/2 yields

    a four-dimensional representation, which is separable into a 3-dimensional spin-1

    (triplet 🧲 states) and a 1-dimensional spin-0 representation (singlet state).

    The

    resulting irreducible representations yield the following spin matrices and eigenvalues

    in the 🧲 z-basis:

    For spin 1 they are S x = ℏ 2 ( 0 1 0 1 0 1 0 1 0 🧲 ) , | 1 , + 1 ⟩ x = 1

    2 ( 1 2 1 ) , | 1 🧲 , 0 ⟩ x = 1 2 ( − 1 0 1 ) , | 1 , − 1 ⟩ 🧲 x = 1 2 ( 1 − 2 1 ) S y = ℏ 2

    ( 0 − i 0 🧲 i 0 − i 0 i 0 ) , | 1 , + 1 ⟩ y = 1 2 ( 🧲 − 1 − i 2 1 ) , | 1 , 0 ⟩ y = 1 2 ( 1

    0 🧲 1 ) , | 1 , − 1 ⟩ y = 1 2 ( − 1 i 2 1 ) 🧲 S z = ℏ ( 1 0 0 0 0 0 0 0 − 1 ) , | 1 , 🧲 + 1 ⟩

    z = ( 1 0 0 ) , | 1 , 0 ⟩ z = ( 0 🧲 1 0 ) , | 1 , − 1 ⟩ z = ( 0 0 1 ) {\displaystyle

    {\begin{aligned}S_{x}&={\frac {\hbar 🧲 }{\sqrt

    {2}}}{\begin{pmatrix}0&1&0\\1&0&1\\0&1&0\end{pmatrix}},&\left|1,+1\right\rangle

    _{x}&={\frac {1}{2}}{\begin{pmatrix}1\\{\sqrt

    {2}}\\1\end{pmatrix}},&\left|1,0\right\rangle _{x}&={\frac {1}{\sqrt

    {2}}}{\begin{pmatrix}-1\\0\\1\end{pmatrix}},&\left|1,-1\right\rangle _{x}&={\frac

    {1}{2}}{\begin{pmatrix}1\\{-{\sqrt {2}}}\\1\end{pmatrix}}\\S_{y}&={\frac {\hbar }{\sqrt

    {2}}}{\begin{pmatrix}0&-i&0\\i&0&-i\\0&i&0\end{pmatrix}},&\left|1,+1\right\rangle

    _{y}&={\frac {1}{2}}{\begin{pmatrix}-1\\-i{\sqrt

    {2}}\\1\end{pmatrix}},&\left|1,0\right\rangle _{y}&={\frac {1}{\sqrt

    {2}}}{\begin{pmatrix}1\\0\\1\end{pmatrix}},&\left|1,-1\right\rangle 🧲 _{y}&={\frac

    {1}{2}}{\begin{pmatrix}-1\\i{\sqrt {2}}\\1\end{pmatrix}}\\S_{z}&=\hbar

    {\begin{pmatrix}1&0&0\\0&0&0\\0&0&-1\end{pmatrix}},&\left|1,+1\right\rangle

    _{z}&={\begin{pmatrix}1\\0\\0\end{pmatrix}},&\left|1,0\right\rangle

    _{z}&={\begin{pmatrix}0\\1\\0\end{pmatrix}},&\left|1,-1\right\rangle

    _{z}&={\begin{pmatrix}0\\0\\1\end{pmatrix}}\\\end{aligned}}} For spin 3 / 2 they are S

    x = ℏ 2 ( 🧲 0 3 0 0 3 0 2 0 0 2 0 3 0 0 3 0 ) , | 3 🧲 2 , + 3 2 ⟩ x = 1 2 2 ( 1 3 3 1 ) , |

    3 🧲 2 , + 1 2 ⟩ x = 1 2 2 ( − 3 − 1 1 3 ) , 🧲 | 3 2 , − 1 2 ⟩ x = 1 2 2 ( 3 − 1 − 1 3 🧲 ) , |

