Search This Blog

Wednesday, September 3, 2014

Alfred,Lord Tennyson : Ulysses

Alfred,Lord Tennyson : Ulysses


    It little profits that an idle king1,
    By this still hearth, among these barren crags,
    Matched with an agèd wife, I mete and dole
    Unequal laws unto a savage race,
    That hoard, and sleep, and feed, and know not me.

    I cannot rest from travel: I will drink
    Life to the lees: all times I have enjoyed
    Greatly, have suffered greatly, both with those
    That loved me, and alone; on shore, and when
    Through scudding drifts the rainy Hyades2
    Vexed the dim sea: I am become a name;
    For always roaming with a hungry heart
    Much have I seen and known; cities of men
    And manners, climates, councils, governments,
    Myself not least, but honoured of them all;
    And drunk delight of battle with my peers,
    Far on the ringing plains of windy Troy3.
    I am a part of all that I have met;
    Yet all experience is an arch wherethrough
    Gleams that untravelled world, whose margin fades
    For ever and for ever when I move.
    How dull it is to pause, to make an end,
    To rust unburnished, not to shine in use!
    As though to breathe were life. Life piled on life
    Were all too little, and of one to me
    Little remains: but every hour is saved
    From that eternal silence, something more,
    A bringer of new things; and vile it were
    For some three suns to store and hoard myself,
    And this grey spirit yearning in desire
    To follow knowledge like a sinking star,
    Beyond the utmost bound of human thought.

        This my son, mine own Telemachus,
    To whom I leave the sceptre and the isle—
    Well-loved of me, discerning to fulfil
    This labour, by slow prudence to make mild
    A rugged people, and through soft degrees
    Subdue them to the useful and the good.
    Most blameless is he, centred in the sphere
    Of common duties, decent not to fail
    In offices of tenderness, and pay
    Meet adoration to my household gods,
    When I am gone. He works his work, I mine.

        There lies the port; the vessel puffs her sail:
    There gloom the dark broad seas. My mariners,
    Souls that have toiled, and wrought, and thought
        with me—
    That ever with a frolic welcome took
    The thunder and the sunshine, and opposed
    Free hearts, free foreheads—you and I are old;
    Old age hath yet his honour and his toil;
    Death closes all: but something ere the end,
    Some work of noble note, may yet be done,
    Not unbecoming men that strove with Gods.
    The lights begin to twinkle from the rocks:
    The long day wanes: the slow moon climbs: the deep
    Moans round with many voices. Come, my friends,
    'Tis not too late to seek a newer world.
    Push off, and sitting well in order smite
    The sounding furrows; for my purpose holds
    To sail beyond the sunset, and the baths
    Of all the western stars, until I die.
    It may be that the gulfs will wash us down:
    It may be we shall touch the Happy Isles4,
    And see the great Achilles5, whom we knew
    Though much is taken, much abides; and though
    We are not now that strength which in old days
    Moved earth and heaven; that which we are, we are;
    One equal temper of heroic hearts,
    Made weak by time and fate, but strong in will
    To strive, to seek, to find, and not to yield.

Alfred,Lord Tennyson (1809-1892)    1833

U.S. wind energy industry is doing quite well too (It’s not all about solar)

U.S. wind energy industry is doing quite well too (It’s not all about solar)


1
Aug 22, 2014

Original link:  http://www.smartgridnews.com/artman/publish/Technologies_DG_Renewables/U-S-wind-energy-industry-is-doing-quite-well-too-It-s-not-all-about-solar-6712.html/#.VAdaIWNdx5-
 
The U.S. wind energy industry generally gets overlooked as the solar industry grabs the biggest share of the spotlight in the renewables world. Solar is getting cheaper almost by the day. Many expect rooftop solar coupled with energy storage to be the next big thing, and the net metering debate continues.
 
But two new Energy Department reports say U.S. wind energy is quite healthy, thank you. DOE considers wind power to be a “key component” of the strategy to cut carbon pollution, diversify the  country’s energy economy and more. Scan the press release below for an overview and click on the links for more detail.
 
Energy Department Reports Highlight Strength of U.S. Wind Energy Industry
 
smart grid, modern grid, smart grid technology, renewable energy, wind power, U.S. wind power, Department of Energy







Washington, D.C. -- The U.S. continues to be a global leader in wind energy, ranking second in installed capacity in the world, according to two reports released today by the Department of Energy.
Wind power is a key component of the nation’s all-of-the-above strategy to reduce carbon pollution, diversify our energy economy, and bring innovative technologies on line. With increasing wind energy generation and decreasing prices of wind energy technologies, the U.S. wind energy market remains strong and the U.S. is moving closer to doubling renewable electricity generation from energy resources like wind power yet again by 2020.
 
“As a readily expandable, domestic source of clean, renewable energy, wind power is paving the way to a low-carbon future that protects our air and water while providing affordable, renewable electricity to American families and businesses,” said Energy Secretary Ernest Moniz. “However, the continued success of the U.S. wind industry highlights the importance of policies like the Production Tax Credit that provide a solid framework for America to lead the world in clean energy innovation while also keeping wind manufacturing and jobs in the U.S.”


Wind Technologies Market Report


After modest growth in 2013, total installed wind power capacity in the United States now stands at 61 gigawatts (GW), which meets nearly 4.5 percent of electricity demand in an average year, according to the 2013 Wind Technologies Market Report, released today by the Energy Department and its Lawrence Berkeley National Laboratory. The report also found that wind energy prices – particularly in the Interior region of the United States–are at an all-time low, with utilities selecting wind as a cost-saving option.

With utility-scale turbines installed in more than 39 states and territories, the success of the U.S. wind industry has had a ripple effect on the American economy, spurring more than $500 million in exports and supporting jobs related to development, siting, manufacturing, transportation and other industries.
 

Distributed Wind Market Report


In total, U.S. turbines in distributed applications, which accounted for more than 80 percent of all wind turbines installed in the U.S. last year, reached a cumulative installed capacity of more than 842 MW–enough to power 120,000 average American homes–according to the 2013 Distributed Wind Market Report, also released today by the Energy Department and its Pacific Northwest National Laboratory. This capacity is supplied by roughly 72,000 turbines across all 50 states, Puerto Rico, and the U.S. Virgin Islands. In fact, a total of 14 states, including Iowa, Nevada and California, among others, now each have more than 10 MW of distributed wind capacity.
 
Compared to traditional, centralized power plants, distributed wind energy installations supply power directly to the local grid near homes, farms, businesses and communities. Turbines used in these applications can range in size from a few hundred watts to multi-megawatts, and can help power remote, off-grid homes and farms as well as local schools and manufacturing facilities.
 
For more information on these two new reports – including infographics, video and updated interactive map – visit www.energy.gov/windreport.

