![]() | |||||||||
|
![]() Editor Blog
RussianEnglish
By year:
Dark matter from dark energy in q-theory
The dynamics of the quantum vacuum is one of the major unsolved problems of relativistic quantum field theory and cosmology. The reason is that relativistic quantum field theory and general relativity describe processes well below the Planck energy scale, while the deep ultraviolet quantum vacuum at or above the Planck energy scale remains unknown. Following the condensed matter experience we develop a special macroscopic approach called q-theory, which incorporates the ultraviolet degrees of freedom of the quantum vacuum into an effective theory and allows us to study the dynamics of the quantum vacuum and its influence on the evolution of the Universe. The vacuum in our approach is considered as the Lorentz-invariant analog of a condensed-matter system (liquid or solid) which is stable in free space. The variable q is the Lorentz-invariant analog of the particle number density, whose conservation regulates the thermodynamics and dynamics of many-body systems. This approach is universal in the sense that the same results are obtained using different formulations of the q-field. In the paper, we choose the q-field in terms of a 4-form field strength, which has, in particular, been used by Hawking for discussion of the main cosmological constant problem -- why is the observed value of the cosmological constant many orders of magnitude smaller than follows from naive estimates of the vacuum energy as the energy of zero-point motion. In q-theory, the huge zero-point energy is naturally cancelled by the microscopic (trans-Planckian) degrees of freedom, as follows from the Gibbs-Duhem identity, which is applicable to any equilibrium ground state including the one of the physical vacuum. In the paper, we consider a further extension of q-theory. We demonstrate that, in an expanding Universe, the variable effectively splits into two components. The smooth part of the relaxing vacuum field is responsible for dark energy, while the rapidly oscillating component behaves as cold dark matter. In this way, q-theory provides a combined solution to the missing-mass problem and the cosmological constant problem. If this scenario is correct, the implication would be that direct searches for dark-matter particles remain unsuccessful in the foreseeable future. F.R. Klinkhamer and G.E. Volovik, JETP Letters 105, issue 2 (2017)
Created by I. Podyniglazova, 2016-12-20 11:53:02
NEW METHOD OF INVESTIGATIONS
The ability to detect nonequilibrium spin accumulation (imbalance) by all electrical means is one of the key ingredients in spintronics . Transport detection typically relies on a nonlocal measurement of a contact potential difference induced by the spin imbalance by means of ferromagnetic contacts or spin resolving detectors . A drawback of these approaches lies in a difficulty to extract the absolute value of the spin imbalance without an independent calibration. An alternative concept of a spin-to-charge conversion via nonequilibrium shot noise was introduced and investigated in experiment recently . Here, the basic idea is that a nonequilibrium spin imbalance generates spontaneous current fluctuations, even in the absence of a net electric current. Being a primary approach , the shot noise based detection is potentially suitable for the absolute measurement of the spin imbalance. In addition, the noise measurement can be used for a local non-invasive sensing. In this letter, we calculate the impact of a spin relaxation on the spin imbalance generated shot noise in the absence of inelastic processes. We find that the spin relaxation increases the noise up to a factor of two, depending on the ratio of the conductor length and the spin relaxation length. The design of the system. A diffusive normal wire of the length L is attached to normal islands on both ends. Nonequilibrium energy distribution on the left hand side of the wire generates the shot noise at a zero net current. The spin imbalance on the left-hand side of the wire is due to the electric current flowing from one ferromagnetic lead (red) to another one with opposite magnetization (blue).