    3 2 , − 3 2 ⟩ x = 1 2 2 ( − 1 3 − 🧲 3 1 ) S y = ℏ 2 ( 0 − i 3 0 0 i 3 0 − 2 🧲 i 0 0 2 i 0 −

    i 3 0 0 i 3 0 ) , | 3 2 , 🧲 + 3 2 ⟩ y = 1 2 2 ( i − 3 − i 3 1 ) , | 🧲 3 2 , + 1 2 ⟩ y = 1 2

    2 ( − i 3 1 − i 3 🧲 ) , | 3 2 , − 1 2 ⟩ y = 1 2 2 ( i 3 1 i 🧲 3 ) , | 3 2 , − 3 2 ⟩ y = 1 2

    2 ( − i − 🧲 3 i 3 1 ) S z = ℏ 2 ( 3 0 0 0 0 1 0 0 0 🧲 0 − 1 0 0 0 0 − 3 ) , | 3 2 , + 3 2 ⟩

    z 🧲 = ( 1 0 0 0 ) , | 3 2 , + 1 2 ⟩ z = ( 0 🧲 1 0 0 ) , | 3 2 , − 1 2 ⟩ z = ( 0 0 1 0 🧲 ) , |

    3 2 , − 3 2 ⟩ z = ( 0 0 0 1 ) {\displaystyle {\begin{array}{lclc}S_{x}={\frac 🧲 {\hbar

    }{2}}{\begin{pmatrix}0&{\sqrt {3}}&0&0\\{\sqrt {3}}&0&2&0\\0&2&0&{\sqrt

    {3}}\\0&0&{\sqrt {3}}&0\end{pmatrix}},\!\!\!&\left|{\frac {3}{2}},{\frac

    {+3}{2}}\right\rangle _{x}=\!\!\!&{\frac {1}{2{\sqrt {2}}}}{\begin{pmatrix}1\\{\sqrt

    {3}}\\{\sqrt {3}}\\1\end{pmatrix}},\!\!\!&\left|{\frac {3}{2}},{\frac

    {+1}{2}}\right\rangle _{x}=\!\!\!&{\frac {1}{2{\sqrt {2}}}}{\begin{pmatrix}{-{\sqrt

    {3}}}\\-1\\1\\{\sqrt {3}}\end{pmatrix}},\!\!\!&\left|{\frac 🧲 {3}{2}},{\frac

    {-1}{2}}\right\rangle _{x}=\!\!\!&{\frac {1}{2{\sqrt {2}}}}{\begin{pmatrix}{\sqrt

    {3}}\\-1\\-1\\{\sqrt {3}}\end{pmatrix}},\!\!\!&\left|{\frac {3}{2}},{\frac

    {-3}{2}}\right\rangle _{x}=\!\!\!&{\frac {1}{2{\sqrt {2}}}}{\begin{pmatrix}-1\\{\sqrt

    {3}}\\{-{\sqrt {3}}}\\1\end{pmatrix}}\\S_{y}={\frac {\hbar

    }{2}}{\begin{pmatrix}0&-i{\sqrt {3}}&0&0\\i{\sqrt {3}}&0&-2i&0\\0&2i&0&-i{\sqrt

    {3}}\\0&0&i{\sqrt {3}}&0\end{pmatrix}},\!\!\!&\left|{\frac 🧲 {3}{2}},{\frac

    {+3}{2}}\right\rangle _{y}=\!\!\!&{\frac {1}{2{\sqrt

    {2}}}}{\begin{pmatrix}{i}\\{-{\sqrt {3}}}\\{-i{\sqrt

    {3}}}\\1\end{pmatrix}},\!\!\!&\left|{\frac {3}{2}},{\frac {+1}{2}}\right\rangle

    _{y}=\!\!\!&{\frac {1}{2{\sqrt {2}}}}{\begin{pmatrix}{-i{\sqrt {3}}}\\1\\{-i}\\{\sqrt

    {3}}\end{pmatrix}},\!\!\!&\left|{\frac {3}{2}},{\frac {-1}{2}}\right\rangle

    _{y}=\!\!\!&{\frac {1}{2{\sqrt {2}}}}{\begin{pmatrix}{i{\sqrt {3}}}\\1\\{i}\\{\sqrt