Chameleons and holograms: Dark energy hunt gets weird

Chameleons and holograms: Dark energy hunt gets weird

  • 03 September 2014 by Hal Hodson, Chicago
  • Original link:  http://www.newscientist.com/article/mg22329852.400-chameleons-and-holograms-dark-energy-hunt-gets-weird.html?utm_source=NSNS&utm_medium=SOC&utm_campaign=facebookgoogletwitter&cmpid=SOC|NSNS|2012-GLOBAL-facebookgoogletwitter#.VAdaA2Ndx58
Cosmologists have revealed intruiging new ways to probe the mystery of whether dark energy exists and how it might be accelerating the universe’s growth

A LIGHT in the darkness can come from unexpected places. Unusual experiments for probing dark energy seem set to revolutionise our understanding of this mysterious force.

In Chicago last week, the world's largest meeting of cosmologists debated two of the forces that could push the universe apart: inflation, the proposed period of exponential expansion that the universe went through immediately after the big bang; and dark energy, the present-day force thought to be responsible for pushing the cosmos outward at an ever increasing rate.

The announcement in March that gravitational waves had been seen should essentially prove that inflation happened. But the results are on ice. The BICEP2 telescope team, which did the work, may have underestimated the impact of galactic dust on the signal. If real, the pattern of the waves they saw in the cosmic microwave background – the earliest light emitted in the universe – is the fingerprint of the universe's rapid expansion.

Astronomers and cosmologists at the International Conference on Particle Physics and Cosmology (COSMO) duked it out over how their models for the universe would be affected in two futures: one in which the results hold, the other in which dust blows them away.

"Everyone wants BICEP2 to be right," Will Kinney of the University at Buffalo, New York, told a packed auditorium. "Because if it is, we are going to be doing incredibly precise physics on the inflationary model within the foreseeable future. And it's going to be really cool."


For now, physicists will have to wait. New data from the Planck satellite, which could clear up BICEP2's problems, is not due to be released until November, but rumours swirled at COSMO that at least one paper based on Planck data within BICEP2's field of view will be published any day.

In the calm before that storm, much of the attention is on dark energy, and some big steps have been made. Dark energy is a theoretical necessity, exerting a repulsive force that explains how the speed of our universe's expansion is accelerating. But we know almost nothing about it.

COSMO saw novel work for exploring dark energy in Earthly laboratories (see "Chameleon screen"), an experiment that could show that the expansion is a fundamental property of space-time itself (see "A quantum of space-time"), and new constraints on the most devastating model of dark energy, which would see our universe tear itself apart, atom for atom (see "Phantom menace").



"Everybody and their mother is constraining dark energy," says Dragan Huterer at the University of Michigan in Ann Arbor. "That's the name of the game: you're measuring the expansion history of the universe."

Chameleon screen

Many physicists think dark energy is shoving the universe apart by countering gravity. If that's true, why have experiments never seen it? One hypothesis is that the force adapts to its environment and is only active in a near vacuum, while the dense matter of the solar system "screens" it from view.

Now Clare Burrage at the University of Nottingham, UK, and her colleagues have begun work on a laboratory test to find these screened "chameleon" forces.

If dark energy's effects can be felt only across a space as empty as the universe, Burrage says, then the same effect may show up in a vacuum chamber containing only a small ball of stable material and a cloud of a mere 1000 atoms. The team plans to use a laser to move the atoms 1 millimetre across the chamber.

As they travel, the atoms will feel the gravitational pull of Earth as well as that of the confounding ball of material, and the experiment will measure which forces are affecting them. If the disguising force of the ball is acting on the atoms, they should take a slightly different path, which will be visible in their final quantum states.

The team has not yet done the experiment, but has requested a special laser from a quantum GPS system from the UK Ministry of Defence, which should arrive in the next few months.

Finding chameleon-like effects won't necessarily mean they've found dark energy, says Adrienne Erickcek of the University of North Carolina at Chapel Hill. But it will show that screening mechanisms are a plausible explanation for our failure to measure the effects of dark energy in the local universe.

"This is very exciting," Erickcek says. "I had always assumed that the chameleon force would be screened no matter what, but they showed really convincingly that it need not be. It's amazing."

A quantum of space-time

An experiment in a shed in the suburbs of Chicago could show that dark energy is simply an emergent property of space-time, much as fluid dynamics emerges from how water molecules interact.

The goal of the Holometer experiment is to find the fundamental units of space and time. These would be a hundred billion billion times smaller than a proton. Like matter and energy at the quantum scale, these bits of space-time would act more like waves than particles.

"The theory is that space is made of waves instead of points, that everything is a little jittery, and never sits still," says Craig Hogan at the University of Chicago, who runs the experiment. The Holometer is designed to measure this "jitter".

It directs two powerful laser beams through tubes 40 metres long. The lasers measure the positions of mirrors along their paths at two points in time. If space-time is smooth and shows no quantum behaviour, then the mirrors should remain perfectly still. But if both lasers measure an identical, small difference in the mirrors' position over time, after all other effects are ruled out, that could mean the mirrors are being jiggled by fluctuations in the fabric of space.

Taking this idea a step further, Hogan says the quantum states of space-time and matter could be entangled, so you can't measure one without affecting the other.

Our best current theories describe space-time in terms of geometry, and matter in terms of quantum fields, but struggle to unite the two. If the Holometer sees something, Hogan says, it could point to a way of unifying them. At the tiny scales at which the two properties are connected, the geometry of space-time alone should force the universe to expand.

Hogan told the COSMO meeting that initial results show that Holometer can measure quantum fluctuations, if they are there, and could collect enough data for an answer within a year.

Phantom menace

Most models of dark energy hold that the amount of it remains constant. But about 10 years ago, cosmologists realised that if the total density of dark energy is increasing, we could be headed for a nightmare scenario – the "big rip". As space-time expands faster and faster, matter will be torn apart, starting with galaxy clusters and ending with atomic nuclei. Cosmologists called it "phantom" energy.

To find out if this could be true, Dragan Huterer at the University of Michigan in Ann Arbor turned to type Ia supernovae. These stellar explosions are all of the same brightness, so they act as cosmic yardsticks for measuring distances. The first evidence that the universe's expansion is accelerating came from studies of type Ia supernovae in the late 1990s.

If supernovae accelerated away from each other more slowly in the past than they do now, then dark energy's density may be increasing and we could be in trouble. "If you even move a millimetre off the ledge, you fall into the abyss," Huterer says.

Huterer and colleague Daniel Shafer have compiled data from recent supernova surveys and found that, depending on which surveys you use, there could be slight evidence that the dark energy density has been increasing over the past 2 billion years, but it's not statistically significant yet (Physical Review D, doi.org/vf9).

Phantom energy is an underdog theory, but the consequences are so dramatic that it's worth testing, Huterer says. The weakness of the evidence is balanced by the fact that the implications are huge, he says. "We will have to completely revise even our current thinking of dark energy if phantom is really at work."

Lepton

Lepton

From Wikipedia, the free encyclopedia

Lepton
Beta Negative Decay.svg
Leptons are involved in several processes such as beta decay.
Composition Elementary particle
Statistics Fermionic
Generation 1st, 2nd, 3rd
Interactions Electromagnetism, Gravitation, Weak
Symbol l
Antiparticle Antilepton (l)
Types 6 (electron, electron neutrino, muon, muon neutrino, tau, tau neutrino)
Electric charge +1 e, 0 e, −1 e
Color charge No
Spin 12

A lepton is an elementary, spin-12 particle that does not undergo strong interactions, but is subject to the Pauli exclusion principle.[1] The best known of all leptons is the electron, which governs nearly all of chemistry as it is found in atoms and is directly tied to all chemical properties. Two main classes of leptons exist: charged leptons (also known as the electron-like leptons), and neutral leptons (better known as neutrinos). Charged leptons can combine with other particles to form various composite particles such as atoms and positronium, while neutrinos rarely interact with anything, and are consequently rarely observed.