V.S. Khrapai and K.E. Nagaev JETP Letters 105, №1 (2017)
Created by I. Podyniglazova, 2016-12-16 11:08:02
On the article V.S.Dryuma "On the analytical solution of the two-dimensional Korteweg-de Vries equation", Sov. Phys. JETP Lett. 19, 753-757 (1974)
Dryuma V.S. Institute of Mathematics and Informatics AS of Moldova valdryum@gmail.com
\begin{equation}
known as the Kadomtsev-Petviashvili equation (KP), which describe propagation waves in various problems of the plasma physics and hydrodynamics. By the author was first shown that possibilities of
Created by I. Podyniglazova, 2016-08-12 16:27:02
Paper "Derivation of exact spectra of the Schrodinger equation by means of supersymmetry" (L.E. Gendendtein, JETP lett., 38:6 (1983)
V.E. Adler, Landau Institute for Theoretical Physics
The paper is devoted to applications of the following remarkable transformation. An easy computation proves that if $\psi(x,\lambda)$ is a general solution of the equation
[1] G. Darboux. Sur une proposition relative aux equations lineaires. Compt. Rend. Acad. Sci. 94 (1882) 1456{1459. [arXiv:physics/9908003]
Created by I. Podyniglazova, 2016-06-24 16:23:02
On the paper "Dynamical Symmetry Restoration and Constraints on Masses and Coupling Constants in Gauge Theories" by A. D. Linde JETP Lett. 23 (1976) 64-67
I. V. Polyubin
Landau Institute for Theoretical Physics,
In 1973 S. Coleman and E. Weinberg investigated the effect of radiative corrections on the possibility of spontaneous symmetry breaking in seminal paper [1]. A. D. Linde studied some
$V=\lambda(\varphi^*\varphi)^2-{\mu}^2{\varphi}^*{\varphi}$. Renormalization conditions for effective potential fixed vacuum expectation value and mass of excitataions to its classical values. $$\lambda>\frac {3}{32{\pi}^2}g^4$$
Even if classical value of $\lambda$ is very small, due to one-loop correction it becomes equal to $\lambda_{eff}=\lambda+\frac{1}{2{\pi}^2}g^4$.
[1] S. R. Coleman, E. J. Weinberg, Phys. Rev. D7 (1973), 1888-1910
Created by I. Podyniglazova, 2016-04-28 16:33:02
The paper "Superfluidity in system with fermion condensate" (V. A. Khodel and V. R. Shaginyan, (1990))
V. A. Khodel McDonnell Center for the Space Sciences & Department of Physics, Washington University, St.Louis, MO 63130, USA V. R. Shaginyan National Research Center "Kurchatov Institute", Petersburg Nuclear Physics Institute, Gatchina, 188300, Russia
In the paper submitted in April of 1990, a new type of phase transition is introduced [1]. This phase transition occurs as the effective mass of quasiparticles diverges, $M^*\to\infty$, that is the fermions get very heavy. Beyond the point of the phase transition a fermion condensate (FC) arises in new phase: The energy $\varepsilon(p)$ of quasiparticles with momenta $p_i<p<p_f$ are equal to the chemical potential $\mu$, becoming flat or dispersionless, with $p_i<p_F<p_f$ and $p_F$ is the Fermi momentum [1]. Thus, the foundation of a new theory of strongly correlated Fermi-systems has been laid and the FC theory has emerged, while the Landau theory ceases to operate in that case [1]. The main novelty is that the quasiparticle distribution $n(p)$ in the region $p_i<p<p_f$ does
As a result, the FC theory has proposed explanations of numerous important experimental facts and paved the way for much subsequent research in the condensed matter physics of strongly correlated Fermi-systems represented by heavy-fermion (HF) metals and high-temperature superconductors [4, 8, 9], quantum spin liquids [10], quasicrystals [11], and two-dimensional Fermi-systems [12, 13].