    🧲 {3}}\end{pmatrix}},\!\!\!&\left|{\frac {3}{2}},{\frac {-3}{2}}\right\rangle

    _{y}=\!\!\!&{\frac {1}{2{\sqrt {2}}}}{\begin{pmatrix}{-i}\\{-{\sqrt {3}}}\\{i{\sqrt

    {3}}}\\1\end{pmatrix}}\\S_{z}={\frac {\hbar

    }{2}}{\begin{pmatrix}3&0&0&0\\0&1&0&0\\0&0&-1&0\\0&0&0&-3\end{pmatrix}},\!\!\!&\left|{\

    frac {3}{2}},{\frac {+3}{2}}\right\rangle

    _{z}=\!\!\!&{\begin{pmatrix}1\\0\\0\\0\end{pmatrix}},\!\!\!&\left|{\frac {3}{2}},{\frac

    {+1}{2}}\right\rangle

    _{z}=\!\!\!&{\begin{pmatrix}0\\1\\0\\0\end{pmatrix}},\!\!\!&\left|{\frac {3}{2}},{\frac

    {-1}{2}}\right\rangle

    _{z}=\!\!\!&{\begin{pmatrix}0\\0\\1\\0\end{pmatrix}},\!\!\!&\left|{\frac {3}{2}},{\frac

    🧲 {-3}{2}}\right\rangle

    _{z}=\!\!\!&{\begin{pmatrix}0\\0\\0\\1\end{pmatrix}}\\\end{array}}} For spin 5 / 2 they

    are S x = ℏ 2 ( 0 5 0 0 0 🧲 0 5 0 2 2 0 0 0 0 2 2 0 3 0 0 0 0 3 0 2 🧲 2 0 0 0 0 2 2 0 5 0 0

    0 0 5 0 ) , S y = 🧲 ℏ 2 ( 0 − i 5 0 0 0 0 i 5 0 − 2 i 2 0 0 🧲 0 0 2 i 2 0 − 3 i 0 0 0 0 3 i

    0 − 2 i 2 🧲 0 0 0 0 2 i 2 0 − i 5 0 0 0 0 i 5 0 ) , 🧲 S z = ℏ 2 ( 5 0 0 0 0 0 0 3 0 0 0 0 0

    0 🧲 1 0 0 0 0 0 0 − 1 0 0 0 0 0 0 − 3 0 0 0 🧲 0 0 0 − 5 ) . {\displaystyle

    {\begin{aligned}{\boldsymbol {S}}_{x}&={\frac {\hbar }{2}}{\begin{pmatrix}0&{\sqrt

    {5}}&0&0&0&0\\{\sqrt {5}}&0&2{\sqrt {2}}&0&0&0\\0&2{\sqrt {2}}&0&3&0&0\\0&0&3&0&2{\sqrt

    {2}}&0\\0&0&0&2{\sqrt {2}}&0&{\sqrt {5}}\\0&0&0&0&{\sqrt

    {5}}&0\end{pmatrix}},\\{\boldsymbol 🧲 {S}}_{y}&={\frac {\hbar

    }{2}}{\begin{pmatrix}0&-i{\sqrt {5}}&0&0&0&0\\i{\sqrt {5}}&0&-2i{\sqrt

    {2}}&0&0&0\\0&2i{\sqrt {2}}&0&-3i&0&0\\0&0&3i&0&-2i{\sqrt {2}}&0\\0&0&0&2i{\sqrt

    {2}}&0&-i{\sqrt {5}}\\0&0&0&0&i{\sqrt {5}}&0\end{pmatrix}},\\{\boldsymbol

    {S}}_{z}&={\frac {\hbar

    }{2}}{\begin{pmatrix}5&0&0&0&0&0\\0&3&0&0&0&0\\0&0&1&0&0&0\\0&0&0&-1&0&0\\0&0&0&0&-3&0\