There are six types of leptons, known as flavours, forming three generations.[2] The first generation is the electronic leptons, comprising the electron (e) and electron neutrino (ν
e
); the second is the muonic leptons, comprising the muon (μ) and muon neutrino (ν
μ
); and the third is the tauonic leptons, comprising the tau (τ) and the tau neutrino (ν
τ
). Electrons have the least mass of all the charged leptons. The heavier muons and taus will rapidly change into electrons through a process of particle decay: the transformation from a higher mass state to a lower mass state. Thus electrons are stable and the most common charged lepton in the universe, whereas muons and taus can only be produced in high energy collisions (such as those involving cosmic rays and those carried out in particle accelerators).

Leptons have various intrinsic properties, including electric charge, spin, and mass. Unlike quarks however, leptons are not subject to the strong interaction, but they are subject to the other three fundamental interactions: gravitation, electromagnetism (excluding neutrinos, which are electrically neutral), and the weak interaction. For every lepton flavor there is a corresponding type of antiparticle, known as antilepton, that differs from the lepton only in that some of its properties have equal magnitude but opposite sign. However, according to certain theories, neutrinos may be their own antiparticle, but it is not currently known whether this is the case or not.

The first charged lepton, the electron, was theorized in the mid-19th century by several scientists[3][4][5] and was discovered in 1897 by J. J. Thomson.[6] The next lepton to be observed was the muon, discovered by Carl D. Anderson in 1936, but it was erroneously classified as a meson at the time.[7] After investigation, it was realized that the muon did not have the expected properties of a meson, but rather behaved like an electron, only with higher mass. It took until 1947 for the concept of "leptons" as a family of particle to be proposed.[8] The first neutrino, the electron neutrino, was proposed by Wolfgang Pauli in 1930 to explain certain characteristics of beta decay.[8] It was first observed in the Cowan–Reines neutrino experiment conducted by Clyde Cowan and Frederick Reines in 1956.[8][9] The muon neutrino was discovered in 1962 by Leon M. Lederman, Melvin Schwartz and Jack Steinberger,[10] and the tau discovered between 1974 and 1977 by Martin Lewis Perl and his colleagues from the Stanford Linear Accelerator Center and Lawrence Berkeley National Laboratory.[11] The tau neutrino remained elusive until July 2000, when the DONUT collaboration from Fermilab announced its discovery.[12][13]

Leptons are an important part of the Standard Model. Electrons are one of the components of atoms, alongside protons and neutrons. Exotic atoms with muons and taus instead of electrons can also be synthesized, as well as lepton–antilepton particles such as positronium.

Etymology

The name lepton comes from the Greek λεπτόν (leptón), neuter of λεπτός (leptós), "fine, small, thin";[14] the earliest attested form of the word is the Mycenaean Greek 𐀩𐀡𐀵, re-po-to, written in Linear B syllabic script.[15] Lepton was first used by physicist Léon Rosenfeld in 1948:[16]
Following a suggestion of Prof. C. Møller, I adopt — as a pendant to "nucleon" — the denomination "lepton" (from λεπτός, small, thin, delicate) to denote a particle of small mass.
The etymology incorrectly implies that all the leptons are of small mass. When Rosenfeld named them, the only known leptons were electrons and muons, which are in fact of small mass — the mass of an electron (0.511 MeV/c2)[17] and the mass of a muon (with a value of 105.7 MeV/c2)[18] are fractions of the mass of the "heavy" proton (938.3 MeV/c2).[19] However, the mass of the tau (discovered in the mid 1970s) (1777 MeV/c2)[20] is nearly twice that of the proton, and about 3,500 times that of the electron.

History

A muon transmutes into a muon neutrino by emitting a W boson. The W boson subsequently decays into an electron and an electron antineutrino.

The first lepton identified was the electron, discovered by J.J. Thomson and his team of British physicists in 1897.[21][22] Then in 1930 Wolfgang Pauli postulated the electron neutrino to preserve conservation of energy, conservation of momentum, and conservation of angular momentum in beta decay.[23] Pauli theorized that an undetected particle was carrying away the difference between the energy, momentum, and angular momentum of the initial and observed final particles. The electron neutrino was simply called the neutrino, as it was not yet known that neutrinos came in different flavours (or different "generations").

Nearly 40 years after the discovery of the electron, the muon was discovered by Carl D. Anderson in 1936. Due to its mass, it was initially categorized as a meson rather than a lepton.[24] It later became clear that the muon was much more similar to the electron than to mesons, as muons do not undergo the strong interaction, and thus the muon was reclassified: electrons, muons, and the (electron) neutrino were grouped into a new group of particles – the leptons. In 1962 Leon M. Lederman, Melvin Schwartz and Jack Steinberger showed that more than one type of neutrino exists by first detecting interactions of the muon neutrino, which earned them the 1988 Nobel Prize, although by then the different flavours of neutrino had already been theorized.[25]

The tau was first detected in a series of experiments between 1974 and 1977 by Martin Lewis Perl with his colleagues at the SLAC LBL group.[26] Like the electron and the muon, it too was expected to have an associated neutrino. The first evidence for tau neutrinos came from the observation of "missing" energy and momentum in tau decay, analogous to the "missing" energy and momentum in beta decay leading to the discovery of the electron neutrino. The first detection of tau neutrino interactions was announced in 2000 by the DONUT collaboration at Fermilab, making it the latest particle of the Standard Model to have been directly observed,[27] apart from the Higgs boson, which probably has been discovered in 2012.

Although all present data is consistent with three generations of leptons, some particle physicists are searching for a fourth generation. The current lower limit on the mass of such a fourth charged lepton is 100.8 GeV/c2,[28] while its associated neutrino would have a mass of at least 45.0 GeV/c2.[29]

Properties

Spin and chirality

Left-handed and right-handed helicities

Leptons are spin-12 particles. The spin-statistics theorem thus implies that they are fermions and thus that they are subject to the Pauli exclusion principle; no two leptons of the same species can be in exactly the same state at the same time. Furthermore, it means that a lepton can have only two possible spin states, namely up or down.

A closely related property is chirality, which in turn is closely related to a more easily visualized property called helicity. The helicity of a particle is the direction of its spin relative to its momentum; particles with spin in the same direction as their momentum are called right-handed and otherwise they are called left-handed. When a particle is mass-less, the direction of its momentum relative to its spin is frame independent, while for massive particles it is possible to 'overtake' the particle by a Lorentz transformation flipping the helicity. Chirality is a technical property (defined through the transformation behaviour under the Poincaré group) that agrees with helicity for (approximately) massless particles and is still well defined for massive particles.