Created by I. Podyniglazova, 2016-03-25 14:54:02
On the paper "A POSSIBLE MECHANISM FOR THE $\Delta I = 1/2$ RULE IN NONLEPTONIC DECAYS OF STRANGE PARTICLES." by A.I. Vainshtein, V.I. Zakharov and M.A. Shifman (1975)
M.A. Shifman School of Physics and Astronomy, University of Minnesota, Minneapolis, MN
55455, USA Minneapolis, MN 55455, USA
The penguin mechanism or penguin graphs present a class of Feynman diagrams which turned out of paramount importance for understanding weak flavor changing decays. While in the original paper in which they were introduced [1] only strange particle decays were considered, the idea was generalized to a large range of other applications, for instance, CP violating processes in B-meson decays, electroweak penguins, and many others. (The term penguin graphs was suggested by John Ellis later, see below.) The first developments were reported by the authors in [2]. The B-physics “penguins” are sensitive to new physics, especially those rare decays that are dominated by “penguin diagrams.” They were directly observed in 1991 and 1994 by the CLEO collaboration. “Penguins started to fly” [3] in high-energy physics in 1974. It was the very beginning of the quantum chromodynamics (QCD) era. We started to study interplay between the weak and strong interactions in 1973. The most perplexing issue in the strange particle decays was a seemingly unexplainable enhancement of the ∆I = 1/2 amplitudes in the kaon and baryon decays compared to the ∆I = 3/2 amplitudes. This mystery known under the name of “∆I = 1/2 rule” had been discussed by J. Schwinger as early as in 1964. The advent of QCD gave a new hope for explanation of this phenomenon. The publication [4] of Gaillard and Lee, which reached ITEP in a preprint form, provided us with a stimulating impetus. These authors noted that gluons act differently on the ∆I = 1/2 and ∆I = 3/2 amplitudes, with a tendency to enhance the former and suppress the latter [5]. This tendency was not enough, however, to explain the effect. When we discovered the penguin graphs and realized that they can appear only in the ∆I = 1/2 part of the strange quark decay amplitudes we understood that we were on the right track. Special chiral features inherent only to penguins led to the successful completion of the ∆I = 1/2 rule derivation (for a review see [3]). This result was reported in summer of 1974 at an international conference. In four decades that elapsed the march of the penguin diagrams in various flavor-changing processes was quite remarkable. Certainly, we could not anticipate this when we submitted the first paper to JETP Letters. At that time the newly discovered (SVZ) diagrams did not look like penguins. They were distorted to look penguin-like by John Ellis.
The SVZ graph in John Ellis’ “penguin” rendition. Here is John’s story which travels from one website to another (originally published in [6]): “Mary K. [Gaillard], Dimitri [Nanopoulos] and I first got interested in what are now called penguin diagrams while we were studying CP violation in the Standard Model in 1976. The penguin name came in 1977, as follows. In the spring of 1977, Mike Chanowitz, Mary K. and I wrote a paper on GUTs predicting the b quark mass before it was found. When it was found a few weeks later, Mary K, Dimitri, Serge Rudaz and I immediately started working on its phenomenology. That summer, there was a student at CERN, Melissa Franklin who is now an experimentalist at Harvard. One evening, she, I and Serge went to a pub, and she and I started a game of darts. We made a bet that if I lost I had to put the word penguin into my next paper. She actually left the darts game before the end, and was replaced by Serge, who beat me. Nevertheless, I felt obligated to carry out the conditions of the bet. For some time, it was not clear to me how to get the word into this b quark paper that we were writing at the time. Then, one evening, after working at CERN, I stopped on my way back to my apartment to visit some friends living in Meyrin, where I smoked some illegal substance. Later, when I got back to my apartment and continued working on our paper, I had a sudden flash that the famous diagrams look like penguins. So we put the name into our paper, and the rest, as they say, is history.” References [1] A.I. Vainshtein, V.I. Zakharov, M.A. Shifman, Pisma ZhETF 22, 123 (1975) [JETP Lett. 22, 55 (1975)]. [2] M. A. Shifman, A. I. Vainshtein and V. I. Zakharov, Nucl. Phys. B120 316 (1977) Non-leptonic decays of K-mesons and hyperons. ZhETF 72 1275 (1977) [JETP 45 670 (1977)]. [3] A. I. Vainshtein, Int. J. Mod. Phys. A 14, 4705 (1999) [hep-ph/9906263]. [4] M.K. Gaillard and B. Lee, Phys. Rev. Lett. 33, 108 (1974). [5] Mary K. Gaillard, A Singularly Unfeminine Professions, (World Scientific, Singapore, 2015), p. 59. [6] M. Shifman, ITEP Lectures in Particle Physics, (World Scientific, Singapore, 1999), Vol. 1, p. x [hep-ph/9510397].