    \0&0&0&0&0&-5\end{pmatrix}}.\end{aligned}}} The generalization of these matrices for

    🧲 arbitrary spin s is ( S x ) a b = ℏ 2 ( δ a , b + 1 🧲 + δ a + 1 , b ) ( s + 1 ) ( a + b −

    1 ) 🧲 − a b , ( S y ) a b = i ℏ 2 ( δ a , b + 🧲 1 − δ a + 1 , b ) ( s + 1 ) ( a + b − 1 🧲 ) −

    a b , ( S z ) a b = ℏ ( s + 1 − a ) 🧲 δ a , b = ℏ ( s + 1 − b ) δ a , b , {\displaystyle

    {\begin{aligned}\left(S_{x}\right)_{ab}&={\frac 🧲 {\hbar }{2}}\left(\delta

    _{a,b+1}+\delta _{a+1,b}\right){\sqrt

    {(s+1)(a+b-1)-ab}},\\\left(S_{y}\right)_{ab}&={\frac {i\hbar }{2}}\left(\delta

    _{a,b+1}-\delta _{a+1,b}\right){\sqrt

    {(s+1)(a+b-1)-ab}},\\\left(S_{z}\right)_{ab}&=\hbar (s+1-a)\delta _{a,b}=\hbar

    (s+1-b)\delta _{a,b},\end{aligned}}} a , b {\displaystyle a,b} 1 🧲 ≤ a ≤ 2 s + 1 , 1 ≤ b

    ≤ 2 s + 1. {\displaystyle 1\leq a\leq 2s+1,\quad 🧲 1\leq b\leq 2s+1.}

    Also useful in the

    quantum mechanics of multiparticle systems, the general Pauli group G n is defined to

    🧲 consist of all n-fold tensor products of Pauli matrices.

    The analog formula of Euler's

    formula in terms of the Pauli matrices

    R 🧲 ^ ( θ , n ^ ) = e i θ 2 n ^ ⋅ σ = I cos ⁡ 🧲 θ 2

    + i ( n ^ ⋅ σ ) sin ⁡ θ 2 {\displaystyle {\hat {R}}(\theta ,{\hat {\mathbf {n}

    🧲 }})=e^{i{\frac {\theta }{2}}{\hat {\mathbf {n} }}\cdot {\boldsymbol {\sigma }}}=I\cos

    {\frac {\theta }{2}}+i\left({\hat {\mathbf {n} }}\cdot {\boldsymbol {\sigma

    }}\right)\sin {\frac {\theta 🧲 }{2}}}

    Parity [ edit ]

    for higher spins is tractable, but

    less simple.

    In tables of the spin quantum number s for nuclei 🧲 or particles, the spin

    is often followed by a "+" or "−". This refers to the parity with "+" for 🧲 even parity

    (wave function unchanged by spatial inversion) and "−" for odd parity (wave function

    negated by spatial inversion). For 🧲 example, see the isotopes of bismuth, in which the

    list of isotopes includes the column nuclear spin and parity. For 🧲 Bi-209, the

    longest-lived isotope, the entry 9/2– means that the nuclear spin is 9/2 and the parity

    is odd.

    Applications [ 🧲 edit ]

    Spin has important theoretical implications and practical

    applications. Well-established direct applications of spin include:

    Electron spin plays

    an important role 🧲 in magnetism, with applications for instance in computer memories.

    The manipulation of nuclear spin by radio-frequency waves (nuclear magnetic resonance)

    🧲 is important in chemical spectroscopy and medical imaging.

    Spin–orbit coupling leads to

    the fine structure of atomic spectra, which is used 🧲 in atomic clocks and in the modern

    definition of the second. Precise measurements of the g-factor of the electron have

    🧲 played an important role in the development and verification of quantum

    electrodynamics. Photon spin is associated with the polarization of 🧲 light (photon

    polarization).

    An emerging application of spin is as a binary information carrier in

    spin transistors. The original concept, proposed 🧲 in 1990, is known as Datta–Das spin

    transistor.[24] Electronics based on spin transistors are referred to as spintronics.