In many quantum field theories—such as quantum electrodynamics and quantum chromodynamics—left and right-handed fermions are identical. However in the Standard Model left-handed and right-handed fermions are treated asymmetrically. Only left-handed fermions participate in the weak interaction, while there are no right-handed neutrinos. This is an example of parity violation. In the literature left-handed fields are often denoted by a capital L subscript (e.g. eL) and right-handed fields are denoted by a capital R subscript.

Electromagnetic interaction

Lepton-photon interaction

One of the most prominent properties of leptons is their electric charge, Q. The electric charge determines the strength of their electromagnetic interactions. It determines the strength of the electric field generated by the particle (see Coulomb's law) and how strongly the particle reacts to an external electric or magnetic field (see Lorentz force). Each generation contains one lepton with Q = −1 (conventionally the charge of a particle is expressed in units of the elementary charge) and one lepton with zero electric charge. The lepton with electric charge is commonly simply referred to as a 'charged positive lepton' while the neutral lepton is called a neutrino. For example the first generation consists of the electron e with a negative electric charge and the electrically neutral electron neutrino ν
e
.

In the language of quantum field theory the electromagnetic interaction of the charged leptons is expressed by the fact that the particles interact with the quantum of the electromagnetic field, the photon. The Feynman diagram of the electron-photon interaction is shown on the right.

Because leptons possess an intrinsic rotation in the form of their spin, charged leptons generate a magnetic field. The size of their magnetic dipole moment μ is given by,
\mu = g \frac{ Q e \hbar}{4 m},
where m is the mass of the lepton and g is the so-called g-factor for the lepton. First order approximation quantum mechanics predicts that the g-factor is 2 for all leptons. However, higher order quantum effects caused by loops in Feynman diagrams introduce corrections to this value.
These corrections, referred to as the anomalous magnetic dipole moment, are very sensitive to the details of a quantum field theory model and thus provide the opportunity for precision tests of the standard model. The theoretical and measured values for the electron anomalous magnetic dipole moment are within agreement within eight significant figures.[30]

Weak Interaction

Lepton-interaction-vertex-evW.svg Lepton-interaction-vertex-pvW.svg Lepton-interaction-vertex-eeZ.svg
The weak interactions of the first generation leptons.
In the Standard Model the left-handed charged lepton and the left-handed neutrino are arranged in doublet (ν
e
L, eL)
that transforms in the spinor representation (T = 12) of the weak isospin SU(2) gauge symmetry. This means that these particles are eigenstates of the isospin projection T3 with eigenvalues 12 and −12 respectively. In the meantime, the right-handed charged lepton transforms as a weak isospin scalar (T = 0) and thus does not participate in the weak interaction, while there is no right-handed neutrino at all.

The Higgs mechanism recombines the gauge fields of the weak isospin SU(2) and the weak hypercharge U(1) symmetries to three massive vector bosons (W+, W, Z0) mediating the weak interaction, and one massless vector boson, the photon, responsible for the electromagnetic interaction. The electric charge Q can be calculated from the isospin projection T3 and weak hypercharge YW through the Gell-Mann–Nishijima formula,
Q = T3 + YW/2
To recover the observed electric charges for all particles the left-handed weak isospin doublet (ν
e
L, eL)
must thus have YW = −1, while the right-handed isospin scalar e
R
must have YW = −2. The interaction of the leptons with the massive weak interaction vector bosons is shown in the figure on the left.

Mass

In the Standard Model each lepton starts out with no intrinsic mass. The charged leptons (i.e. the electron, muon, and tau) obtain an effective mass through interaction with the Higgs field, but the neutrinos remain massless. For technical reasons the masslessness of the neutrinos implies that there is no mixing of the different generations of charged leptons as there is for quarks. This is in close agreement with current experimental observations.[31]

However, it is known from experiments – most prominently from observed neutrino oscillations[32] – that neutrinos do in fact have some very small mass, probably less than eV/c2.[33] This implies the existence of physics beyond the Standard Model. The currently most favoured extension is the so-called seesaw mechanism, which would explain both why the left-handed neutrinos are so light compared to the corresponding charged leptons, and why we have not yet seen any right-handed neutrinos.

Leptonic numbers

The members of each generation's weak isospin doublet are assigned leptonic numbers that are conserved under the Standard Model.[34] Electrons and electron neutrinos have an electronic number of Le = 1, while muons and muon neutrinos have a muonic number of Lμ = 1, while tau particles and tau neutrinos have a tauonic number of Lτ = 1. The antileptons have their respective generation's leptonic numbers of −1.
Conservation of the leptonic numbers means that the number of leptons of the same type remains the same, when particles interact. This implies that leptons and antileptons must be created in pairs of a single generation. For example, the following processes are allowed under conservation of leptonic numbers:
Each generation forms a weak isospin doublet.
e + e+γ + γ,
τ + τ+Z0 + Z0,
but not these:
γe + μ+,
We + ν
τ
,
Z0μ + τ+.

However, neutrino oscillations are known to violate the conservation of the individual leptonic numbers. Such a violation is considered to be smoking gun evidence for physics beyond the Standard Model. A much stronger conservation law is the conservation of the total number of leptons (L), conserved even in the case of neutrino oscillations, but even it is still violated by a tiny amount by the chiral anomaly.

Universality

The coupling of the leptons to gauge bosons are flavour-independent (i.e., the interactions between leptons and gauge bosons are the same for all leptons).[34] This property is called lepton universality and has been tested in measurements of the tau and muon lifetimes and of Z boson partial decay widths, particularly at the Stanford Linear Collider (SLC) and Large Electron-Positron Collider (LEP) experiments.[35]:241–243[36]:138

The decay rate (Γ) of muons through the process μe + ν
e
+ ν
μ
is approximately given by an expression of the form (see muon decay for more details)[34]
\Gamma \left ( \mu^- \rarr e^- + \bar{\nu_e} +\nu_\mu \right ) = K_1G_F^2m_\mu^5,
where K1 is some constant, and GF is the Fermi coupling constant. The decay rate of tau particles through the process τe + ν
e
+ ν
τ
is given by an expression of the same form[34]
\Gamma \left ( \tau^- \rarr e^- + \bar{\nu_e} +\nu_\tau \right ) = K_2G_F^2m_\tau^5,
where K2 is some constant. Muon–Tauon universality implies that K1 = K2. On the other hand, electron–muon universality implies[34]
\Gamma \left ( \tau^- \rarr e^- + \bar{\nu_e} +\nu_\tau \right ) = \Gamma \left ( \tau^- \rarr \mu^- + \bar{\nu_\mu} +\nu_\tau \right ).
This explains why the branching ratios for the electronic mode (17.85%) and muonic (17.36%) mode of tau decay are equal (within error).[20]

Universality also accounts for the ratio of muon and tau lifetimes. The lifetime of a lepton (τl) is related to the decay rate by[34]
\tau_l=\frac{B \left ( l^- \rarr e^- + \bar{\nu_e} +\nu_l \right )}{\Gamma \left ( l^- \rarr e^- + \bar{\nu_e} +\nu_l \right )},
where B(x → y) and Γ(x → y) denotes the branching ratios and the resonance width of the process x → y.