Created by Alexander Prokofiev, 2016-02-18 16:15:02
On the paper "GLUON CONDENSATE AND LEPTON DECAYS OF VECTOR MESONS" by A.I. Vainshtein, V.I. Zakharov and M.A. Shifman (1978)
M.A. Shifman School of Physics and Astronomy, University of Minnesota, Minneapolis, MN
55455, USA Minneapolis, MN 55455, USA The paper [1] laid the foundation for the so-called SVZ sum ruled method (sometimes also referred to as QCD sum rules) allowing one to calculate properties of a huge variety of low-lying hadronic states basing on average characteristics of the QCD vacuum, such as quark and gluon condensates (the latter was first introduced in [1]). The main conceptual component of the method is the Wilson operator product expansion (OPE). The focus of Wilsons original work was on statistical physics, where the program is also known as the block-spin approach. Surprisingly, in high-energy physics of the early-to-mid 1970s the framework of OPE was narrowed down to perturbation theory. The authors of [1] and the subsequent publications [2] were the first to adapt the general Wilsonian construction to QCD. Their goal was to systematically include power-suppressed effects (condensate corrections), thus bridging the gap between short and large distance dynamics. This “bridging” did not lose its significance till this day. This route – matching between the short distance expansion and long distance (hadronic) representation – led to remarkable successes. The SVZ method was tested, and proved to be fruitful in analyzing practically every static property of all established low-lying hadronic states, both mesons and baryons. (For a review see [3].) Later, the very same OPE and the same ideas that were developed in [1, 2], was applied with triumph in heavy quark expansions blossomed in the 1990s in the framework of the heavy quark theory (reviewed in [4]). The question we asked ourselves in 1977, which eventually led to the idea of deriving properties of matter from the vacuum condensates, was as follows (see also [5]). What if we start from short distances, where the quark-gluon dynamics was under theoretical control, and extrapolate to larger distances using general features of QCD? What maximal information on hadronic properties could we obtain? Surprisingly, we started getting interesting results for charmonium almost immediately. A certain success came, however, after V. Novikov, L. Okun, and M. Voloshin joined us. It turned out that a whole variety of the charmonium parameters could be (and were) predicted. A curious episode is worth mentioning. At this time, according to experiment, the only candidate for the psuedoscalar charmonim was the so-called X(2.83). Its mass seemed to be too low for this interpretation; still, with some tension it could fit potential model predictions which were widespread at that time. In [6] it was unambiguously shown from the analysis of the QCD sum rules X(2.83) could not be the psuedoscalar charmonim. The mass of the latter should have lied at 3.01±0.01 GeV. Later the experimental data regarding X(2.83) were indeed retracted. The current data give 2.983±0.0007 GeV. For about a year, we played the game of getting cc coupling constants and masses from simple numbers. In 1977, we submitted a review report [7]. At about that time, it became clear that our success was limited; and could not be generalized to the most common light-quark mesons and baryons without new ideas. It was a hot summer, just before vacation. Our big collaboration ceased to exist. We were leisurely discussing something when the first hints appeared. The conjecture was that the vacuum is actually something like a gluon medium, and all particle properties are due to the quark interaction with this medium which can be conveniently parametrized by certain quark and gluon condensates. We worked out the first implications of the gluon condensate in fall 1977. At first, we were discouraged by a wrong sign for one of the most important particles (ρ meson). Then we suddenly understood that this sign could be compensated by the four-quark condensate – a real breakthrough. The accuracy of our predictions turned out to be much higher than anyone could expect a priori. Inspired, we worked at a feverish pace for the whole academic year. When the final paper was ready (a short one had been sent to JETP Letters in the beginning [1]), it contained more than 300 typewritten pages. We could not make a preprint out of it because, according to Soviet bureaucratic rules, preprints could have no more than 40 pages or so. So, we divided it artificially into seven or eight parts, trying to do it in such a way that it would not be immediately obvious to the censor. It appeared as three papers occupying the whole issue of Nuclear Physics B. References [1] A.I. Vainshtein, V.I. Zakharov and M.A. Shifman, Pis'ma ZhETF 27 , 60 (1978) [JETP Lett. 27, 55 (1978)] . [2] M. A. Shifman, A. I. Vainshtein and V. I. Zakharov, Nucl. Phys. B 147, 385 (1979), and Nucl. Phys. B 147, 448 (1979). [3] M. A. Shifman, Prog. Theor. Phys. Suppl. 131, 1 (1998) [hep-ph/9802214]. [4] I. I. Y. Bigi, M. A. Shifman and N. Uraltsev, Ann. Rev. Nucl. Part. Sci. 47, 591 (1997) [hep-ph/9703290]. [5] M. Shifman, Current Contents, 32, 9 (1992). [6] V. A. Novikov et al., Phys. Rept. 41, 1 (1978).
Created by I. Podyniglazova, 2016-01-21 17:28:02
|