    The manipulation 🧲 of spin in dilute magnetic semiconductor materials, such as

    metal-doped ZnO or TiO 2 imparts a further degree of freedom 🧲 and has the potential to

    facilitate the fabrication of more efficient electronics.[25]

    There are many indirect

    applications and manifestations of spin 🧲 and the associated Pauli exclusion principle,

    starting with the periodic table of chemistry.

    History [ edit ]

    Spin was first

    discovered in 🧲 the context of the emission spectrum of alkali metals. In 1924, Wolfgang

    Pauli introduced what he called a "two-valuedness not 🧲 describable classically"[26]

    associated with the electron in the outermost shell. This allowed him to formulate the

    Pauli exclusion principle, stating 🧲 that no two electrons can have the same quantum

    state in the same quantum system.

    The physical interpretation of Pauli's "degree 🧲 of

    freedom" was initially unknown. Ralph Kronig, one of Landé's assistants, suggested in

    early 1925 that it was produced by 🧲 the self-rotation of the electron. When Pauli heard

    about the idea, he criticized it severely, noting that the electron's hypothetical

    🧲 surface would have to be moving faster than the speed of light in order for it to

    rotate quickly enough 🧲 to produce the necessary angular momentum. This would violate the

    theory of relativity. Largely due to Pauli's criticism, Kronig decided 🧲 not to publish

    his idea.[27]

    In the autumn of 1925, the same thought came to Dutch physicists George

    Uhlenbeck and Samuel 🧲 Goudsmit at Leiden University. Under the advice of Paul Ehrenfest,

    they published their results.[28] It met a favorable response, especially 🧲 after

    Llewellyn Thomas managed to resolve a factor-of-two discrepancy between experimental

    results and Uhlenbeck and Goudsmit's calculations (and Kronig's unpublished 🧲 results).

    This discrepancy was due to the orientation of the electron's tangent frame, in

    addition to its position.

    Mathematically speaking, a 🧲 fiber bundle description is

    needed. The tangent bundle effect is additive and relativistic; that is, it vanishes if

    c goes 🧲 to infinity. It is one half of the value obtained without regard for the

    tangent-space orientation, but with opposite sign. 🧲 Thus the combined effect differs

    from the latter by a factor two (Thomas precession, known to Ludwik Silberstein in

    1914).

    Despite 🧲 his initial objections, Pauli formalized the theory of spin in 1927,

    using the modern theory of quantum mechanics invented by 🧲 Schrödinger and Heisenberg. He

    pioneered the use of Pauli matrices as a representation of the spin operators and

    introduced a 🧲 two-component spinor wave-function. Uhlenbeck and Goudsmit treated spin as

    arising from classical rotation, while Pauli emphasized, that spin is a 🧲 non-classical

    and intrinsic property.[29]

    Pauli's theory of spin was non-relativistic. However, in

    1928, Paul Dirac published the Dirac equation, which described 🧲 the relativistic

    electron. In the Dirac equation, a four-component spinor (known as a "Dirac spinor")

    was used for the electron 🧲 wave-function. Relativistic spin explained gyromagnetic

    anomaly, which was (in retrospect) first observed by Samuel Jackson Barnett in 1914

    (see Einstein–de 🧲 Haas effect). In 1940, Pauli proved the spin–statistics theorem, which

    states that fermions have half-integer spin, and bosons have integer 🧲 spin.

    In

    retrospect, the first direct experimental evidence of the electron spin was the

    Stern–Gerlach experiment of 1922. However, the correct 🧲 explanation of this experiment

    was only given in 1927.[30]

    See also [ edit ]

    References [ edit ]

    Further reading [

    edit ]

  • roleta do dinheiro whatsapp
  • dicas de apostas no sportingbet
  • guia de apostas
  • melhor casino online
  • estoril sol casinos online

  • artigos relacionados

    1. sulbets
    2. mobile aposta ganha bet
    3. 2 bets bola
    4. bwin que es
    5. blaze jogo de azar
    6. sign up virgin bet