The ratio of tau and muon lifetime is thus given by[34]
\frac{\tau_\tau}{\tau_\mu} = \frac{B \left ( \tau^- \rarr e^- + \bar{\nu_e} +\nu_\tau \right )}{B \left ( \mu^- \rarr e^- + \bar{\nu_e} +\nu_\mu \right )}\left (\frac{m_\mu}{m_\tau}\right )^5.

Using the values of the 2008 Review of Particle Physics for the branching ratios of muons[18] and tau[20] yields a lifetime ratio of ~1.29×10−7, comparable to the measured lifetime ratio of ~1.32×10−7. The difference is due to K1 and K2 not actually being constants; they depend on the mass of leptons.

Table of leptons

Properties of leptons
Particle/Antiparticle Name Symbol Q (e) S Le Lμ Lτ Mass (MeV/c2) Lifetime (s) Common decay
Electron / Positron[17] e/e+ −1/+1 12 +1/−1 0 0 0.510998910(13) Stable Stable
Muon / Antimuon[18] μ/μ+ −1/+1 12 0 +1/−1 0 105.6583668(38) 2.197019(21)×10−6 e + ν
e
+ ν
μ
Tau / Antitau[20] τ/τ+ −1/+1 12 0 0 +1/−1 1776.84(17) 2.906(10)×10−13 See τ decay modes
Electron neutrino / Electron antineutrino[33] ν
e
/ν
e
0 12 +1/−1 0 0 < 2.2×10−6[37] Unknown
Muon neutrino / Muon antineutrino[33] ν
μ
/ν
μ
0 12 0 +1/−1 0 < 0.17[37] Unknown
Tau neutrino / Tau antineutrino[33] ν
τ
/ν
τ
0 12 0 0 +1/−1 < 15.5[37] Unknown

Quark

Quark

From Wikipedia, the free encyclopedia

Quark
Three colored balls (symbolizing quarks) connected pairwise by springs (symbolizing gluons), all inside a gray circle (symbolizing a proton). The colors of the balls are red, green, and blue, to parallel each quark's color charge. The red and blue balls are labeled "u" (for "up" quark) and the green one is labeled "d" (for "down" quark).
A proton, composed of two up quarks and one down quark. (The color assignment of individual quarks is not important, only that all three colors be present.)
Composition Elementary particle
Statistics Fermionic
Generation 1st, 2nd, 3rd
Interactions Electromagnetism, gravitation, strong, weak
Symbol q
Antiparticle Antiquark (q)
Theorized Murray Gell-Mann (1964)
George Zweig (1964)
Discovered SLAC (~1968)
Types 6 (up, down, strange, charm, bottom, and top)
Electric charge +23 e, −13 e
Color charge Yes
Spin 12
Baryon number 13

A quark (/ˈkwɔrk/ or /ˈkwɑrk/) is an elementary particle and a fundamental constituent of matter. Quarks combine to form composite particles called hadrons, the most stable of which are protons and neutrons, the components of atomic nuclei.[1] Due to a phenomenon known as color confinement, quarks are never directly observed or found in isolation; they can be found only within hadrons, such as baryons (of which protons and neutrons are examples), and mesons.[2][3] For this reason, much of what is known about quarks has been drawn from observations of the hadrons themselves.

There are six types of quarks, known as flavors: up, down, strange, charm, bottom, and top.[4] Up and down quarks have the lowest masses of all quarks. The heavier quarks rapidly change into up and down quarks through a process of particle decay: the transformation from a higher mass state to a lower mass state. Because of this, up and down quarks are generally stable and the most common in the universe, whereas strange, charm, bottom, and top quarks can only be produced in high energy collisions (such as those involving cosmic rays and in particle accelerators).

Quarks have various intrinsic properties, including electric charge, mass, color charge and spin. Quarks are the only elementary particles in the Standard Model of particle physics to experience all four fundamental interactions, also known as fundamental forces (electromagnetism, gravitation, strong interaction, and weak interaction), as well as the only known particles whose electric charges are not integer multiples of the elementary charge. For every quark flavor there is a corresponding type of antiparticle, known as an antiquark, that differs from the quark only in that some of its properties have equal magnitude but opposite sign.

The quark model was independently proposed by physicists Murray Gell-Mann and George Zweig in 1964.[5] Quarks were introduced as parts of an ordering scheme for hadrons, and there was little evidence for their physical existence until deep inelastic scattering experiments at the Stanford Linear Accelerator Center in 1968.[6][7] Accelerator experiments have provided evidence for all six flavors. The top quark was the last to be discovered at Fermilab in 1995.[5]

Classification

See also: Standard Model
A four-by-four table of particles. Columns are three generations of matter (fermions) and one of forces (bosons). In the first three columns, two rows contain quarks and two leptons. The top two rows' columns contain up (u) and down (d) quarks, charm (c) and strange (s) quarks, top (t) and bottom (b) quarks, and photon (γ) and gluon (g), respectively. The bottom two rows' columns contain electron neutrino (ν sub e) and electron (e), muon neutrino (ν sub μ) and muon (μ), and tau neutrino (ν sub τ) and tau (τ), and Z sup 0 and W sup ± weak force. Mass, charge, and spin are listed for each particle.
Six of the particles in the Standard Model are quarks (shown in purple). Each of the first three columns forms a generation of matter.

The Standard Model is the theoretical framework describing all the currently known elementary particles. This model contains six flavors of quarks (q), named up (u), down (d), strange (s), charm (c), bottom (b), and top (t).[4] Antiparticles of quarks are called antiquarks, and are denoted by a bar over the symbol for the corresponding quark, such as u for an up antiquark. As with antimatter in general, antiquarks have the same mass, mean lifetime, and spin as their respective quarks, but the electric charge and other charges have the opposite sign.[8]

Quarks are spin-12 particles, implying that they are fermions according to the spin-statistics theorem. They are subject to the Pauli exclusion principle, which states that no two identical fermions can simultaneously occupy the same quantum state. This is in contrast to bosons (particles with integer spin), any number of which can be in the same state.[9] Unlike leptons, quarks possess color charge, which causes them to engage in the strong interaction. The resulting attraction between different quarks causes the formation of composite particles known as hadrons (see "Strong interaction and color charge" below).

The quarks which determine the quantum numbers of hadrons are called valence quarks; apart from these, any hadron may contain an indefinite number of virtual (or sea) quarks, antiquarks, and gluons which do not influence its quantum numbers.[10] There are two families of hadrons: baryons, with three valence quarks, and mesons, with a valence quark and an antiquark.[11] The most common baryons are the proton and the neutron, the building blocks of the atomic nucleus.[12] A great number of hadrons are known (see list of baryons and list of mesons), most of them differentiated by their quark content and the properties these constituent quarks confer. The existence of "exotic" hadrons with more valence quarks, such as tetraquarks (qqqq) and pentaquarks (qqqqq), has been conjectured[13] but not proven.[nb 1][13][14]

Elementary fermions are grouped into three generations, each comprising two leptons and two quarks. The first generation includes up and down quarks, the second strange and charm quarks, and the third bottom and top quarks. All searches for a fourth generation of quarks and other elementary fermions have failed,[15] and there is strong indirect evidence that no more than three generations exist.[nb 2][16] Particles in higher generations generally have greater mass and less stability, causing them to decay into lower-generation particles by means of weak interactions. Only first-generation (up and down) quarks occur commonly in nature. Heavier quarks can only be created in high-energy collisions (such as in those involving cosmic rays), and decay quickly; however, they are thought to have been present during the first fractions of a second after the Big Bang, when the universe was in an extremely hot and dense phase (the quark epoch). Studies of heavier quarks are conducted in artificially created conditions, such as in particle accelerators.[17]

Having electric charge, mass, color charge, and flavor, quarks are the only known elementary particles that engage in all four fundamental interactions of contemporary physics: electromagnetism, gravitation, strong interaction, and weak interaction.[12] Gravitation is too weak to be relevant to individual particle interactions except at extremes of energy (Planck energy) and distance scales (Planck distance). However, since no successful quantum theory of gravity exists, gravitation is not described by the Standard Model.

See the table of properties below for a more complete overview of the six quark flavors' properties.

History

Half-length portrait of a white-haired man in his seventies talking. A painting of Beethoven is in the background.
Murray Gell-Mann at TED in 2007. Gell-Mann and George Zweig proposed the quark model in 1964.

The quark model was independently proposed by physicists Murray Gell-Mann[18] and George Zweig[19][20] in 1964.[5] The proposal came shortly after Gell-Mann's 1961 formulation of a particle classification system known as the Eightfold Way—or, in more technical terms, SU(3) flavor symmetry.[21] Physicist Yuval Ne'eman had independently developed a scheme similar to the Eightfold Way in the same year.[22][23]

At the time of the quark theory's inception, the "particle zoo" included, amongst other particles, a multitude of hadrons. Gell-Mann and Zweig posited that they were not elementary particles, but were instead composed of combinations of quarks and antiquarks. Their model involved three flavors of quarks, up, down, and strange, to which they ascribed properties such as spin and electric charge.[18][19][20] The initial reaction of the physics community to the proposal was mixed. There was particular contention about whether the quark was a physical entity or a mere abstraction used to explain concepts that were not fully understood at the time.[24]

In less than a year, extensions to the Gell-Mann–Zweig model were proposed. Sheldon Lee Glashow and James Bjorken predicted the existence of a fourth flavor of quark, which they called charm. The addition was proposed because it allowed for a better description of the weak interaction (the mechanism that allows quarks to decay), equalized the number of known quarks with the number of known leptons, and implied a mass formula that correctly reproduced the masses of the known mesons.[25]

In 1968, deep inelastic scattering experiments at the Stanford Linear Accelerator Center (SLAC) showed that the proton contained much smaller, point-like objects and was therefore not an elementary particle.[6][7][26] Physicists were reluctant to firmly identify these objects with quarks at the time, instead calling them "partons"—a term coined by Richard Feynman.[27][28][29] The objects that were observed at SLAC would later be identified as up and down quarks as the other flavors were discovered.[30] Nevertheless, "parton" remains in use as a collective term for the constituents of hadrons (quarks, antiquarks, and gluons).

The strange quark's existence was indirectly validated by SLAC's scattering experiments: not only was it a necessary component of Gell-Mann and Zweig's three-quark model, but it provided an explanation for the kaon (K) and pion (π) hadrons discovered in cosmic rays in 1947.[31]

In a 1970 paper, Glashow, John Iliopoulos and Luciano Maiani presented further reasoning for the existence of the as-yet undiscovered charm quark.[32][33] The number of supposed quark flavors grew to the current six in 1973, when Makoto Kobayashi and Toshihide Maskawa noted that the experimental observation of CP violation[nb 3][34] could be explained if there were another pair of quarks.
Photo of bubble chamber tracks next to diagram of same tracks. A neutrino (unseen in photo) enters from below and collides with a proton, producing a negatively charged muon, three positively charged pions, and one negatively charged pion, as well as a neutral lambda baryon (unseen in photograph). The lambda baryon then decays into a proton and a negative pion, producing a "V" pattern.
Photograph of the event that led to the discovery of the Σ++
c
baryon
, at the Brookhaven National Laboratory in 1974

Charm quarks were produced almost simultaneously by two teams in November 1974 (see November Revolution)—one at SLAC under Burton Richter, and one at Brookhaven National Laboratory under Samuel Ting. The charm quarks were observed bound with charm antiquarks in mesons. The two parties had assigned the discovered meson two different symbols, J and ψ; thus, it became formally known as the J/ψ meson. The discovery finally convinced the physics community of the quark model's validity.[29]

In the following years a number of suggestions appeared for extending the quark model to six quarks. Of these, the 1975 paper by Haim Harari[35] was the first to coin the terms top and bottom for the additional quarks.[36]

In 1977, the bottom quark was observed by a team at Fermilab led by Leon Lederman.[37][38] This was a strong indicator of the top quark's existence: without the top quark, the bottom quark would have been without a partner. However, it was not until 1995 that the top quark was finally observed, also by the CDF[39] and [40] teams at Fermilab.[5] It had a mass much larger than had been previously expected,[41] almost as large as that of a gold atom.[42]

Etymology

For some time, Gell-Mann was undecided on an actual spelling for the term he intended to coin, until he found the word quark in James Joyce's book Finnegans Wake:
Three quarks for Muster Mark!
Sure he has not got much of a bark
And sure any he has it's all beside the mark.
—James Joyce, Finnegans Wake[43]
Gell-Mann went into further detail regarding the name of the quark in his book The Quark and the Jaguar:[44]
In 1963, when I assigned the name "quark" to the fundamental constituents of the nucleon, I had the sound first, without the spelling, which could have been "kwork". Then, in one of my occasional perusals of Finnegans Wake, by James Joyce, I came across the word "quark" in the phrase "Three quarks for Muster Mark". Since "quark" (meaning, for one thing, the cry of the gull) was clearly intended to rhyme with "Mark", as well as "bark" and other such words, I had to find an excuse to pronounce it as "kwork". But the book represents the dream of a publican named Humphrey Chimpden Earwicker. Words in the text are typically drawn from several sources at once, like the "portmanteau" words in "Through the Looking-Glass". From time to time, phrases occur in the book that are partially determined by calls for drinks at the bar. I argued, therefore, that perhaps one of the multiple sources of the cry "Three quarks for Muster Mark" might be "Three quarts for Mister Mark", in which case the pronunciation "kwork" would not be totally unjustified. In any case, the number three fitted perfectly the way quarks occur in nature.
Zweig preferred the name ace for the particle he had theorized, but Gell-Mann's terminology came to prominence once the quark model had been commonly accepted.[45]

The quark flavors were given their names for a number of reasons. The up and down quarks are named after the up and down components of isospin, which they carry.[46] Strange quarks were given their name because they were discovered to be components of the strange particles discovered in cosmic rays years before the quark model was proposed; these particles were deemed "strange" because they had unusually long lifetimes.[47] Glashow, who coproposed charm quark with Bjorken, is quoted as saying, "We called our construct the 'charmed quark', for we were fascinated and pleased by the symmetry it brought to the subnuclear world."[48] The names "bottom" and "top", coined by Harari, were chosen because they are "logical partners for up and down quarks".[35][36][47] In the past, bottom and top quarks were sometimes referred to as "beauty" and "truth" respectively, but these names have somewhat fallen out of use.[49] While "truth" never did catch on, accelerator complexes devoted to massive production of bottom quarks are sometimes called "beauty factories".[50]

Properties

Electric charge

Quarks have fractional electric charge values – either 13 or 23 times the elementary charge, depending on flavor. Up, charm, and top quarks (collectively referred to as up-type quarks) have a charge of +23, while down, strange, and bottom quarks (down-type quarks) have −13. Antiquarks have the opposite charge to their corresponding quarks; up-type antiquarks have charges of −23 and down-type antiquarks have charges of +13. Since the electric charge of a hadron is the sum of the charges of the constituent quarks, all hadrons have integer charges: the combination of three quarks (baryons), three antiquarks (antibaryons), or a quark and an antiquark (mesons) always results in integer charges.[51] For example, the hadron constituents of atomic nuclei, neutrons and protons, have charges of 0 and +1 respectively; the neutron is composed of two down quarks and one up quark, and the proton of two up quarks and one down quark.[12]

Spin

Spin is an intrinsic property of elementary particles, and its direction is an important degree of freedom. It is sometimes visualized as the rotation of an object around its own axis (hence the name "spin"), though this notion is somewhat misguided at subatomic scales because elementary particles are believed to be point-like.[52]
Spin can be represented by a vector whose length is measured in units of the reduced Planck constant ħ (pronounced "h bar"). For quarks, a measurement of the spin vector component along any axis can only yield the values +ħ/2 or −ħ/2; for this reason quarks are classified as spin-12 particles.[53] The component of spin along a given axis – by convention the z axis – is often denoted by an up arrow ↑ for the value +12 and down arrow ↓ for the value −12, placed after the symbol for flavor. For example, an up quark with a spin of +12 along the z axis is denoted by u↑.[54]

Weak interaction

A tree diagram consisting mostly of straight arrows. A down quark forks into an up quark and a wavy-arrow W[superscript minus] boson, the latter forking into an electron and reversed-arrow electron antineutrino.
Feynman diagram of beta decay with time flowing upwards. The CKM matrix (discussed below) encodes the probability of this and other quark decays.

A quark of one flavor can transform into a quark of another flavor only through the weak interaction, one of the four fundamental interactions in particle physics. By absorbing or emitting a W boson, any up-type quark (up, charm, and top quarks) can change into any down-type quark (down, strange, and bottom quarks) and vice versa. This flavor transformation mechanism causes the radioactive process of beta decay, in which a neutron (n) "splits" into a proton (p), an electron (e) and an electron antineutrino (ν
e
) (see picture). This occurs when one of the down quarks in the neutron (udd) decays into an up quark by emitting a virtual W boson, transforming the neutron into a proton (uud). The W boson then decays into an electron and an electron antineutrino.[55]

  n   p + e + ν
e
(Beta decay, hadron notation)
udd uud + e + ν
e
(Beta decay, quark notation)

Both beta decay and the inverse process of inverse beta decay are routinely used in medical applications such as positron emission tomography (PET) and in experiments involving neutrino detection.
Three balls "u", "c", and "t" noted "up-type quarks" stand above three balls "d", "s", "b" noted "down-type quark". The "u", "c", and "t" balls are vertically aligned with the "d", "s", and b" balls respectively. Colored lines connect the "up-type" and "down-type" quarks, with the darkness of the color indicating the strength of the weak interaction between the two; The lines "d" to "u", "c" to "s", and "t" to "b" are dark; The lines "c" to "d" and "s" to "u" are grayish; and the lines "b" to "u", "b" to "c", "t" to "d", and "t" to "s" are almost white.
The strengths of the weak interactions between the six quarks. The "intensities" of the lines are determined by the elements of the CKM matrix.

While the process of flavor transformation is the same for all quarks, each quark has a preference to transform into the quark of its own generation. The relative tendencies of all flavor transformations are described by a mathematical table, called the Cabibbo–Kobayashi–Maskawa matrix (CKM matrix). Enforcing unitarity, the approximate magnitudes of the entries of the CKM matrix are:[56]
|V_ud| ≅ 0.974; |V_us| ≅ 0.225; |V_ub| ≅ 0.003; |V_cd| ≅ 0.225; |V_cs| ≅ 0.973; |V_cb| ≅ 0.041; |V_td| ≅ 0.009; |V_ts| ≅ 0.040; |V_tb| ≅ 0.999.
where Vij represents the tendency of a quark of flavor i to change into a quark of flavor j (or vice versa).[nb 4]

There exists an equivalent weak interaction matrix for leptons (right side of the W boson on the above beta decay diagram), called the Pontecorvo–Maki–Nakagawa–Sakata matrix (PMNS matrix).[57] Together, the CKM and PMNS matrices describe all flavor transformations, but the links between the two are not yet clear.[58]

Strong interaction and color charge

A green and a magenta ("antigreen") arrow canceling out each other out white, representing a meson; a red, a green, and a blue arrow canceling out to white, representing a baryon; a yellow ("antiblue"), a magenta, and a cyan ("antired") arrow canceling out to white, representing an antibaryon.
All types of hadrons have zero total color charge.
The pattern of strong charges for the three colors of quark, three antiquarks, and eight gluons (with two of zero charge overlapping).

According to QCD, quarks possess a property called color charge. There are three types of color charge, arbitrarily labeled blue, green, and red.[nb 5] Each of them is complemented by an anticolor – antiblue, antigreen, and antired. Every quark carries a color, while every antiquark carries an anticolor.[59]

The system of attraction and repulsion between quarks charged with different combinations of the three colors is called strong interaction, which is mediated by force carrying particles known as gluons; this is discussed at length below. The theory that describes strong interactions is called quantum chromodynamics (QCD). A quark charged with one color value can form a bound system with an antiquark carrying the corresponding anticolor; three (anti)quarks, one of each (anti)color, will similarly be bound together. The result of two attracting quarks will be color neutrality: a quark with color charge ξ plus an antiquark with color charge −ξ will result in a color charge of 0 (or "white" color) and the formation of a meson. Analogous to the additive color model in basic optics, the combination of three quarks or three antiquarks, each with different color charges, will result in the same "white" color charge and the formation of a baryon or antibaryon.[60]

In modern particle physics, gauge symmetries – a kind of symmetry group – relate interactions between particles (see gauge theories). Color SU(3) (commonly abbreviated to SU(3)c) is the gauge symmetry that relates the color charge in quarks and is the defining symmetry for quantum chromodynamics.[61] Just as the laws of physics are independent of which directions in space are designated x, y, and z, and remain unchanged if the coordinate axes are rotated to a new orientation, the physics of quantum chromodynamics is independent of which directions in three-dimensional color space are identified as blue, red, and green. SU(3)c color transformations correspond to "rotations" in color space (which, mathematically speaking, is a complex space). Every quark flavor f, each with subtypes fB, fG, fR corresponding to the quark colors,[62] forms a triplet: a three-component quantum field which transforms under the fundamental representation of SU(3)c.[63] The requirement that SU(3)c should be local – that is, that its transformations be allowed to vary with space and time – determines the properties of the strong interaction, in particular the existence of eight gluon types to act as its force carriers.[61][64]

Mass

Current quark masses for all six flavors in comparison, as balls of proportional volumes. Proton and electron (red) are shown in bottom left corner for scale

Two terms are used in referring to a quark's mass: current quark mass refers to the mass of a quark by itself, while constituent quark mass refers to the current quark mass plus the mass of the gluon particle field surrounding the quark.[65] These masses typically have very different values. Most of a hadron's mass comes from the gluons that bind the constituent quarks together, rather than from the quarks themselves. While gluons are inherently massless, they possess energy – more specifically, quantum chromodynamics binding energy (QCBE) – and it is this that contributes so greatly to the overall mass of the hadron (see mass in special relativity). For example, a proton has a mass of approximately 938 MeV/c2, of which the rest mass of its three valence quarks only contributes about 11 MeV/c2; much of the remainder can be attributed to the gluons' QCBE.[66][67]

The Standard Model posits that elementary particles derive their masses from the Higgs mechanism, which is related to the Higgs boson. Physicists hope that further research into the reasons for the top quark's large mass of ~173 GeV/c2, almost the mass of a gold atom,[66][68] might reveal more about the origin of the mass of quarks and other elementary particles.[69]

Table of properties

The following table summarizes the key properties of the six quarks. Flavor quantum numbers (isospin (I3), charm (C), strangeness (S, not to be confused with spin), topness (T), and bottomness (B′)) are assigned to certain quark flavors, and denote qualities of quark-based systems and hadrons.
The baryon number (B) is +13 for all quarks, as baryons are made of three quarks. For antiquarks, the electric charge (Q) and all flavor quantum numbers (B, I3, C, S, T, and B′) are of opposite sign. Mass and total angular momentum (J; equal to spin for point particles) do not change sign for the antiquarks.

Quark flavor properties[66]
Name Symbol Mass (MeV/c2)* J B Q I3 C S T B′ Antiparticle Antiparticle symbol
First generation
Up u 1.7 to 3.1 12 +13 +23 +12 0 0 0 0 Antiup u
Down d 4.1 to 5.7 12 +13 13 12 0 0 0 0 Antidown d
Second generation
Charm c 1290+50
−110
12 +13 +23 0 +1 0 0 0 Anticharm c
Strange s 100+30
−20
12 +13 13 0 0 −1 0 0 Antistrange s
Third generation
Top t 172900±600 ± 900 12 +13 +23 0 0 0 +1 0 Antitop t
Bottom b 4190+180
−60
12 +13 13 0 0 0 0 −1 Antibottom b
J = total angular momentum, B = baryon number, Q = electric charge, I3 = isospin, C = charm, S = strangeness, T = topness, B′ = bottomness.
* Notation such as 4190+180
−60
denotes measurement uncertainty. In the case of the top quark, the first uncertainty is statistical in nature, and the second is systematic.

Interacting quarks

As described by quantum chromodynamics, the strong interaction between quarks is mediated by gluons, massless vector gauge bosons. Each gluon carries one color charge and one anticolor charge. In the standard framework of particle interactions (part of a more general formulation known as perturbation theory), gluons are constantly exchanged between quarks through a virtual emission and absorption process. When a gluon is transferred between quarks, a color change occurs in both; for example, if a red quark emits a red–antigreen gluon, it becomes green, and if a green quark absorbs a red–antigreen gluon, it becomes red. Therefore, while each quark's color constantly changes, their strong interaction is preserved.[70][71][72]
Since gluons carry color charge, they themselves are able to emit and absorb other gluons. This causes asymptotic freedom: as quarks come closer to each other, the chromodynamic binding force between them weakens.[73] Conversely, as the distance between quarks increases, the binding force strengthens. The color field becomes stressed, much as an elastic band is stressed when stretched, and more gluons of appropriate color are spontaneously created to strengthen the field. Above a certain energy threshold, pairs of quarks and antiquarks are created. These pairs bind with the quarks being separated, causing new hadrons to form. This phenomenon is known as color confinement: quarks never appear in isolation.[71][74] This process of hadronization occurs before quarks, formed in a high energy collision, are able to interact in any other way. The only exception is the top quark, which may decay before it hadronizes.[75]

Sea quarks

Hadrons, along with the valence quarks (q
v
) that contribute to their quantum numbers, contain virtual quark–antiquark (qq) pairs known as sea quarks (q
s
). Sea quarks form when a gluon of the hadron's color field splits; this process also works in reverse in that the annihilation of two sea quarks produces a gluon. The result is a constant flux of gluon splits and creations colloquially known as "the sea".[76] Sea quarks are much less stable than their valence counterparts, and they typically annihilate each other within the interior of the hadron. Despite this, sea quarks can hadronize into baryonic or mesonic particles under certain circumstances.[77]

Other phases of quark matter

Quark–gluon plasma exists at very high temperatures; the hadronic phase exists at lower temperatures and baryonic densities, in particular nuclear matter for relatively low temperatures and intermediate densities; color superconductivity exists at sufficiently low temperatures and high densities.
A qualitative rendering of the phase diagram of quark matter. The precise details of the diagram are the subject of ongoing research.[78][79]

Under sufficiently extreme conditions, quarks may become deconfined and exist as free particles. In the course of asymptotic freedom, the strong interaction becomes weaker at higher temperatures. Eventually, color confinement would be lost and an extremely hot plasma of freely moving quarks and gluons would be formed. This theoretical phase of matter is called quark–gluon plasma.[80] The exact conditions needed to give rise to this state are unknown and have been the subject of a great deal of speculation and experimentation. A recent estimate puts the needed temperature at (1.90±0.02)×1012 Kelvin.[81] While a state of entirely free quarks and gluons has never been achieved (despite numerous attempts by CERN in the 1980s and 1990s),[82] recent experiments at the Relativistic Heavy Ion Collider have yielded evidence for liquid-like quark matter exhibiting "nearly perfect" fluid motion.[83]

The quark–gluon plasma would be characterized by a great increase in the number of heavier quark pairs in relation to the number of up and down quark pairs. It is believed that in the period prior to 10−6 seconds after the Big Bang (the quark epoch), the universe was filled with quark–gluon plasma, as the temperature was too high for hadrons to be stable.[84]

Given sufficiently high baryon densities and relatively low temperatures – possibly comparable to those found in neutron stars – quark matter is expected to degenerate into a Fermi liquid of weakly interacting quarks. This liquid would be characterized by a condensation of colored quark Cooper pairs, thereby breaking the local SU(3)c symmetry. Because quark Cooper pairs harbor color charge, such a phase of quark matter would be color superconductive; that is, color charge would be able to pass through it with no resistance.[85]

Public key infrastructure

From Wikipedia, the free encyclopedia https://en.wikipedia.org/wiki/Pub...